Swart 2015
Swart 2015
Peter K. Swart,
pswart@rsmas.miami.edu
ABSTRACT
Stable carbon and oxygen isotopes (18O and 13C values) and trace elements have been
applied to the study of diagenesis of carbonate rocks for over 50 years. As valuable as these
insights have been, many problems regarding the interpretation of geochemical signals within
mature rocks remain. For example, while the 18O values of carbonate rocks are dependent
both upon the temperature and the 18O value of the fluid and additional information,
including trace element composition aids in interpreting such signals, direct evidence of either
the temperature or the composition of the fluids is required. Such information can be obtained
This is an Accepted Article that has been peer-reviewed and approved for publication in the
Sedimentology, but has yet to undergo copy-editing and proof correction. Please cite this article
as an “Accepted Article”; doi: 10.1111/sed.12205
method such as the ‘clumped’ isotope technique. Such data speak directly to a large number of
problems in interpreting the oxygen isotope record including the well-known tendency for 18O
values of carbonate rocks to decrease with increasing age. Unlike the 18O, 13C values of
carbonates are considered to be less influenced by diagenesis and more a reflection of primary
changes in the global carbon cycle through time. However, many studies have not sufficiently
emphasized the effects of diagenesis and other post-depositional influences on the eventual
carbon isotopic composition of the rock with the classic paradigm that the present is the key to
the past being frequently ignored. Finally, many additional proxies are poised to contribute to
exciting point with an explosion of new proxies and methods, care should be taken to
understand both old and new proxies before applying them to the ancient record.
INTRODUCTION
Although geochemical changes during the diagenesis of carbonate sediments and rocks have
been recognised for a considerable period of time, the development and refinement of the
techniques for the analysis of stable isotopes of oxygen (18O and 16O) and carbon (13C and 12C)
(18O and 13C values) in carbonates (Epstein et al., 1951; Epstein et al., 1953; McCrea, 1950)
have provided insights into the use of isotopes as geochemical tracers in the interpretation of
alteration (diagenesis) of carbonate sediments. Early papers (Gross, 1964; Gross & Tracey,
1966; Hudson, 1977; Land, 1967; Matthews, 1968) provided the foundation for subsequent
work which refined and explored new interpretations and applications. In addition, the
combination of trace elements (Banner, 1995; Brand & Veizer, 1980; Brand & Veizer, 1981;
Kinsman, 1973) and stable C and O isotopes to unravel diagenetic histories in carbonate rocks.
There have been several important and valuable texts dealing with this topic (Bathurst, 1971;
Berner, 1980; Milliman, 1974; Moore, 1989; Morse & Mackenzie, 1990) as well as numerous
review papers (Budd, 1997; Drever, 1982; Emerson & Hedges, 2003; Holland, 2007; Lerman &
Clauer, 2007; Martin & Sayles, 2003; Morse, 2003; Veizer & Mackenzie, 2003) and others.
Carbonates are considered to include any mineral with a structural CO32- group. Although many
metals form carbonate minerals, the most common carbonates are those formed by Ca, Mg,
and a combination of these two elements. The basic distribution of carbon and oxygen
isotopes in nature is well covered in numerous publications and text books (Friedman & O'Neil,
1977; Hoefs, 1980; Sharp, 2007; Zeebe & Wolf-Gladrow, 2001) and will not be considered here
in detail. However, a brief introduction has been included in order to guide the reader to
Oxygen
The 18O value of carbonate minerals is controlled by: (i) the temperature of formation (Epstein
et al., 1951; Friedman and O'Neil, 1977; O'Neil et al., 1969; Tremaine et al., 2011; Urey, 1947);
(ii) the 18O value (18Ow) of the precipitating fluid (Epstein et al., 1953; Epstein & Mayeda,
1953; Urey, 1947); (iii) mineralogy (Emrich et al., 1970; Tarutani et al., 1969); (iv) the pH of the
Temperature
Since the paper of Urey (1947), the focus of the 18O technique has been mainly as a
temperature and for those carbonates forming at sedimentary temperatures the relationship is
Fluids
The 18Ow is controlled by the origin of rainfall, the amount of evaporation (Craig & Gordon,
1965; Craig et al., 1963; Gonfiantini, 1986), mixing with seawater and mineral–water reactions
(Lawrence, 1989) . Typically meteoric waters have 18Ow values less than 0‰ and these
become more negative towards polar regions, with increasing altitude, distance from the
evaporation source and/or decreasing temperature (Rozanski et al., 1993; Fig. 1).
Compared to meteoric fluids marine waters have more positive 18Ow values and, although
during evaporation water bodies becomes more enriched in 18O, in practice there is a limit to
the extent to which 18Ow values can increase because the 18O and 16O in the water vapour in
the atmosphere isotopically exchanges with the evaporating water pool (Craig & Gordon, 1965).
In addition, the 18Ow value can become altered by changes in the thermodynamic activity of
water, as a consequence of the hydration of ions in solution, and from isotope fractionation
during the crystallization of salts (Gonfiantini, 1986). As a result, during the final stages of
evaporation the 18Ow value may reach a maximum value and then become more negative as
relative humidity of the atmosphere, the temperature of evaporation, the 18O of the water
Mineralogy
Inorganically formed aragonite (Grossman, 1984) and dolomite (Fritz & Smith, 1970; Matthews
& Katz, 1977; Northrop & Clayton, 1966; O'Neil et al., 1969; O'Neil & Epstein, 1966 ; Sheppard &
Schwarcz, 1970; Vasconcelos et al., 2005) have different equations compared to low Mg calcite
that relate temperature to the 18O value of the mineral. The 18O of aragonite, for example, is
typically about 1‰ more positive than that of low Mg calcite (LMC) formed at the same
temperature, while the 18O value of high Mg calcite (HMC) increases by about 0.06‰ for
every mol% MgCO3 (Tarutani et al., 1969). Interestingly, the 0.06‰ difference corresponds to
the theoretical and expected difference of 3‰ between the 18O values of calcite and dolomite
Kinetics and pH
The distribution of the inorganic oxygen bearing carbon species in solution is controlled by pH.
At high pH the speciation is dominated by CO32- changing to HCO3- and H2CO3 as the pH is
lowered. Because there are well-known isotopic fractionations for both C and O between these
species (Beck et al., 2005; Emrich et al., 1970; Zeebe & Wolf-Gladrow, 2001), the 18O value of
the sum of these species changes as a function of pH assuming that pH is the only factor driving
the speciation (Zeebe, 2005a; Zeebe, 2007; Zeebe & Wolf-Gladrow, 2001). At low pH the major
species, H2CO3, is about 41 ‰ more positive than the oxygen in the water at 25oC (Bottinga &
Craig, 1969). As pH increases and HCO3- and CO32- become the dominant species, the 18O
value of carbonate precipitated reflects the 18O value of the sum of the carbonate species and
that equilibrium is not maintained between the various species (Zeebe, 2005a), then at low pH
it would be expected that carbonates would have more positive 18O values than those
Kinetic controls are considered together with pH because the rate of precipitation is often
related to the amount of CO32- in a solution which in turn is controlled by the pH. Generally
faster rates of precipitation will occur at higher pH and favour the incorporation of the light
isotopes, 16O and 12C relative to 18O and 13C. Kinetic and pH controls are particularly important
in the precipitation of biogenic carbonates, but also play a role during the precipitation of some
inorganic carbonates (i.e. speleothems; Affek et al., 2008b; Tremaine et al., 2011).
Carbon
The 13C values in carbonates are controlled by the: (i) 13C values of ambient dissolved
inorganic carbonate (DIC; Mook, 1968); (ii) pH of precipitation (Zeebe & Wolf-Gladrow, 2001);
(iii) rate of precipitation (kinetics; McConnaughey, 2003); (iv) mineralogy (Emrich et al., 1970;
Rubinson & Clayton, 1969); and (v) temperature (Deines et al., 1974; Emrich et al., 1970).
The major control on the 13C values of marine carbonates is exerted by the 13C value of
oceanic DIC; this in turn is determined by the fixation of CO2 during photosynthesis (Craig,
1953; Park & Epstein, 1961), the weathering of carbonate rocks (Berner et al., 1983) and the
oxidation of organic material (OM) within the marine environment (Weber & Woodhead,
utilized by the plant (C3, C4 or CAM.) Most photosynthetic terrestrial vegetation utilizes the C3
pathway and has a 13C value between -25‰ and -30‰, while marine vascular plants utilizing
the same pathway have significantly more positive values (-8 to -12‰; Deines, 1980). The
differences in the 13C values of aqueous and terrestrial vascular plants reflects the fact that the
majority of inorganic carbon in the aqueous realm is usually present as HCO3-, and is ca 8‰
enriched in 13C compared to atmospheric CO2 (Vogel et al., 1970). In contrast to marine
vascular plants, marine algae have much more negative 13C values (-20 to -24‰; Degens et al.,
1968; Sackett et al., 1965). Variations in the 13C values of DIC are relatively small in the open
ocean, but become increasingly important in coastal marine environments, such as coral reefs,
isolated embayments and shallow carbonate shelves, where the residence time of the water is
longer (Weber & Woodhead, 1971). The 13C values of DIC in these environments can
therefore be influenced by the addition of isotopically negative CO2 from respiration as well as
Kinetics and pH
The pH and kinetic control of carbon in water operates in a similar manner to that of oxygen
except that the pool of carbon is much smaller. As a result, the range of 13C values over which
the different species vary is restricted compared to that of 18O. Consequently, between pH
values of ca 4 to 10 there is a strong covariance between 13C and 18O values in carbonates
affected by changes in pH (Zeebe & Wolf-Gladrow, 2001). Above pH values of ca 10 the 13C
value of the carbon species no longer varies, while the 18O value of the sum of the carbonate
The 13C value of aragonite is typically 1 to 2‰ more positive than co-occurring LMC and HMC
(Emrich et al., 1970; Romanek et al., 1992; Rubinson & Clayton, 1969). Dolomites may be ca
1‰ more positive than LMC formed under similar conditions (Sheppard & Schwarcz, 1970).
Temperature
In contrast to 18O, the 13C value of carbonates is not thought to be directly influenced by
temperature. However, it is clear that the relationships between the various inorganic bearing
species are temperature-dependent (Deines et al., 1974; Emrich et al., 1970). Therefore, in
spite of experimental data which suggests that the 13C values of carbonates is independent of
temperature (Romanek et al., 1992; Rubinson & Clayton, 1969), theoretically the 13C value of
carbonate should change. Part of the problem may be the limited number of studies measuring
the fractionation factors between various species and the inconsistent values reported. For
example, in the study of Deines et al. (1974), the fractionation between calcite and HCO3-
decreased with decreasing temperature (contrary to the expected pattern), while the data
presented by Emrich et al. (1970) were based on carbonates of mixed aragonite and calcite
mineralogy. Regardless of whether the 13C values of inorganic precipitates are temperature
temperature will play an important role in governing the 13C value of DIC and therefore
Clumped Isotopes
The major problem with using 18O values in minerals such as carbonates is that the values
depend both upon temperature and the 18Ow value (Fig. 3). This dilemma is potentially
independently of the 18Ow. This method has already been successfully applied to a wide range
of previously used palaeotemperature proxies (Affek et al., 2008a; Affek et al., 2008b; Eagle et
al., 2011; Ghosh et al., 2007a; Ghosh et al., 2007b; Huntington et al., 2011; Peters et al., 2012;
Snell et al., 2007; Thiagarajan et al., 2011; Tripati et al., 2010; Wacker et al., 2014). Although
the application of the clumped isotope technique to diagenetic studies is still in the early
stages, several papers (Bristow et al., 2011; Budd et al., 2013; Dennis & Schrag, 2010;
Fernandez et al., 2014; Ferry et al., 2011; Loyd et al., 2012; Loyd et al., 2013; Sena et al., 2014;
Van De Velde et al., 2013) have suggested that this could be a promising approach. However,
there have been no studies examining the behaviour of clumped isotopes in well-constrained
diagenetic settings. The original calibration (Ghosh et al., 2006), applied to a range of different
carbonates, has been augmented by a theoretical calibration (Guo et al., 2009) and other
empirical calibrations (Dennis & Schrag, 2010). The slight differences in slopes and intercepts
between these are sufficient to cause significant error in the calculated temperature. The
promise of the clumped isotope method is that once the temperature has been calculated, the
18Ow value of the formation or diagenetic fluid can be estimated using known relationships
between the fluid and mineral. Although the uncertainty in the various equations relating 47
to temperature is still significant, reasonable estimates can be made using the equations
Solid-state Resetting
The clumped isotope method also allows investigation of the phenomenon of isotopic resetting
its ratio of 13C–18O bonds at higher temperature and, if so, at what temperature and at what
rate? Alternatively, do minerals formed at high temperatures continually reorganize their 13C–
18
O bonds until a ratio is locked in at lower temperatures? Such a phenomenon would be
important because it would constrain the burial history of a carbonate and perhaps the original
depositional temperatures could be back calculated assuming that there was information on
the rate of change at specific temperatures and the burial temperature. If solid diffusion was
responsible for these changes, then the 18Ow calculated using such temperatures would in fact
be erroneous unless the temperatures were corrected for the burial history. Studies
investigating this phenomenon indicate that such resetting does in fact take place at
geologically significant rates, particularly at temperatures higher than 300oC (Henkes et al.,
During periods of enhanced preservation and burial of organic matter (OM) the 13C value of
the DIC increases. During the Early Carboniferous for example, when abundant coal deposits
were formed, the 13C value of the DIC in the oceans, as recorded by carbonate minerals,
increased by 3 to 4‰ (Saltzman et al., 2004; Veizer et al., 1999). Conversely, during times of
enhanced oxidation of OM, the 13C value of the DIC decreases. Over the past 50 Myr, the 13C
values of pelagic carbonates have decreased by 2 to 3‰ (Shackleton, 1985; Tipple et al., 2010)
reflecting the transfer of organic carbon into the ocean-atmosphere system. Enhanced
oxidation can also occur during periods of low sea level, when the continental shelves are
continental margin, thereby releasing large volumes of CH4 with low 13C values into the
atmosphere. Such CH4 ultimately oxidizes to CO2, causing the atmosphere and the oceanic DIC
to return to more negative values. Similar changes can also be associated with extreme events
such as rapid warming, extra-terrestrial impacts, or man-made events such as the burning of
fossil fuel.
Marine Cements
Inorganically precipitated carbonates which form in equilibrium with their environment possess
18O and 13C values which agree with those calculated from theoretically and/or empirically
derived equations. Because different types of carbonate (HMC, low-Mg LMC, dolomite and
aragonite) have distinctive equations which describe equilibrium, these differences will be
evident in the 18O and 13C values. Differences in the 18O and 13C values of carbonate with
the aragonites which form in the cavities of steep marine slopes in the Bahamas (Gonzalez &
Lohmann, 1985; Grammer et al., 1993) or HMC cements which are found within coral reefs in
the Pacific (Aissaoui, 1988; Gonzalez & Lohmann, 1985). The 13C values of these cements are
slightly more negative than those of inorganic aragonite precipitated on margins of the Great
Bahama Bank (GBB), perhaps as a result of microbial processes occurring within the interstices
Biogenic Carbonates
Biogenically produced carbonates can either be precipitated in isotopic equilibrium with their
environments or can exhibit so-called ‘vital effects’ in which the 18O and 13C values are
18O and 13C values in carbonate secreting organisms arise as a result of differences in the
calcification (Adkins et al., 2003; McConnaughey, 1989; McConnaughey, 2003). The 18O and
13C values of such skeletal material have been reported in numerous publications and can vary
widely (Fig. 6). As an example, aragonite from photosymbiont-bearing, shallow water coral
skeletons typically have 18O values between -5‰ and -4‰ and 13C values between -1‰ and
+1‰ (Swart, 1983) . Such 18O values are too negative to be indicative of equilibrium in the
environments in which coral reefs are normally found. The 13C values of the skeletons are also
too variable when compared to the relatively constant 13C values of the ambient DIC. In
contrast to shallow-water corals, deep-water varieties, which lack photosymbionts, have 18O
values between -8‰ and -2‰ and 13C values between -8‰ and +1‰ (Adkins et al., 2003;
Emiliani et al., 1978; Land et al., 1977; Fig. 6). These wide ranges in the deep-water corals
cannot be caused by changes in the 18O and 13C values of the water or the temperature of the
environment and therefore they must be produced by changes in the calcification process, such
as calcification rate (McConnaughey, 2003), pH (Adkins et al., 2003), or both. In contrast, the
photosymbionts in the shallow-water corals modify the calcification environment, limiting wide
variations in pH during calcification and therefore have more restrictive 18O and 13C ranges
compared to non-photosymbiont bearing corals. Insight into the behaviour of the 18O and
13C values of the non-symbiont bearing corals can be obtained by comparing data with the
relating to the calcification processes (Figs 5 and 6). Green algae tend to have more positive
13C values, although different portions of the algae can have different compositions. For
example the head of algae such as Penicillus sp. and Rhipocephalus sp. tend to have more
positive 18O and 13C values compared to the stalk. In Halimeda sp. the 13C values of the
newly formed algal segments can be up to 6‰ heavier than the basal portion (Wefer, 1981). A
similar enrichment was observed in Penicillus sp. by these authors. Although the authors
offered no explanation for the observed pattern in 13C values, it is likely that it results from
secondary calcification within the algal segments as they age. In contrast calcareous red algae
are much more depleted in 13C than green algae and can be considered to form out of isotopic
Most modern carbonates, whether they are considered non-biogenic (ooids and peloids) or
Muscatine et al., 2005; Wainwright, 1964). Such OM can be produced by the organism itself as
a result of the calcification process, or be secondary as a result of the action of algae, fungi,
bacteria, sponges and other boring organisms. Such secondary OM contributes significantly to
the total organic content of many skeletons with, for example, boring organisms observed in
intimate contact with recently formed coral skeletons (Golubic, 1969; Lukas, 1974). In fact it is
probable that the OM originating from secondary contributors is greater than that supplied by
the host organism. The pervasiveness of such material can be observed readily through the use
of thin sections and scanning electron microscopy (SEM) in live and recently dead skeletons
subject to degradation causing lower levels of oxygen and possible sulphate reduction. The
importance of such material is that its 15N and 13C value has been used as a proxy for a
variety of different environmental parameters (Marion et al., 2006; Wang et al., 2014). Non-
skeletal materials such as ooids and peloids contain higher amounts of OM mainly derived from
endolithic organisms. Processes such as sulphate reduction and denitrification have been
identified in these particles using genetic markers (Diaz et al., 2014) and confirmed in the
geochemical signatures of pore water leachates derived from these particles (Diaz et al., 2013).
Carbonate Platforms
different carbonate environments tend to have characteristic 18O and 13C values. For
example, within a reef, the 18O and 13C values of the sediments change between different
sub-environments (i.e. reef flat, lagoon, reef crest and fore reef; Weber & Woodhead, 1969)
by non-skeletal grains (peloids, ooids, etc.), such as the Bahamas, tend to have more positive
18O and 13C values (18O = ca 0‰ and 13C = ca +3 to +5‰) and appear to be in isotopic
equilibrium (Lowenstam & Epstein, 1957; Shinn et al., 1989; Swart et al., 2009) with ambient
waters (Gischler et al., 2009; Weber & Woodhead, 1969; Fig. 5). For comparison, other skeletal
and non-skeletal components are shown in Fig. 6. The nature of the basin (closed or open) and
its connection to the open marine environment also affects the 18O and 13C values of the
waters and hence the values of carbonates precipitated from those waters (Patterson & Walter,
(Lloyd, 1964; Swart & Price, 2002) as a result of evaporation. The 13C values, however, will be
lowered following input of waters containing organic carbon or the products of the oxidation of
organic carbon from adjacent terrestrial areas, which are depleted in 13C (Halley & Roulier,
1999). A similar scenario prevails in ramp settings, both in modern and ancient carbonates
Freshwater Carbonates
carbonates includes travertines (generally used for carbonates deposited in higher temperature
regimes related to hydrothermal activity), tufas (lower temperature deposits formed in lakes,
springs and waterfalls) and speleothems (precipitates formed in cave systems; Capezzuoli et al.,
2014; Ford & Pedley, 1996; Fouke, 2011; Sanders & Friedman, 1967). In contrast to marine
settings, freshwater carbonates, whether organic or inorganic in nature, generally have more
negative 18O and 13C values. The 18O values tend to be related to water temperature and
the values of the local meteoric fluids, with higher latitudes and altitudes having more negative
18Ow values that closely follow those of global precipitation patterns. The only exception to
the more negative patterns in 18Ow values would be terminal lakes such as Pyramid Lake
(Benson et al., 2013) and marsh areas such as the Everglades (Meyers et al., 1993; Price &
Swart, 2006) where the principal water loss is through evaporation (Gonfiantini, 1986; Leng &
Marshall, 2004; Yuan et al., 2011). Evaporation would produce freshwater carbonates with
more positive 18Ow values than expected considering the low salinity of the water bodies.
Such reverse patterns in the 18Ow values are observed in the transition between freshwater
reflected in elevated 18O values in carbonates found in the region (Halley & Roulier, 1999;
Lloyd, 1964). This is opposite to the expected trend if a major river such as the Mississippi with
negative 18Ow values flowed into a marine estuary with more positive 18Ow values. In the
ancient a similar interpretation to that used for the Everglades has been placed on the 18O
values in Jurassic molluscs (Hendry & Kalin, 1997) where specimens with the most negative
18O values were found furthest from shore. The 13C values of lake carbonates are more
negative than marine carbonates, a phenomenon linked to the quantity of OM relative to the
size of the water body, its 13C value and subsequent oxidation. In addition, many tufas are
associated with microbial activity (Chafetz & Folk, 1984) which can influence the 13C value of
precipitated carbonate and often result in material with negative values, while travertines
forming from high temperature geothermal springs have more elevated values (Fouke, 2011;
Renaut et al., 2013; Turi, 1986). Despite the complexities, sedimentary records obtained from
such localities are sensitive to variations in water balance, input of OM and productivity. This
combined with the anoxic bottom conditions mean that sedimentary records frequently
Speleothems form as ground waters charged with CO2 seep into caves. Here the waters degas
CO2, forcing the precipitation of calcite. In instances where the ground waters have high Mg/Ca
ratios, aragonite can also form (Lambert & Aharon, 2011; Schwarcz, 1986). Because
speleothems have been used extensively for palaeoclimate interpretations, numerous reviews
exist covering the principals and caveats controlling the 13C and 18O values and trace element
The clumped isotope method has been applied to speleothems, but most studies show that
values are offset towards higher than expected temperatures (Affek et al., 2008a; Affek et al.,
2008b; Kluge & Affek, 2012). Because the formation of speleothems may be analogous to the
environments.
Hydrogen
Isotopes of H (1H and 2H) are useful in understanding carbonate diagenesis because H is
incorporated as water into fluid inclusions within carbonates and is available for comparison
with the oxygen in the same inclusion. Because H is not incorporated into the carbonate
structure to any great extent, the stable isotopes of H are not fractionated during carbonate
deposition and therefore the isotopic composition of H reflects the origin and evaporation
history of the fluid. Normally the 2H value of water, which has not been fractionated during
evaporative processes, has a defined relationship relative to the 18Ow value, namely the
meteoric water line (MWL; Craig & Gordon, 1965). Evaporated fluids deviate from the MWL
depending on the relative humidity in the ambient environment (Gonfiantini, 1986). Waters
evaporating in a humid environment fall closer to the MWL compared to those in more arid
localities. Because the magnitude of the fractionation effects is different for hydrogen and
oxygen, an evaporating solution can take different pathways depending upon the initial
As a result of the long residence time of B, the oceans have a fairly homogeneous 10B and 11B
ratio, with a 11B value of ca 39.5‰ (Foster et al., 2013). This value represents a balance
between the major input of B from the continents (-8 to -15‰; Ishikawa & Nakamura, 1993)
and removal by processes which act to increase the 11B value, such as adsorption onto clays,
exchanges with the oceanic crust and incorporation into carbonates. The incorporation of B
into modern carbonates, which have 11B values of +22±3‰ (Hemming & Hanson, 1992)
represents about 20% of the loss of B from the oceans (Vengosh et al., 1991). Within natural
waters B exists complexed as either B(OH)3 or B(OH4)- depending upon the pH of the solution.
Because there is significant isotopic fractionation of 11B between these species (Zeebe, 2005b)
and because it is generally believed that only the B(OH4)- ion is incorporated into the crystal
structure of carbonates, probably substituting for the CO32- ion, the 11B values of the
carbonates become more positive as the pH increases. The promise is that the 11B value of
unaltered carbonates will provide information on both recent and ancient oceanic pH. There
has been a significant amount of work to test this hypothesis concentrating on calibrating the
11B values of the skeletal carbonate of various organisms in situations where the pH of the
unclear how the 11B values of the skeletons of organisms such as corals, which have been
relates to pH variations in the external environment. There have been measurements of the
11B values of unaltered Mesozoic fossil material (Foster et al., 2010) in attempts to derive
Sulphur
Sulphur has three stable isotopes, 32S, 33S and 34S, and is mainly present in the ocean as
sulphate (SO42-), a species involved in the oxidation of OM in the absence of oxygen. The
significant fractionation of 34S that occurs during this process (sulphate reduction) results in 32S
being preferentially incorporated into H2S while the residual SO42- is enriched in 34S (Chambers
& Trudinger, 1979). The fractionation accompanying this step is between 40‰ and 70‰. In
the presence of Fe, pyrite or other related iron sulphide minerals are formed. The present 34S
of the oceanic sulphate is approximately +20‰ and because SO42- is incorporated into
evaporite minerals, such as gypsum and anhydrite with only a minimal amount of fractionation
(Holser & Kaplan, 1966; Lloyd, 1968), the 34S values of evaporites can be used to monitor the
changing 34S values of the oceans through time (see later discussion). An additional proxy of S
in the oceans is recorded in the 34S values of barite (BaSO4), a mineral formed throughout the
water column by a variety of different organisms (Griffith & Paytan, 2012; Paytan et al., 1998)
and also as a diagenetic mineral formed within the interstitial pore water.
Although the S cycling is intimately linked to the carbon cycle, SO42- is also included as a trace
(Lyons et al., 2005). Workers using this proxy have expended significant effort to make sure
that all traces of the S adhering to the exterior of the carbonate have been removed before
extracting the S and converting it into a form suitable for S-isotopic analysis (Gill et al., 2008;
therefore the34S value of this CAS is thought to reflect oceanic compositions, and
consequently has been used as an additional indicator of the 34S value of the oceans through
time (Gill et al., 2011; Paris et al., 2014). The abundance of 33S within CAS can also be measured
and its concentration has implications regarding the evolution of oxygen in the atmosphere
Magnesium
The behaviour of Mg isotopes (24Mg, 25Mg and 26Mg) during modern carbonate deposition has
been studied by Wombacher et al. (2011) who determined small, but significant, differences
between the fractionation exerted by certain aragonitic (-0.9 ± 0.2‰) and calcitic organisms
(-2.6 ± 0.3‰), with aragonite generally being less depleted than calcite. In contrast Mg-clays
are slightly enriched in 26Mg. Initial work on 26Mg reported a small temperature dependence in
both inorganic aragonite (Wang et al., 2013) and corals (Saenger et al., 2014). While limited in
extent, studies so far appear to show discrimination against the heavier isotopes during
diagenesis, so that diagenetic carbonates such as dolomites are isotopically more negative than
modern carbonates (Carder et al., 2005; Galy et al., 2002; Geske et al., 2015; Higgins & Schrag,
2010; Higgins & Schrag, 2012).The relative discrimination of Mg isotopes in carbonates and clay
minerals make them a powerful tool for studying the processes of geochemical cycling of Mg
Calcium has five stable isotopes (40Ca, 42Ca, 43Ca, 44Ca and 46Ca). The most common minor
isotope investigated is mass 44, reported either to mass 40 or mass 42 [44/40Ca or 44/42Ca;
mass 42 is used as a result of interferences at mass 40 from the argon used in the plasma on
the ICP–MS (inductively coupled plasma – mass spectrometer)]. The modern oceans appear to
sources and sinks (De La Rocha & DePaolo, 2000; Zhu & Macdougall, 1998) and variations in the
rate of cycling of calcium through sedimentary rocks (Holmden, 2009; Holmden et al., 2012)
producing secular variation of about 2‰ in the 44Ca value of unaltered fossil calcareous
organisms (De La Rocha & DePaolo, 2000; Farkaš et al., 2007). The 44Ca value has been
investigated in a range of different modern calcareous organisms. Some, such as corals, show a
small amount of enrichment (0.1 to 0.6‰; Pretet et al., 2013) while others, such as planktonic
foraminifera, show more negative values (ca -1.5‰; Griffith et al., 2008). It has been suggested
that the 44Ca value of such organisms is related to the temperature of calcification (Griffith et
al., 2008; Gussone et al., 2005; Heuser et al., 2005; Immenhauser et al., 2005), with similar
relationships being observed in calcite and aragonite. However, temperature dependence does
not appear to be ubiquitous (Sime et al., 2005) and some workers have claimed that the effects
are ‘insignificant’ when they do occur (Steuber & Buhl, 2006) . Diagenetic processes have been
noted by some workers (Steuber & Buhl, 2006), the most important of which is fractionation
during dissolution and precipitation reactions (Fantle & DePaolo, 2007). This allows the 44Ca
Although strontium has three stable isotopes, 86Sr, 87Sr and 88Sr, the most common
measurement made in carbonate materials is that of 87Sr relative to 86Sr (87Sr/86Sr). Because
carbonate organisms do not fractionate the 87Sr/86Sr when they form skeletons, the 87Sr/86Sr of
the oceans can be measured through time in unaltered carbonate material (Burke et al., 1982;
DePaolo & Ingram, 1985; McArthur et al., 2001). The temporal change in the Sr-isotope ratio of
the oceans reflects the erosion of continental crust which has relatively high 87Sr/86Sr ratios
(87Sr/86Sr > 0.7140) with the 87Sr being supplemented by the decay of 87Rb and increased cycling
of ocean water through oceanic ridges which have relatively low 87Sr/86Sr values (87Sr/86Sr=
0.7040; Brass, 1976). The steady increase in 87Sr/86Sr over the past 60 Myr for example has
been linked to Himalayan uplift that has delivered an increased flux of radiogenic Sr to the
oceans (Raymo & Ruddiman, 1992). Similar arguments have been used to explain the evolution
of the 87Sr/86Sr ratio of seawater during earlier periods in Earth history (Montanez et al., 1996).
Measurements have also been made of the 88Sr/86Sr ratio in carbonates. Because this ratio is
not altered by radioactive decay, its measurement provides an additional constraint for
understanding carbonate deposition and diagenesis, as well as the Sr cycle in the oceans
(Boehm et al., 2012; Krabbenhoft et al., 2010). The 88Sr/86Sr ratio is reported relative to NIST
SRM 987 using the conventional delta notation (88Sr), with the total range of reported values
less than 0.2‰. Carbonates measured to date have a 88Sr value of ca 0.2‰ compared to
Additional isotopic systems which are potentially available for future research into carbonate
diagenesis include elements such as Cr (Wille et al., 2013), Fe , Cu, Zn, (Conway et al., 2013), Mo
(Wille et al., 2013) and Hg (Kritee et al., 2009; Kwon et al., 2013). The isotopic systematics of
these elements have only recently been explored and are not therefore widely applied to
carbonate systems. For example, the isotopic systematics of Mo and Cr in non-carbonate facies
have helped to define the redox state during the Archean (Crowe et al., 2013; Wille et al., 2013)
Systems which have more than two stable isotopes can exhibit mass independent fractionation,
so that the behaviour of one of the minor isotopes is not directly predictable from the
measurement made on the second minor isotope. The classic example of mass independent
fractionation is that exhibited by the 18O, 17O and 16O system. On the Earth’s surface the
fractionation of 17O/16O can be related to that of 18O/16O through the terrestrial fractionation
line (TFL) which has a slope of 0.53 equating to the difference in the zero-point energy of the
two isotopic systems. Hence, a plot of the 17O value against the 18O value in most terrestrial
systems will produce samples which fall on a line with a slope of 0.53. Samples which are
contain materials enriched in one of the stable isotopes and therefore fall off of the TFL.
Although early work suggested that deviations from the TFL were solely explained by
nucleosynthetic processes (Clayton et al., 1977; Clayton & Mayeda, 1977), it is now recognised
much of the observed mass independent behaviour (Thiemens, 2006; Thiemens & Heidenreich,
1983).
The mechanisms producing different fractionation are not precisely known although there may
processes (Thiemens, 2006). During the modern day non-mass dependent fractionation is
restricted to the upper atmosphere where ultraviolet light has not been diminished by the
presence of ozone. As a result, atmospheric O2 has a small but measureable 17O anomaly.
Similar behaviour has been recognised in the sulphur isotopic system prior to the Great
Oxygenation Event (GOE), around 2.3 Ga, so the deviations from the expected behaviour of 33S
from that of 34S have been interpreted to reflect the evolution of ozone in the atmosphere
(Farquhar et al., 2000). Once sufficient oxygen was produced from the activity of
photosynthetic organisms, ozone reduced the amount of ultraviolet light reaching the Earth’s
surface, thereby inducing a mass independent isotopic effect in the sulphur system. Hence, in
the fossil record, sulphate extracted from sediments prior to the GOE frequently showed
deviations from the TFL, while, without exception, all sulphate samples analyzed in sediments
deposited after the GOE show no mass independent fractionation. How and whether mass
independent fractionation theory might be applied to diagenetic carbonate systems has not
been explored in detail. Clearly sediments formed before the GOE, might have the sulphur
isotopic anomaly removed through diagenetic processes, but the reverse is not possible. In the
case of oxygen, although the majority of meteoric waters fall on the TFL some precipitation
occurring at high latitudes and recorded in ice cores contain small anomalies (Landais et al.,
during evaporation of water (Angert et al., 2004) and such signatures might be incorporated
into carbonates. In addition, carbonates diagenetically altered prior to the GOE might have an
oxygen isotopic anomaly as precipitation during these time periods would have been strongly
Trace and minor elements contained within carbonate minerals can be incorporated either as a
substitute for one of the major structural groups, Ca2+ or CO32-, adsorbed onto the external
crystal surfaces, included as lattice defects and/or as contaminant mineral phases or fluid
inclusions (McIntire, 1963). In the case of Mg and Sr, only 1% of the concentration appears to
be in exchangeable sites within the carbonate (Amiel et al., 1973a) , the rest being substituted
for Ca. Although in practice all elements are incorporated into carbonates to some extent, they
can be separated into the minor elements (Sr, Mg and Na), usually present at concentrations
greater than 100 to 1000 ppm and the trace elements, usually present at concentrations of less
than 1 to 10 ppm.
In open marine environments the concentrations of trace elements in carbonates are generally
very low (<1 ppm) and it becomes difficult to separate elements which are truly substituted for
Ca in the crystal lattice from those which have absorbed onto the exterior crystal surfaces or
those which are present as contaminants. Many elements which are quite common in the
Earth’s crust, such as Fe, are only present at very low concentrations in seawater (<5 ppb) and
consequently have low concentrations in carbonates. This is because while the elements may
metal hydroxides in seawater. Even the soluble portion is largely believed to consist of colloids
complexed to OM (Wu & Boyle, 1998). Such oxides and colloids are deposited rapidly near river
mouths, but also travel long distances behaving very much like clay particles. Many of the
studies which report high values of elements such as Mn, Fe and others (Friedman, 1968;
contamination arises from fluvial input and/or atmospheric dust (Swart et al., 2014). The
metals are either coprecipitated with the carbonates, form coatings on the surfaces of the
carbonates or are precipitated as authigenic minerals and present within fluid inclusions. True
trace element values within the carbonate itself are probably better represented by the
concentrations of Fe, Zn and Cu (<1 ppm; St. John, 1974), and Pb and Cd (ca 10 ppb; Shen &
The most useful environmental information is generally obtained from elements which have
(Doerner & Hoskins, 1925) which relates the ratio of the element (M) under consideration to
(1)
If the value of D is one then there is no preferential accumulation of the contaminant element,
if D is greater than one the mineral preferentially incorporates the metal and if D is less than
one the metal is discriminated against. As a general rule, elements with ionic radii larger than
incorporation into LMC and HMC. During carbonate precipitation and diagenesis, smaller radii
trace elements (Fe, Mn, Zn, Cu and Cd) have distribution coefficients greater than one (Crocket
& Winchester, 1966; Pingitore, 1978) and are therefore concentrated into the diagenetic phase.
The larger radii elements (Pb, Ba and U) act more like Sr and are discriminated against
(Pingitore, 1984; Pingitore & Eastman, 1985). Although the range of concentrations and
distribution coefficients for elements incorporated into the three common forms of calcium
carbonate (aragonite, LMC and HMC) have been tabulated by Veizer (1983), there are wide
ranges in the estimates of the distribution coefficient for most trace and minor elements,
Such uncertainty arises from the differences in the analytical procedures used in determining
the D values, as well as an absence of research in this area. For example, although Veizer
(1983) reported a range of DMn values of between 5 and 30, it is clear that this range is
controlled by the kinetics of the precipitation with the low values being manifested at very
rapid rates and the high values at very slow rates (Lorens, 1981). Although similar inverse
correlations between precipitation rate and D values were reported for Co and Cd, it is likely
In contrast to the smaller radii elements, the DSr for calcite showed a positive correlation with
precipitation rate (Lorens, 1981). Strontium has a distribution coefficient into aragonite of
approximately one yielding concentrations of ca 7000 ppm in the mineral precipitated from
seawater (Banner, 1995; Kinsman, 1969). However, the distribution coefficient is temperature
(Beck et al., 1992; Kinsman & Holland, 1969; Smith et al., 1979). It has also been suggested that
the D value is significantly higher in inorganically precipitated aragonites suggesting that the Sr
aragonite (Milliman et al., 1993). Some biogenic aragonites (i.e. Pteropods) have low Sr
concentrations (ca 1000 ppm) suggesting that organisms can exercise control on the mineralogy
calcification mechanisms can exclude certain ions. Biogenic LMC organisms typically have Sr
concentrations between 1000 and 1200 ppm (corresponding to a DSr of ca 0.1), a value close to
that measured in inorganically precipitated LMC (Holland et al., 1964). Values for the
alteration of aragonite to LMC or biogenic LMC to inorganic LMC have been measured to be ca
0.05 (Baker et al., 1982; Katz et al., 1972a). Such differences are related to the rate of
precipitation of calcite. The values correspond extremely closely to the data from Lorens
(1981) who found the DSr for LMC ranged from values close to 0.1 [similar to the values
determined by Kinsman (1969)] at high rates to very low values at low precipitation rates
[<0.05; similar to but even less than the values reported by Baker et al. (1982) and Katz et al.
(1972b)]. High-Mg calcites have higher Sr concentrations than LMC, perhaps as a result of
Barium
Barium in carbonates is believed to be related to the Ba/Ca ratio in seawater (Lea et al., 1989)
and, because of its nutrient type behaviour (Lea & Boyle, 1991), has been used as a proxy
coral skeletons, which normally have concentrations of ca 2 to 3 ppm Ba, have also been used
as a tracer of riverine runoff (Alibert et al., 2003; McCulloch et al., 2003) although non-fluvial
signals have been observed in some corals (Sinclair, 2005). Similar riverine influences on Ba
Lead
Lead is present in low concentrations in all modern marine carbonates. Lead levels are
sensitive to variations in the external environment, and changes as a result of the addition of Pb
to petrol can be seen clearly in many carbonate records such as corals and sclerosponges
Uranium
Uranium shows relatively high concentrations (2 to 8 ppm) in aragonite, but low values in LMC
(0.02 ppm) and HMC (Amiel et al., 1973b; Chung & Swart, 1990; Gvirtzman et al., 1973; Schoepf
et al., 2014; Schroeder et al., 1970). Because U is mainly complexed with carbonate ions in
seawater, it has been suggested as a potential proxy of the pCO2 in the oceans (Swart &
Hubbard, 1982).
The rare earth elements (REE) refer to a series of 15 elements of increasing atomic weight and
decreasing atomic radii (1.14 Å for La to 0.84 Å for Lu) which generally occur in the same
mineral deposits and have similar geochemical properties. The REE distribution or pattern
usually refers to the REE concentration normalized to a standard reference material such as
Modern carbonates incorporate REEs in ratios similar to those in seawater (Sholkovitz & Shen,
1995) and the REE signal or REY (REE with yttrium) has been used as a proxy for runoff
(Wyndham et al., 2004), pollution (Fallon et al., 2002) and bioproductivity (Wyndham et al.,
2004). Certain REEs such as Ce and Eu exist in multiple valence states. Cerium which has both
Ce3+ and Ce4+ states is mainly present in seawater as the Ce4+ form and is relatively insoluble.
anomaly in the normalized REE pattern. Europium also has multiple valence states and can be
present as either Eu2+ or Eu3+. Because Eu has an ionic radius similar to Ca2+, higher
Another category of trace elements, including univalent elements such as Na+, K+ and Cl-, as
well as anionic complexes such as SO42-, PO43-, NO3- and B(OH)-, are also incorporated in
carbonates (Ishikawa & Ichikuni, 1984; Land & Hoops, 1973; Lyons et al., 2005; Prokopenko et
al., 2013; Staudt et al., 1993; Staudt & Schoonen, 1995; Veizer et al., 1977; White, 1977). These
are termed non-conventional indicators (NCIs) because it is not known precisely where or how
they become incorporated in the mineral structure. Univalent cations might ‘fit’ into the same
locations in the crystal lattice as Ca2+, but there is a problem of charge balance. Where anions
might end up is a matter of speculation. For example, complexes such as SO42- or PO43 might
substitute for a CO32- group in the carbonate structure. Undoubtedly, a large proportion of the
total concentration of these species is either adsorbed onto the exterior of the crystal surface,
is present as contamination or is contained within fluid inclusions. The main use of variations in
concentration of the element, the higher the salinity value (Ishikawa & Ichikuni, 1984; Land &
Hoops, 1973; Veizer et al., 1977; White, 1977). However, as a result of contamination issues,
the equilibrium concentrations of these elements have not yet been defined. Variations in the
concentration and stable isotopes (N and S) of P, N and S are suggested to be related directly to
the concentration of these elements in the environment (Lyons et al., 2005; Montagna et al.,
2006). However, the concentrations of all of these constituents are easily influenced by
contamination and prior to analysis samples must be rigorously cleaned. Even when such
cleaning takes place it is uncertain whether all contamination, or in some instances too much,
has been removed. In addition, microbial processes taking place within the carbonate have
been shown to drastically alter the concentration and isotopic composition of some of these
proxies. For example NO3- is present in the open oceans at very low concentrations (<0.1 M),
yet within skeletons organic material can degrade to NH4+ and then NO3- substantially
increasing its concentration within the interstices. The interior of carbonate particle may
become anoxic leading to denitrification and thereby causing the 15N value to become
elevated. Because during such processes carbonate dissolution and precipitation also takes
place, the NO3- produced and incorporated into the carbonate will reflect diagenetic rather
Effect of Temperature
minerals. In the case of DSr in aragonite there is a strong inverse correlation between
temperature and Sr content in both inorganic (Kinsman & Holland, 1969) and biogenic systems
biogenic systems with the slope for biogenic aragonite, for example corals, being about twice as
steep as that in inorganic aragonite. Although less information exists for LMC, the experiments
of Baker et al. (1982) and Katz et al. (1972a) showed that the DSr was not influenced by
with increasing temperature can be explained by a kinetic control on the DSr for calcite, with the
DSr increasing at higher temperatures (Lorens, 1981). Higher rates of precipitation cause higher
DSr values thus offsetting the thermodynamic tendency for DSr to decrease at higher
temperatures. The distribution of other smaller radii elements (Mn, Co and Cd) tends to
In contrast to Sr, the distribution coefficient of elements such as Mg in biogenic calcite (Chave,
1954) increases with temperature. This property has led to widespread use of the Mg/Ca ratio
as an indicator of temperature in organisms such as foraminifera (Lea, 2002; Lea et al., 1999;
Lear et al., 2000). Once the temperature has been obtained using the Mg/Ca ratio, it can be
combined with the 18O value of the same organism to determine the 18Ow. Salinity can then
be estimated using a published relationship between the 18Ow value and salinity (Flower et
al., 2004; Schmidt et al., 2004). However, while the Mg/Ca ratio is related to temperature,
properties such as salinity, carbonate ion content and pH also appear to influence the ratio
(Arbuszewski et al., 2010), thereby confusing the relationship at least in the tests of
foraminifera. Calculation of salinity using this method is also hampered by large errors
associated with the equation relating the Mg/Ca ratio and 18O value to temperature as well as
uncertainties regarding the relationship between salinity and the 18Ow value.
including U (Min et al., 1995; Shen & Dunbar, 1995), Mg (Mitsuguchi et al., 1996) and Ba
(Gaetani & Cohen, 2006). However, because there are frequently a number of other controls,
the relationships with temperature have proven to be variable. For example, in the study of
Fallon et al. (2003) a large amount of variability was determined in the relationships between
temperature and Sr, Mg, U and B in a number of corals from a transect stretching from the
inshore to the outer barrier. It has also been proposed that a combination of Sr, Mg and Ba
could be used to arrive at a more reliable coral thermometer (Gaetani et al., 2011). Because
coral shows different amounts of discrimination against these elements, it has been argued that
a Rayleigh based approach can be used to make accurate and precise determination of sea-
surface temperature (SST) without the need for species specific calibrations.
In some instances the precipitation of secondary minerals within the interstices of ‘live’ skeletal
materials can contribute significantly to the trace element concentration of carbonates. For
example the mineral brucite (Mg(OH2) has been recorded within coral skeletons (Buster &
Holmes, 2006; Schmalz, 1965) and some calcareous algae (Weber & Kaufman, 1965) thereby
increasing the Mg content. Other secondary minerals within skeletons, such as LMC and
aragonite, have also been widely documented and are known to alter primary elemental and
McGregor & Gagan, 2003). Recently, the presence of a carbonate phase containing
concentrations of Mg approaching that found in dolomite has been noted in the skeletons of
some calcareous red algae with skeletons composed of HMC (Nash et al., 2013). Finally,
algae, sponges and fungi (Golubic et al., 1975; Perkins, 1977; Perkins & Sentas, 1976). These
remove primary skeletal material and the resultant boring frequently becomes filled with fine
mud and/or secondary cements (Bathurst, 1966). Such borings, typically most noticeable
around the exterior of the carbonate grain or skeleton as micritic rinds (Bathurst, 1966), add to
Diagenesis affects the individual components of the rock differentially and therefore different
geochemical results can be obtained by analyzing rocks at different scales. For example,
cements often form between grains without affecting the chemical composition of the grains
crystals forming from the same solution may exhibit different trace element and isotopic
compositions in different growth sectors (Dickson, 1991; Reeder & Grams, 1987; Reeder &
Paquette, 1989; TenHave & Heijnen, 1985) (compositional sector zoning). Bulk stable isotopic
measurements, which are common in many studies, may not capture the chemical signatures
of diagenetic processes unless the entire rock has been altered. The microsampling of
carbonates has progressed from the initial attempts of Dickson & Coleman (1980) who used a
scalpel to excise sufficient material from thin sections, to more modern use of computer-
controlled microdrilling methods. However, these methods are still rather crude and do not
offer the precision necessary to sample material at spatial resolutions of less than ca 100 to 200
m. In addition, there have been attempts to use laser ablation coupled to mass spectrometers
(Dickson, 1991; Dickson et al., 1991; Larson & Longstaffe, 2007; Sharp & Cerling, 1996; Smalley
al., 2003; Treble et al., 2005; Treble et al., 2007) to measure stable isotopes. These latter two
methods offer promise, but are still not used widely being limited by the rarity of the
equipment or poor precision of the results. More widely available methods for microsampling
minor and trace elements in carbonates include the electron microprobe, which typically can
analyze materials at concentrations down to ca 500 to 1000 ppm (Moberly, 1968), secondary
ion mass spectrometry, which can analyze at the ppm level (Swart, 1990; Veizer et al., 1987)
and measure stable C and O isotopes (Vetter et al., 2013; Williford et al., 2013), or nuclear track
methods which can analyze spatial distribution of elements such as U using fission tracks
(Chung & Swart, 1990; Swart, 1988; Swart & Hubbard, 1982) or B and Li using alpha tracks. As
stated by Hudson (1977): “…we must take our limestone to pieces” if we want to adequately
Fluid Inclusions
Fluid inclusions generally represent water and gas trapped during the formation of crystals
which have remained fluid during cooling to normal temperatures (Roedder, 1984). Inclusions
range in size from <2 m to >1 mm and may include water with associated salts as well as
vapour, hydrocarbons, precipitated phases and organic matter. Inclusions can be classified as
being one-phase all-liquid, one-phase all-gas, two-phase or even three-phase. Inclusions can be
primary (i.e. forming at the same time the mineral was precipitated) or secondary, associated
with alteration after the crystal growth. Fluids entrapped at low temperature should remain a
single-phase inclusion while those formed at temperatures higher than 50oC are two phase at
surface temperatures and pressures (Roedder, 1984). In addition, primary inclusions can be
interpretations can be controversial. Assuming that primary fluid inclusions represent samples
of the formation fluid and that the inclusions have remained unaltered, information can be
obtained regarding the temperature and chemical composition of the mineral using a variety of
The simplest technique is to examine a doubly polished wafer of rock with fluid inclusions in a
heating and cooling stage attached to a petrographic microscope. As a normal two phase fluid
inclusion (liquid and vapour) is heated, the temperature at which the two phases become
pressure). Conversely when a sample is frozen and then warmed, the temperature at which the
solid phase (‘ice’) melts is related to the density of the fluids in the inclusion (Potter et al.,
1978). Although the salinity of the fluid can be approximated using a variety of assumptions, it
is important to note that the ions present are not likely to have behaved conservatively and
therefore the estimated salinity may be incorrect. Such temperature and salinity data can be
combined with the 18O value of the carbonate to calculate the 18Ow (assuming a defined
relationship between salinity and 18Ow). Ideally such information should conform to the
inclusions and the temperature derived from the 18O value (using the salinity estimate
obtained from freezing) are usually ascribed to problems associated with the origin of the fluid
inclusions themselves, unknown pressure corrections, or problems relating salinity to the 18Ow
(1990)]. These methods have been utilized both to determine the fluids involved in diagenetic
processes and also within evaporite minerals to determine ratios of ions in evaporative fluids.
Based on these ratios, inferences have been made regarding the chemical composition of
seawater through time (Demicco et al., 2005; Hardie, 2003; Horita et al., 2002). Assuming that
the 18O and 2H values of the fluid inclusion represent the formation fluid, then the 18O value
of the carbonate can be used to estimate the temperature of formation or the evolution of the
18Ow value and the chemistry of seawater through time (Knauth & Beeunas, 1986).
It has been suggested that the predominant form of inorganic marine cementation has changed
from its current favoured form, aragonite, to LMC and then back again to aragonite several
times over the past 600 Myr. This ‘aragonite-calcite’ seas phenomenon, as it is known, was
documented first in the inferred mineralogy of ooids (Sandberg, 1983) and subsequently
considered to have weak control on their calcification processes (Stanley, 2006). It has also
been suggested that some organisms are able to change the mineralogy of their skeletons (Ries
et al., 2006) depending upon the chemistry of the oceans, specifically the ratio of Mg/Ca in the
precipitating fluid. During intervals when the Mg/Ca ratio of seawater was similar to the
modern day, precipitation of aragonite was favoured while during periods of low Mg/Ca ratios
calcite was the preferred form. This notion is supported by inhibited precipitation of LMC in
solutions with high Mg/Ca ratios, the preservation of secular variations in the Mg/Ca ratio of
fluid inclusions from primary halite (Lowenstein et al., 2003), variation in the mineralogy of late
and secular variations in the Mg/Ca ratio of molluscs (Steuber & Rauch, 2005) and echinoderms
(Dickson, 2004). The Mg/Ca ratio in the oceans responds to a combination of changes in the
concentration of both elements, which in turn are thought to be related to changes in sea-floor
spreading and the circulation of seawater from mid-oceanic ridges (Demicco et al., 2005;
Hardie, 2003). In addition the change in Mg/Ca ratio is believed to be accompanied by small
changes in the 26Mg (Higgins & Schrag, 2015) and 44Ca values of seawater (Gothmann et al.,
In Review). As a consequence of the change in the Mg/Ca ratio of seawater, the nature of early
marine cement should change through time as should the 18O, 13C values and minor and trace
element composition. In addition, modification of the Mg/Ca ratio within marine pore fluids,
even close to the sea floor, can result in precipitation of marine cements which are LMC or HMC
For many years it was believed that crystal mineralogy and habit indicated whether cements
formed either in marine or meteoric environments (Folk, 1974; James & Choquette, 1983;
James & Choquette, 1984; Longman, 1980). Acicular aragonite crystals were considered
characteristic of marine dominated fluids with high Mg/Ca ratios, while LMC equant spar and
overgrowths were associated with freshwater. However, several studies have documented the
occurrence of supposed freshwater cements and fabrics in environments which could only be
influenced by marine fluids (Melim et al., 1995; Schlager & James, 1978). Because the
formation of LMC is determined by the Mg/Ca ratio, it is probable that the LMC cements
1996). An additional control on mineralogy can be exerted by changes in the saturation state of
the oceans, related to circulation and depth. For example, the formation of dolomite occurs in
many marine settings and reduces the Mg2+ content of the pore fluids thereby allowing LMC to
be precipitated, while in deeper waters which are undersaturated with respect to aragonite
Sea Level
Earth history is marked by periods when there were only minor changes in sea level (10 to 20
m) while during others changes of 100 m or more occurred over 20 to 100 kyr time scales.
Because the principal cause of large sea-level change is glaciation, warmer time periods in
which the amount of continental ice was reduced, coincided with more stable sea level and
temperatures which were generally warmer than at present. These greenhouse and icehouse
periods tend to coincide with calcite and aragonite seas, respectively. Hence, Cretaceous
carbonates were mainly formed of LMC without the large glacial-eustatic sea-level changes
which characterized the Pleistocene. The absence of such sea-level changes would have limited
the presence of a primary mineralogy of calcite, rather than aragonite, means that the
geochemical signatures of carbonates formed during calcite seas are more likely to be
The use of geochemical tracers to study dolomitisation has been reviewed by various workers
(Budd, 1997; Machel & Mountjoy, 1986; Warren, 2000) since the seminal paper of Land (1980).
Although more than 30 years have passed since the Land paper, and many important
geochemical studies have been published, knowledge of the factors controlling geochemistry of
Dolomites have equilibrium 18O values which are believed to be about 3 to 6‰ heavier than
LMC formed under the same conditions at 25oC. The 3‰ uncertainty arises because there are
at present nine different equations linking the fractionation between the 18Ow and the 18O
value of the dolomite () to temperature (Clayton et al., 1968; Fritz & Smith, 1970; Horita,
2014; Katz & Matthews, 1977; Land, 1983; Northrop & Clayton, 1966; O'Neil & Epstein, 1966 ;
Sheppard & Schwarcz, 1970; Vasconcelos et al., 2005; Zheng, 1999). Some of these equations
were derived experimentally at high temperatures (>200oC) and then extrapolated to lower
temperatures; others are based on the observations of naturally occurring high temperature
dolomites and extrapolating these data to lower temperatures (Sheppard & Schwarcz, 1970). In
contrast, the work of Vasconcelos et al. (2005) determined 18O values in microbial induced
dolomite formed at temperatures between 25oC and 50oC, while the experiments of Fritz &
Smith (1970) generated dolomites between 50oC and 70oC. The study of Zheng (1999) was
based on theoretical considerations. Although the Vasconcelos et al. (2005) equation should
seemingly be the most appropriate at low temperatures, the dolomites produced in those
experiments possessed very poor ordering peaks (or none at some of the temperatures), were
Mg calcite (VHMC; Sibley et al., 1994). Because variations in stoichiometry and ordering can
produce artifacts and lead to variability in 18O values unrelated to environmental conditions
(Land, 1980; Vahrenkamp & Swart, 1994), these low-temperature equations might not be
appropriate for the interpretation of temperature and fluid composition in more mature types
of dolomite.
The 13C value of dolomite is thought to be slightly more positive (ca 1‰) than co-occurring
LMC (Sheppard & Schwarcz, 1970), but is in most instances similar to the sediment from which
it formed. The exception to this occurs where dolomitisation takes place in carbonate poor
sediments and/or the system is heavily influenced by the oxidation of OM (Baker & Burns,
1985; Compton & Siever, 1986; Irwin et al., 1977; Kelts & McKenzie, 1982). In spite of the slight
difference in the 13C value between LMC and dolomite, the processes affecting the 13C values
of dolomites are similar to those affecting other diagenetic carbonates. However, since
abundant, dolomites from these environments may frequently have negative 13C values (Kelts
& McKenzie, 1982) if there are relatively low concentrations of carbonate materials in the
system. For example, dolomitisation is seen in areas where abundant upwelling promotes high
productivity generating siliceous sediments rich in OM (Compton & Siever, 1986; Malone et al.,
1994). Degradation of OM produces dissolved inorganic carbon in pore water with extremely
negative 13C, high alkalinity and low concentrations of sulphate. All of these factors have been
Dolomite exhibits a broad range in crystal ordering and composition ranging from an excess of
10% Ca relative to Mg, to a composition in which there are equal molar amounts. Similar to
LMC, the principal trace elements of interest in dolomites are Sr, Mn and Fe, and only a few
studies have critically examined other trace elements and REEs in dolomites (Shaw &
Wasserburg, 1985; Swart, 1988). The concentration of Sr in dolomite is generally lower than in
calcite formed under the same conditions as Sr is preferentially substituted into the Ca sites in
the crystal lattice (Kretz, 1982; McIntire, 1963); this predicts that Ca-rich dolomites will have
dolomites from a wide range of ages (Vahrenkamp & Swart, 1990). Other elements with larger
radii (Ba and Pb) should show similar patterns to Sr, while elements with smaller radii (Ni, Cu,
Co, Zn, Mn and Fe) should be elevated in stoichiometric dolomites. While this may be the case,
natural variability and input of Mn and Fe from sources unrelated to the control of the Mg/Ca
ratio in dolomites probably overwhelms the stoichiometric control (Vahrenkamp & Swart,
1994). In the case of Mn, whether it is substituted within the Mg or Ca sites, can be measured
using electron spin resonance (ESR; Wildeman, 1970). This and later studies (Lumsden & Lloyd,
1984) indicate that there is more Mn in the Mg sites (compared to Ca sites) by factors of
between ca 1 to 70, a considerably greater range than the value of 1.5 presented in the model
of Kretz (1982). Lumsden & Lloyd (1988) classified the ESR spectra obtained from dolomites
into three categories depending upon whether the position of Mn could be resolved or not.
Broadly speaking these authors determined that stoichiometric dolomites tended to have the
spectral gamma ray methods, that dolomitised reservoirs are frequently elevated in uranium
relative to co-occurring limestone (Swart, 1988). A possible explanation for this is that modern
aragonite dominated carbonate sediments are elevated in U (Chung & Swart, 1990), and if
these sediments are dolomitised early then the U concentration, which is principally complexed
with the carbonate, is preserved. In contrast, if the sediments are altered to LMC first in an
open system, then U is lost and any subsequent dolomitisation is likely to produce dolomites
low in U.
As stoichiometric dolomite incorporates equal amounts of Ca and Mg, attention has been paid
to the fractionation of Mg isotopes during dolomitisation (Higgins & Schrag, 2010; Higgins &
Schrag, 2012). The principal effect appears to be the formation of dolomite in which the lighter
more abundant isotope is preferentially incorporated, leading to more negative 26Mg values in
Strontium Isotopes
The 87Sr/86Sr ratio of dolomites has been used to help constrain the timing of dolomitisation
provided the depositional age is known and the sources of Sr constrained (Saller, 1984). In
situations where the only possible sources of Sr are provided by the original sediments and the
dolomitising fluids are seawater of a later age, the oldest possible age of dolomitisation can be
The 34S of CAS in dolomites can be used in conjunction with the concentration of NCIs to
elucidate the environment of dolomitisation. For example, because it is well known that many
dolomites are formed within the sulphate reduction zone, these should have lower
concentrations of S, normal concentrations of Na, K and Cl, but slightly elevated 34S values.
Dolomites associated with evaporite minerals might have low concentrations of SO42- (as SO42- is
removed during the formation of evaporite minerals), normal 34S values, but elevated values
of Na+, K+ and Cl-. Dolomites formed from brines which have not attained saturation with
respect to gypsum or anhydrite might be expected to show elevated concentrations of all NCIs,
The clumped isotope technique should be able to ascertain which of the various equations
which link the 18O value of dolomite to temperature and the 18Ow value is correct. While a
number of studies have measured the 47 in dolomites (Ferry et al., 2011; Loyd et al., 2012),
reporting both temperatures and 18Ow values, most of these have used a theoretical equation
(Guo et al., 2009) to calculate the temperature from 47 and the 18O–18Ow to temperature
equation of Vasconcelos et al. (2005) to calculate the 18Ow value. Other studies have used
alternative 47 temperature equations, typically the original Ghosh et al. (2006) equation
adjusted for the absolute reference framework (Dennis et al., 2011) or the Passey & Henkes
(2012) equation (Geske et al., 2015; Murray et al., 2014; Sena et al., 2014). These equations are
combined with a range of different 18O–18Ow to temperature equations. A recent study also
suggests that the acid fractionation of 47 at 90oC for dolomite may be greater than the value of
The use of this higher fractionation factor combined with different 47- temperature and 18O–
18Ow temperature equations produces a range of temperatures and 18Ow values which makes
The formation of dolomite requires a source of Mg2+ (usually seawater), a mechanism of Mg2+
supply and a sediment body with sufficient permeability. Although seawater is approximately
1000 times oversaturated with respect to dolomite, any process which raises the saturation
state of dolomite has the potential to increase the efficiency and rate of dolomitisation. Such
processes would include elevation of the salinity of the fluid and/or an increase in alkalinity
caused by oxidation of OM. Clues regarding the interpretation of the geochemical signals left
behind during dolomitisation can be found in recently formed material from areas such as Abu
Dhabi and the Bahamas where dolomites have formed under reasonably well-constrained
conditions. In Abu Dhabi, Holocene dolomites are unambiguously associated with the
evaporation of seawater and the formation of evaporites (Patterson & Kinsman, 1982), and in
some places the formation of microbialites (Bontognali et al., 2012). Both the waters and
dolomites associated with the evaporites have elevated 18O values (18Ow = +2 to +6‰ and
18O =+2 to +3‰; McKenzie, 1981; McKenzie et al., 1980). However, curiously the dolomites
from the modern evaporites in Abu Dhabi are not that much more elevated in 18O than the
deposition (Dawans & Swart, 1988; Fig. 8) and which are generally considered to have formed
from normal seawater with a 18O value of about +1‰. It is possible that the absence of higher
temperatures are initially very poorly ordered and are Ca-rich. Such dolomites might initially
have more negative 18O values than well-ordered and stoichiometric dolomites eventually
recrystallizing to become more ordered and stoichiometric (Vahrenkamp et al., 1991). The
similarity of the 18O value of the dolomites from the Bahamas, which formed from normal
seawater, and those from Abu Dhabi, which are hypersaline, emphasizes the uncertainty in the
use of 18O values to ascertain the nature of the environment of dolomitisation. Older
dolomites and those formed at greater depths tend to have more negative 18O values
reflecting formation or recrystallisation under higher temperatures. One possible model for
dolomite formation maybe that the mineral starts as disordered dolomite or VHMC and
gradually recrystallises to more ordered and stoichiometric forms. In this way the Mg
necessary for dolomite formation is provided early, but the final mature mineral is formed
much later. Along this pathway geochemical patterns, such as stable isotopes and trace
Some indication of this can be observed in studies which show an increase in ordering with
depth (Gregg et al., 1992) although more convincing evidence of the transformation of early
Freshwater
When sea-level is lowered, shallow water carbonates are exposed to the influence of meteoric
fluids and a series of recognizable diagenetic zones are formed (Fig. 9). Such zones include
those associated with sub-aerial exposure surfaces, vadose and freshwater diagenesis, the
mixing-zone and the marine phreatic zone. As sea-level oscillates, the partially-lithified
sediments and rocks are sequentially affected by each of these diagenetic environments,
conditions are usually a composite of a large number of superimposed physical and chemical
diagenetic changes. However, it is likely that the first diagenetic episode will be the most
influential and that subsequent episodes will not affect material already stabilized, but rather
affect only those remaining carbonates not yet influenced by the initial diagenetic event.
Mineralogy
Rainwater contains only small quantities of dissolved salts, derived from aerosols and
atmospheric CO2. The high CO2 and low concentration of Ca2+ causes this water to be corrosive
to all carbonate minerals. Once in the vadose zone, the rainwater (now groundwater) acquires
additional CO2 from the decay and respiration of local OM and initially dissolves the local
calcium carbonate until saturation is attained with respect to the ambient carbonate
mineralogy. In geologically young terrains, the carbonate sediments are mainly composed of
occur before dissolution of aragonite and HMC is complete (Budd, 1988). Precipitation of LMC
causes further undersaturation with respect to aragonite and HMC and the system could
theoretically dissolve the metastable minerals and precipitate the more stable ones until all the
aragonite and HMC are consumed. Further additions of freshwater to the system will then
Although carbonate cementation in limestones was believed by many to result primarily from
exposure to freshwater (Shinn, 1969; Shinn, 2013), Clayton & Degens (1959) suggested that
13C values could help to distinguish between freshwater and marine limestones. The first
studies to utilize 13C and 18O values within diagenetic carbonates were published by Gross
(1964) using samples from the island of Bermuda. These studies showed that the 18O values of
secondary carbonates were controlled primarily by the 18O value of rainfall on the island, while
the 13C value was governed by CO2 derived from the soil zone. Gross (1964) concluded that:
“The 18O/16O and 13C/12C ratios of the limestone appear to be useful in determining their
diagenetic history where the geologic relationships are relatively simple and when the original
isotopic composition of the carbonates and water is known.”. The recognition that diagenetic
carbonates acquire the 18O signature of their recrystallizing fluids is the basis for numerous
other interpretations of the behaviour of 18O values in diagenetic carbonates. However, since
the 18O value of the carbonate also depends on the temperature of alteration there has always
temperature estimate, such as that obtained from fluid inclusions or clumped isotopes, has
the relationship between temperature and the 18O values of calcite or dolomite (see below for
discussion of dolomite) for fluids with varying 18O values. Plotting a data set on such a
diagram necessitates knowledge of the 18O value of the solid as well as either the original
temperatures of deposition, or the 18Ow. Ignoring for the moment the problem of
distinguishing variations in the 18O value of the carbonate caused by temperature as opposed
to the 18Ow, most of the work dealing with early diagenesis has interpreted the changes in the
18O value of the carbonates as originating from the fluids. This is because most early
diagenetic changes are considered to take place at temperatures between 20oC and 30oC which
would introduce an uncertainty of only a few per mille in the 18O value. As a consequence,
variations in the 18Ow values are considered to dominate the early diagenesis of carbonates.
Changes also occur in the composition of trace and minor elements during diagenesis (Budd &
Land, 1990; Land & Epstein, 1970; Saller & Moore, 1991). Such changes usually involve the
dissolution of metastable carbonates (aragonite and HMC) releasing trace elements into the
pore water, followed by precipitation of more stable phases such as LMC and dolomite.
Variations in the concentrations of Mn and Fe in the pore fluids or varying rates of precipitation
can lead to the formation of cements with luminescent characteristics (Frank et al., 1996;
Meyers, 1974), characteristics which have been used to establish paragenetic sequences and
identify freshwater diagenesis. It is believed that high concentrations of Mn (> ca 50 ppm) lead
to luminescence, but this can be altered in the presence of Fe which tends to quench the
phenomenon. However, some workers have induced changes in luminescence without altering
The diagenetic behaviour of REEs has been investigated in a few studies (Johannesson, 2012;
Scherer & Seitz, 1980; Shaw & Wasserburg, 1985; Webb et al., 2012; Webb et al., 2009). The
most recent study shows that meteorically altered carbonates appear to have similar values
when compared to the original sedimentary components (Webb et al., 2009) although a Ce
anomaly was noted together with a slight depletion in the light REEs. However, ubiquitous
contamination from oxide phases and diagenetic incorporation from pore waters confound the
use of REEs (Della Porta et al., 2015; Haley et al., 2004; Reynard et al., 1999). Laser ablation
methods, coupled to ICP-MS, eliminates the ubiquitous contamination seen in bulk samples and
allows the REE signatures of individual cements to be targeted. Hence characteristic primary
signatures can be confirmed in cements believed to be of marine origin (Della Porta et al.,
2015).
Uranium
Uranium generally follows the behaviour expected of a large radii element and freshwater
cements generally have much lower concentrations compared to the original biogenic
components. However, it has been noted that U concentrations with freshwater phreatic
cements are significantly higher than in vadose cements, a phenomenon which was ascribed to
accumulation of uranium and the longer residence time of water here (Chung & Swart, 1990).
A secondary cause related to the lower pH in the phreatic zone which lower the concentration
of CO32- ions and hence raised the UO2(CO3)22-/CO32- ratio of the fluid.
The impact of marine and freshwater diagenesis upon the 34S (CAS) and 11B values has not
regimes which are not anoxic does not appear to alter the 34S of the CAS (Gill et al., 2008).
There is a suggestion that the 11B values are altered in the freshwater phreatic zone associated
Calcretes
Calcretes or caliche surfaces form in a variety of ways and represent the sub-aerial weathering
of carbonate surfaces (Esteban & Klappa, 1983; James, 1972; Rossinsky & Swart, 1993;
Rossinsky et al., 1992; Sheldon & Tabor, 2009; Watts, 1980) correlating with periods of
emergence (Beach & Ginsburg, 1982; Kievman, 1998; McNeill et al., 1988). Calcretes can form
on the surfaces of bare carbonate rocks and within soil zones (Fig. 10A and B). On bare
carbonate rocks, laminae form in response to dissolution of the surface by corrosive rain and
then the subsequent precipitation of carbonate minerals during drying events. During the
wetting events small depressions in the rock are colonized by microalgae and cyanobacteria
which photosynthesize and then decay, imparting an overall negative 13C signature to the
water. Evaporation causes the water to become isotopically enriched in 18O and therefore
these laminated carbonates and/or calcretes have negative 13C and occasionally positive 18O
values. Caliches found within the soil zone are characterized by root structures (rhizoliths)
(Klappa, 1980; Fig. 10C) and also have more negative 13C values, but are rarely enriched in 18O
(Rossinsky et al., 1992); this is because meteoric waters rapidly penetrate the soil layer thus
protecting the fluids from evaporation. Soil carbonates therefore tend to reflect patterns of
workers (Rozanski et al., 1993). The magnitude of 13C depletion within the soil zone has been
related to the nature of the vegetation and atmospheric pCO2. For example, the 13C values of
diagenetic carbonates from soil profiles in East Africa allowed the timing of grassland
Palaeo-atmospheric pCO2 levels have been inferred from the 13C value of soil carbonates
based on the work of Cerling (1991) who recognised that soil CO2 is a mixture of atmospheric
and soil-respired components. This approach has been used by a number of workers to
estimate atmospheric pCO2 through time (Ekart et al., 1999; Montanez et al., 2007), although
the model is based on many different assumptions and the constancy of such assumptions
through time is difficult to verify. One of these assumptions is that the soil carbonate formed at
the 18O values (Dworkin et al., 2005), interpretation is complicated as a result of changing
18Ow values and the fact that carbonate seems to form mainly during dry periods (Breecker et
al., 2009). The solution to this problem might be to apply a technique such as the clumped
isotope thermometer (Passey et al., 2010) although some studies, supported by petrographic
analyses, have suggested that subsequent alteration of the soil carbonates may be a significant
Calcretes and caliche crusts are typically elevated in Fe, Mn, Si and Al (Machusak & Kump, 1997;
Rossinsky et al., 1992). In the Bahamas such elements are derived from atmospheric dust
1970). Carbonates forming deeper in soil horizons (termed penetrative calcretes) also have
elevated Fe and Mn concentrations, but typically lower than calcretes forming near the surface
(Rossinsky et al., 1992). These elements are present not only substituting for Ca in the calcite
Vadose Zone
The vadose zone is defined as the portion of the rock or regolith above the water table.
Technically this zone also includes the calcrete and caliche portion of the soil profile and
consequently receives abundant contributions from the decay of OM and CO2 from the
respiration of roots. Rainwater entering the vadose zone contains only small quantities of
dissolved salts, derived from aerosols and atmospheric CO2. Here the water initially dissolves
local calcium carbonate until saturation is attained with respect to the ambient carbonate
precipitation of LMC can occur before dissolution of aragonite and HMC is complete (Budd,
1988). Within the vadose zone, LMC precipitates as meniscus and pendant cements as the
water drips between grains (Halley & Harris, 1979), with the shape of the cements reflecting
the concave nature of the water surfaces (Fig. 10D). Allan & Matthews (1982) described this
zone as showing a wide range in 13C values, but with a rather constant and negative 18O value
(Fig. 11). Changes also occur in the composition of trace and minor elements during diagenesis.
Such changes usually involve the dissolution of metastable carbonates (aragonite and HMC)
releasing trace elements into the pore water, followed by precipitation of more stable phases
such as LMC. Because the vadose zone is located immediately below the soil horizon, cements
to 1000 ppm; Kievman, 1998). These elements can be derived from atmospheric dust in
locations remote from terrestrial influences (Rossinsky et al., 1992; Swart et al., 2014).
The fresh-water phreatic zone is located below the water table and is characterized by both
lateral and vertical flow. Here isopachous equant cements grow into pore space filled with
water, forming a characteristic dog-toothed spar (Halley & Harris, 1979; Fig. 10E). At the water
table itself diagenetic reactions are particularly aggressive as OM, which has filtered through
the vadose zone, accumulates at the water-air interface. Here the concentration of oxygen is
lowered as OM is oxidized and sulphate reduction (utilizing the small amount of sulphate in the
meteoric fluid derived from aerosols) causes pH to decrease, and dissolution and precipitation
reactions to take place (McClain et al., 1992). Based on changes in the water chemistry (SO42-,
alkalinity, Ca2+ and Sr2+) the reaction zone in the study of McClain et al. (1992) appeared to be
localized to within 10 to 20 cm of the water table. The sediment in the majority of the
freshwater phreatic zone in that study was uncemented leading McClain to argue that a
changing sea level caused the active diagenetic zone to move through the sedimentary column
promoting cementation (Budd & Land, 1990; McClain et al., 1994). In contrast Budd et al.
(1988) suggested that the original aragonite and high-Mg calcite sediments within the
freshwater phreatic zone would eventually stabilize to LMC as a result of the mineralogical
drive (James & Choquette, 1982), the difference in solubility between metastable minerals such
as aragonite and HMC and LMC. Regardless of the mechanism of stabilization, within the fresh-
water phreatic zone the 18O values are still more negative (similar to the vadose zone; Allan &
compared to the vadose zone, although still isotopically negative relative to the original marine
sediments. Although it is generally believed that the freshwater phreatic zone is characterized
by the presence of cements as shown in Fig. 10E, muddier portions of the section frequently do
not show such cements, yet still possess relatively negative 13C and 18O values (Fig. 10F).
The pattern of alteration of marine sediments by meteoric fluids has been described by
Lohmann (1987) as following an inverted ‘J’ pattern (Fig. 11) in 13C and 18O space. Original
carbonate sediments, which fall within a range of 13C and 18O values depending on their
origin (Fig. 5), undergo dissolution and precipitation reactions mediated by meteoric fluids
changing the bulk 18O of the rock towards more negative values. In an open system there is
abundant fluid moving through the rock and therefore the 18O value of the original carbonate
becomes quickly masked by the local meteoric fluids as the rock is altered. At the same time
the 18O value is altered, the 13C value of the diagenetic carbonate is influenced by oxidized
OM. As a result of the large amount of carbon in the sediment/rock compared to the water
(contrasting with oxygen), alteration of the 13C value takes significantly longer than the 18O
value. The evolution of the carbonate as traced on a 13C versus 18O plot shows the 18O value
moving into equilibrium with the groundwater at a specific temperature while the 13C values
span a wide range. This is the so-called inverted ‘J’ trend (Fig. 11). The vertical lines on this
diagram are described as the Meteoric Calcite Lines (MCL) and represent the 18O values of
calcite in equilibrium at a specific temperature with ground waters of different 18O values. As
the MCL varies as a function of the 18O content of the rainfall, it also varies relative to latitude
the 18O value of calcretes varies throughout the Caribbean (Rossinsky & Swart, 1993) relative
to the amount of rainfall and its 18O value. As a final observation many diagenetic pathways
might not follow the inverted ‘J’ if the original sediment has a 18O value close to or even more
negative than that represented by the MCL. In addition, the transition from the original
sediment might take place rapidly and might not be represented in the rock record. This is
Mixing-zone
The mixing-zone describes the transition between fresh and marine waters occurring along
coastlines and small oceanic islands. As a result of the non-linear behaviour of activity
coefficients of ions in fluids of varying salinity, the mixing of fresh and saline waters, which are
both oversaturated with respect to carbonate minerals such as calcite, can produce waters
which are undersaturated relative to the same minerals (Badiozamani, 1973). Such behaviour
can cause extensive dissolution (Back et al., 1986; Whitaker & Smart, 1997). Mixing-zone
associated caves are therefore common within carbonate terrains that are close to sea level.
These mixing-zones might also be regions of dolomite formation because while fluids in the
mixing-zone are undersaturated with respect to calcite, these same fluids are still highly
supersaturated relative to dolomite (Badiozamani, 1973). While this notion was briefly very
popular with several authors claiming to have documented the occurrence of mixing-zone
dolomites based on stable C and O isotopes and minor elements (Humphrey, 2000; Ward &
Halley, 1985), the evidence is not compelling. The 18O values for example cannot be used to
dolomitisation has fallen out of favour principally as a result of the absence of dolomite in most
modern day mixing-zone environments (Budd, 1997; Hardie, 1987; Machel & Mountjoy, 1986;
Melim et al., 2004). Notwithstanding the arguments regarding its significance, Allan &
Matthews (1982) have shown the mixing-zone to be characterized by a strong linear covariation
between 13C and 18O values. It has been suggested that this results from alteration in fluids
with progressively more positive 18O values as one progresses from fresh water to marine
Blue Holes
Blue Holes are collapse features observed in modern carbonate terrains such as the Bahamas;
they are believed to form as a result of dissolution related to the movement of the mixing-zone
through the carbonate platform as sea-level oscillates. Modern Blue Holes in the Bahamas have
a layer of freshwater water overlying seawater. Unlike groundwater settings, the OM in the
Blue Holes appears to sink through the freshwater, accumulating at the seawater interface
(Jones et al., 2014). Here the OM is rapidly degraded utilizing all the available oxygen and then,
with help from the sulphate in the underlying seawater, copious amounts of H2S are produced.
The presence of H2S, combined with the naturally corrosive characteristics of the mixing zone,
produces enhanced dissolution in these caves as the interface oscillates under the influence of
In the previous example, the mixing zone has been defined to represent interacting solutions
with varying salinities. However, non-conservative behaviour can also take place when two
strengths, but with different CO2 fugacities, are mixed. Undersaturation with respect to
carbonate minerals occurs as the changes in the concentrations of the various carbonate ions
are not conservative. Such behaviour is common in groundwaters charged with CO2 after
passing through soil zones and then mixed with a second body of water with a lower pCO2.
Evaporative Waters
Climate, principally the amount of rainfall, can also control surface related diagenesis. Two
good modern examples are the Bahamas and the Persian Gulf, both located at about the same
latitude. In the Persian Gulf, rainfall is very low and evaporation extremely high. This leads to
the development of hypersaline conditions in tidal flats and lagoons and the formation of a
suite of minerals associated with the evaporation of seawater (evaporites). In contrast, the
Bahamas experiences greater precipitation and low evaporation. Consequently, evaporites are
rarely found associated with modern sediments in the Bahamas. Despite the prevalence of
evaporites in arid settings, saline fluids can also form in water bodies with restricted circulation.
Although such fluids might reach more modest salinities (40 to 120), they are highly saturated
relative to carbonate minerals and in some instances are associated with higher amounts of
organic degradation or higher rates of photosynthesis. Such saline waters could be dense
enough to sink or reflux through underlying sediments causing diagenetic alteration as they
migrate. Generally such fluids would have positive 18O values, but might have either positive
or negative 13C values depending upon the relative amounts of photosynthesis and
of Ca2+ and CO3- from solution, the remaining fluids are enriched in Mg2+, Na+, Cl-, K+ and SO42-.
During the later stages, minerals such as gypsum, dolomite and halite can eventually be formed
along with aragonite and a host of other evaporative minerals (Schreiber & El Tabakh, 2000).
Because the distribution coefficient for the incorporation of Sr into evaporite minerals is lower
than one (Holser, 1979), the amount of Sr in the residual brine increases during precipitation of
these minerals and celestite (SrSO4) can be found as an accessory mineral. Although other
trace elements can also change concentration depending upon their distribution coefficients
and the minerals being formed, there are no data concerning the values of the distribution
coefficients for elements such as Mn, Fe and Mg into common evaporite minerals.
Evaporite associated carbonates are frequently dolomitised as a result of the high Mg/Ca ratio
of the evaporative fluids (McKenzie, 1981; McKenzie et al., 1980), the presence of abundant
OM which promotes high alkalinity (Schreiber et al., 2001) and certain types of bacteria. All of
these conditions are known to promote the formation of dolomite and protodolomite
(Vasconcelos & McKenzie, 1997; Warthmann et al., 2000). Since depletion of O2 and the use of
sulphate as an oxidant of OM leads to the production of CO2 with negative 13C values,
dolomites (and other carbonates) formed under such conditions tend to have more negative
13C values than marine carbonates. The 34S values of evaporate associated sulphate minerals
are generally believed to reflect the 34S of the sulphate in the oceans. However, wide ranges
in 34S values found in evaporite deposits during certain times probably reflect processes of
Sea Floor
Evidence of the influence of early marine diagenesis on the 18O and 13C values of carbonate
particles is poorly documented in the literature. Processes such as micritisation and the
precipitation of aragonite and HMC cement undoubtedly affect the 18O and 13C values of the
original sediments, but such processes are difficult to assess as the cements and areas of
alteration are generally too small to allow effective sampling without cross-contamination of
the original material. It is likely, however, that the addition of cements to a marine framework
originally possessing more negative 13C values will tend to enrich the samples in 13C (as
inorganic aragonite and HMC cements tend to have more positive 13C values; Aissaoui, 1988;
Grammer et al., 1993). Studies that have sampled such cements confirm their more positive
13C values (compared with normal skeletal aragonite) and 18O values close to equilibrium
(Braithwaite & Camoin, 2011). Processes influenced by photosynthesis, such as those adding
cement within stromatolites, might have more positive values as a result of the preferential
removal of 12C or alternatively more negative values as a result of CO2 produced during
sulphate reduction (Andres et al., 2006). The 18O value of the inorganic cements is also likely
to be different depending on the nature of the original grains and the temperature of
precipitation. The trace and minor element concentrations might also be altered by
micritisation and precipitation of aragonite cements. Micritised borings are often filled with
HMC and therefore would be lower in Sr and higher in Mg, if the grain was originally aragonite.
During periods of non-deposition, cementation takes place near or at the sea floor resulting in
the formation of hardgrounds and firmgrounds. In cases where carbonate cementation occurs
in colder and deeper waters than the environment in which the sediment formed, the 18O
value of the subsequent cements will be more positive than the host sediment, overprinting or
modifying the original bulk sedimentary signature. The 13C value of the precipitate may also be
positive as inorganic aragonite precipitated in equilibrium with normal ocean water would be
expected to have a 13C value of +4‰. This process has been argued to account for positive
13C and 18O values at hardground surfaces in ancient rocks (Marshall & Ashton, 1980).
However, cementation in hardgrounds can also lead to more negative 13C values as products
this can be found in cores drilled in the Bahamas (Fig. 12), where a surface at a depth of 536 m
in the Clino core, representing a hiatus of 2 to 3 Myr, became cemented with carbonates
possessing more negative 13C and more positive 18O values than those of the original
sediment (Swart & Melim, 2000; Fig. 13). These sediments were originally derived from the
surface of GBB and had 13C and 18O values in equilibrium with shallow, warm water with
relatively positive 13C values. During the period represented by the hiatus, they became
burrowed and enriched in OM. This OM was eventually degraded by oxic and anoxic processes
leading to the production of CO2 depleted in 13C and cementation in cooler waters. Additional
examples of negative 13C values in hardgrounds are provided by Dickson et al. (2008). In this
Implications
The Deep Sea Drilling Project (DSDP), the Ocean Drilling Program (ODP) and their successors
have substantially increased current knowledge of the diagenesis of carbonates in the deep
marine environment. Pore waters can be retrieved from such sediments by squeezing the
described by Martin & Sayles (2003). The analysis of the geochemistry of such pore waters is
potentially much more powerful than examining the sediments themselves. For example, if
dolomite is formed in a sediment with 50% porosity utilizing the local CaCO3 and Mg2+ derived
from the pore water, then only ca 0.1 % of dolomite could be generated utilizing all the
available Mg2+ in the pore fluid. Such a drastic change in the pore water chemistry could easily
be measured, but it might be very difficult to find the corresponding 0.1% of dolomite in the
bulk sediment.
Marine Burial
In the upper portion of the sedimentary column, close to the sediment/water interface,
oxidation of OM proceeds using O2 as the terminal electron acceptor. This process generates
CO2 which, in turn, causes the pH to decrease and the pore waters to become undersaturated
with respect to the more soluble carbonate minerals. During low rates of sedimentation the
products of oxidation escape into the overlying water column and the pore fluids become
of OM, but little physical or chemical alteration of the sediments, although some sea-floor
cementation may take place under low rates of sedimentation. As rate of sedimentation
increases, the supply of O2 and the removal of CO2 are limited by diffusion. The most
favourable alternate oxidant under anoxic conditions is Mn4+, followed by NO3- and Fe3+.
Compared to O2, which provides energy of 3190 kJ for every mole of glucose oxidized, these
alternate oxidants yield ca 3000, 2750 and 1400 kJ mol-1, respectively (Froelich et al., 1979).
However, in practice the concentrations of these species (Mn4+, NO3- and Fe3+) in the modern
oceans are very low and they are not normally important in the degradation of OM.
Volumetrically the most important alternative oxidant to O2 is SO42-, even though the energy
yield of 380 kJ mol-1 is considerably lower than that of the other oxidants. Typically in modern
sediments there is a transition from the zone in which oxygen is the principal oxidant to one
where sulphate is utilized by bacterial sulphate reduction (BSR) to degrade OM [anoxic sulphic
zone (Berner, 1981)]. Regardless of the oxidant utilized, all these processes contribute CO2 with
negative 13C values, similar to the parent OM. In the SO42- reduction zone abundant H2S is also
produced. Once SO42- has been completely utilized the oxidation of OM proceeds through the
process of methanogenesis [anoxic non-sulphic zone (Berner, 1981)] which results in a large
fractionation between CH4 and CO2, resulting in CO2 with positive 13C values (+10‰) and CH4
with very negative 13C values (-40 to -60‰; Games et al., 1978). It is believed that
methanogenesis terminates once the geothermal gradient elevates the temperature above the
viable limit for BSR (Coleman et al., 1979). Below this depth the OM degrades under the
influence of thermal processes which produce isotopically negative CO2 and CH4, as well as
upwards through the sediments and be utilized by bacteria within the BSR and oxic zones
The transitions as described above occur not only within the pore water in sediments, but also
in stratified anoxic water bodies such as lakes or basins, for example, the Black Sea (Murray et
al., 1991) or Cariaco Basin (Scranton et al., 1987). During certain periods of Earth history, such
as the Cretaceous, the number of anoxic basins was significantly greater than today (Jenkyns,
2010) and during the early Earth all oceans were anoxic (Busigny et al., 2014; Lyons et al.,
2014).
Limitations imposed by the supply of oxidants have important consequences regarding the
extent to which carbonate sediments can be influenced by the degradation of OM. Consider 1 g
of carbonate sediment composed of aragonite (density = 2.9 g/cm3) with 50% porosity. Based
on a seawater SO42- concentration of 28 mM, within the pore space there would be
approximately 9.6*10-7 moles of SO42- available for the oxidation of any OM present in the pore
fluids. That would produce twice that amount of HCO3-. In contrast there are 0.01 moles of C in
the 1 g of carbonate sediment. By simple mass balance calculation it is apparent that the
oxidation of OM within carbonate sediment cannot alter the 13C value of the original sediment
significantly unless additional oxidants are supplied. Although such a supply is possible through
the process of diffusion, considerable time would be necessary to allow sufficient SO42- to
diffuse downward from the overlying seawater. In addition, the greater the depth over which
diffusion takes place, the less likely it will be that the 13C value of the bulk sediments will be
realm, it will be difficult to significantly alter the 13C value of the original sediments, even if
100% of the original sediment is dissolved and reprecipitated. The exception to this would be
hiatus, or in sediments with a low rate of deposition. Alternatively if carbonates are a minor
contributor to the sediment, then the 13C value of any diagenetic component will be strongly
influenced by the oxidation of OM. If the system were open, sufficient electron acceptors may
be available to oxidize all the OM. However, by definition, in an open system the products of
diagenesis would be removed from the source of oxidation. The influence on the diagenetic
carbonate would be balanced by the rate of alteration and the rate of fluid flow.
In the majority of cases the geochemical parameters in the pore waters of fine-grained and
muddy sediments are controlled by diffusive rather than advective processes. A diffusive
between the interstitial pore waters and an external source such as the overlying seawater or
an underlying reservoir. An ion diffuses from a region of high concentration to one of low
Berner (1980). In this equation the change of the concentration of any component (C) as a
function of depth (z) and time (t), bulk diffusion coefficient (Db) and diagenetic change (J) is
expressed by Eq. 2:
C C (vC)
Db z z ( J i ...J n ) (2)
t z
and trace elements in pore waters squeezed from sediments, but also in the geothermal
gradients measured in the sediments. There are instances, however, where advective
processes control the temperature and geochemical profiles. An advective process is one in
which there is movement of fluids driven by external processes unrelated to the actual
concentrations of ions. While a certain amount of advection occurs during compaction, this is
slow relative to diffusion. Significant advective processes can be recognised as either concave
the sediments, or concave downward profiles, indicating movement of fluid into the sediments.
Typically in oceanic sediments, these processes occur associated with hydrothermal processes
near spreading centres (Kastner et al., 1986), in carbonate platforms, where large-scale
circulation is perhaps driven by differences in temperature between the platforms and the
adjacent oceans (Kohout, 1967), density difference between the waters on the platform and
the adjacent oceans (Simms, 1984) and along continental margins, where there is a hydrological
head associated with the adjacent mainland (Sayles & Manheim, 1975). The movement of
seawater throughout carbonate platforms by Kohout convection and reflux of brines with
slightly higher salinity than normal seawater has been discussed in numerous, mainly
theoretical studies (Jones et al., 2002; Simms, 1984). Such processes are potentially more
important in controlling diagenesis than diffusive ones, in that advection is able to supply
reactants needed for the alteration of sediments as well removing the products of these
diagenetically active elements such as Ca2+, Sr2+ and Mg2+ to one another, i.e. Mg2+/Ca2+ and
Sr2+/Ca2+. The excess or deficit of an element or species (x) relative to a conservative element in
(3)
If the concentration of an element behaves in a conservative manner then the excess is zero; if
the element is consumed then the excess is negative and if the element is generated the value
is positive. An example of this approach using data from ODP Site 1005 located on the margin
of GBB (see Fig. 12 for location) is shown in Fig. 14. At this site, excess (deficit) concentrations
of all minor species cluster near zero in the upper 50 m of the core. This indicates either an
absence of geochemical reactions in this portion of the core or a rapid flushing of fluids so that
the products of diagenesis are removed. Below this depth Ca2+, Mg2+ and SO42- all decrease in
concentration, while the alkalinity increases. Decreases in Ca2+ and Mg2+ reflect the
precipitation of LMC and dolomite, while changes in alkalinity and sulphate concentration
(4)
In the absence of: (i) reactions which utilize HCO3-; and (ii) preferential diffusion of HCO3-
relative to SO42-, the deficit of SO42- should be 28 mM and the excess of HCO3- 56mM (utilizing
Eq. 4; Fig. 15). The precipitation of carbonate minerals can also be seen in changes in the ratios
data are plotted from two localities, one from cores in Florida Bay (Fig. 12; Burns & Swart,
1992; Swart et al., 1987a) and from ODP Site 1005 (Kramer et al., 2000) drilled at a water depth
of 888 m. Despite the difference in core locations, the diagenetic reactions which occur at
these two localities are similar. For example, the precipitation of LMC at both sites results in a
reduction of the Ca2+/Cl- ratio (as Ca2+ is consumed) and an increase in the Sr2+/Ca2+ ratio (as
Sr2+ is preferentially rejected as a consequence of the low distribution coefficient for the
formation of LMC; Fig. 16). Changes in the concentration of Mg2+ and Ca2+ resulting from the
precipitation of dolomite also take place, but will depend upon the stoichiometry of the
reaction involved. In the case of dolomitisation using Eq. 5 there will be a reduction in the
Mg2+/Cl- and Ca2+/Cl – ratios and an increase in the Sr2+/Ca2+. The formation of dolomite by this
equation should not change the Mg2+/Ca2+ ratio of the pore fluids as Mg2+ and Ca2+ are used in
(5)
In contrast dolomitisation utilizing Eq. 6 involves the dissolution of existing calcium carbonate
and therefore this reaction increases the Ca2+/Cl- ratio while decreasing the Mg2+/Ca2+ ratio:
(6)
Pathways for other possible diagenetic reactions such as aragonite and HMC dissolution and
preferentially excluded and consequently this ion builds up in the pore fluids. An increasing
Sr2+/Ca2+ can suggest which of the reactions are actually taking place. Such changes are readily
visible by using plots of the Mg2+/Cl- with respect to Ca2+/Cl -, SO42-/Cl- and other ions as
As a first approximation, the diagenesis of carbonate buried in the marine realm can be
separated into those which overly basaltic basement, those found along the margins of
Modern oceanic sediments are dominated by LMC producing organisms. At the seawater
interface the sediments form an ooze and are uncemented with porosities of 60 to 70% being
typical. With increasing depth the sediments become compacted, thereby reducing porosity
and the processes of dissolution and precipitation eventually transform the ooze into a hard
rock.
Oceanic carbonates typically show an increase in the concentration of Ca2+ and a decrease in
Mg2+ in the interstitial pore fluids with increasing depth [between 1 m and 300 m below sea
floor (mbsf)]. These changes arise partially as a result of dissolution and precipitation within
at the interface between the sediments and the underlying igneous rocks (McDuff & Gieskes,
1976). In addition secular changes in the Ca2+ and/or Mg2+ of seawater can modify the profile
(Higgins & Schrag, 2012). Reactions at the basalt-sediment interface produce Ca2+ and consume
Mg2+ setting up a strong diffusion gradient in the interstitial pore water from the sea floor to
values of the pore waters. The two examples shown in Fig. 18 are taken from DSDP Site 504
(Mottl et al., 1983), situated above basaltic basement on the Costa Rica Rise and DSDP Site 541
(Gieskes et al., 1984), located on the Barbados accretionary prism. Interstitial-pore fluid
mM, while Mg2+ values decrease from 55 mM to less than 20 mM close to the boundary
between the sediment and the underlying basalts. The 44Ca in pore waters in oceanic
sediment typically become more negative as the carbonates are altered to LMC and impart
their original negative 44Ca value to the pore fluids (Fantle & DePaolo, 2007) .
Strontium
Although Cenozoic oceanic sediments mainly consist of foraminifera and coccoliths composed
of LMC, such biogenic carbonates tend to be metastable compared to inorganic LMC and over
time dissolve and precipitate forming inorganic LMC (Baker et al., 1982). Oxidation of OM by
consequent precipitation of the more stable inorganic carbonate phases. With depth , there is a
steady increase in the concentration of Sr2+ in the pore fluids (Baker et al., 1982; Gieskes, 1976)
The Sr2+/Ca2+ ratio in the pore fluids can either decrease, increase or remain the same,
depending on the magnitude of Ca2+ and Sr2+ changes. The concentration of Sr2+ in the pore
fluid increases until it reaches saturation with respect to the mineral celestite (SrSO4).
Consequently celestite is frequently found coincident with the Sr2+ maximum in the pore water
(Baker & Bloomer, 1987), suggesting direct precipitation from the pore fluids. In the absence of
sulphate reduction, the maximum concentration of Sr2+ which can be attained in the pore fluid
abundant sulphate reduction the concentration of Sr2+ can reach much higher values (see
below). In some instances external sources of sulphate diffuse into the reduction zone from
underlying sediments (Kramer et al., 2000). In these cases Sr2+ in the initial pore fluids is
removed (as a result of the precipitation of celestite) causing a sink of Sr2+ deeper in the
sedimentary column. Strontium in such situations can diffuse both upward and downward
from the Sr2+ maximum zone (Gieskes et al., 1986). The 87Sr/86Sr ratio is also an indicator of Sr2+
diffusion within the interstitial pore fluids. Strontium released to the interstitial pore fluids
initially has a 87Sr/86Sr ratio similar to that of the local sediment. As the Sr2+ diffuses up or down
through the sediment column this signature changes along with the transported Sr2+, not only
because the 87Sr/86Sr ratio is different, but also as a result of preferential diffusion of 87Sr
relative to 86Sr. Hence if one considers the last ca 60 Myr, during which the 87Sr/86Sr ratio of the
oceans has become increasingly more radiogenic towards the present day, then in the
sediments above the pore fluid Sr maximum, the 87Sr/86Sr ratio of the pore fluids is typically less
radiogenic than the sediments themselves. Below the region of maximum concentration the
can be seen in the example from Site 541 (Fig. 18). Here the concentration of Sr2+ reaches a
maximum of 400 M around 300 mbsf. This corresponds to the point at which Sr2+ diffuses
upwards imparting a less radiogenic (older) signal to the pore waters in the younger sediments
Oxygen Isotopes
The18Ow values in pore fluids decrease with increasing depth at oceanic sites (Fig. 18), as a
result of the alteration of the underlying basalts to clay minerals (Lawrence et al., 1975). This
trend for the pore fluids to assume more negative 18Ow values is, however, ameliorated to
some extent by the tendency for pore waters to become enriched in 18O as a result of
Carbonate precipitated along such a gradient would be depleted in 18O, not only as a result of
changes in the 18Ow values, but also as a result of the increasing temperature. In the upper
portion of oceanic sediments there is a slight increase in the 18Ow values which is not related
to carbonate dissolution and reprecipitation reactions. This change is a result of the increase in
the 18Ow value of glacial seawater which is now preserved in the pore fluid profile. With time
this increase will diffuse away, but the current 18Ow profile has been used to model the change
in the 18Ow values of the oceans during the last glacial period (Adkins & Schrag, 2003; Schrag,
1996).
The extent to which SO42- is utilized in the pore fluids depends upon the supply of OM to the
location and the rate of sedimentation. In the open ocean where sedimentation is low, OM is
effectively oxidized near the sediment/water interface and the proportion of OM which
survives is reduced. As a consequence, decreases in SO42- and increases in alkalinity in the pore
waters are low. For example, at Site 709 drilled in the Indian Ocean, SO42- decreases only
slightly in the upper portions of the sedimentary column and the concentration of SO42- and
(Backman et al., 1988). As a result, rates of dissolution and precipitation are relatively low
leading to only a small increase in the concentration of Sr2+ in the pore fluids. As this particular
site overlies basaltic basement there is still a large increase in the concentration of Ca2+ and
Diagenetic carbonates formed during the burial of deep ocean carbonates tend to show
decreasing 18O values with burial, both as a result of the increased temperature and as a
consequence of the decreasing 18Ow values of the pore fluids caused by basalt–water
interaction. In contrast, the 13C values of the diagenetic carbonate should not be significantly
different than the 13C values of the original sediments (see previous discussion). Pore water
Sr2+ concentrations increase with depth in oceanic carbonates, but a combination of more
rapidly rising Ca2+ concentrations and an increasing temperature would tend to lower the Sr
concentration of the diagenetic calcite. The combination of a low 18O value and low Sr
Using Pore water profiles for Estimating the Amount of Dissolution and Precipitation
Various workers have utilized pore water profiles, particularly Sr and 18Ow but more recently of
Ca, Mg and U isotopes, in order to understand and model rates of recrystallization in deep sea
carbonates (Baker et al., 1982; Fantle et al., 2010; Killingley, 1983; Lawrence et al., 1975;
Richter & DePaolo, 1987; Schrag et al., 1995; Stout, 1985). The most recent review of the
Strontium: The concentration of Sr2+ in the pore waters has been used to estimate the amount
of recrystallization in oceanic sediments, by comparing the total amount of Sr lost from the
sediments over a defined period of time with the total amount of Sr available from a given
amount of sediment (Baker et al., 1982). An alternative model is to compare the expected
Sr/Ca (or Mg/Ca) ratio of the sediments with the expected values based on the Sr, Mg and Ca
concentrations in the pore water. At the sediment-water interface the Sr/Ca in the sediment
will be substantially different to that predicted from the pore waters [using a DSr of 0.04 (Baker
et al., 1982)], but with depth these values will converge and complete replacement of the
sediments by a diagenetic derived calcite is considered to have occurred when these two values
enabled better constraints to be placed on the diagenesis and hence accurately predict
dissolution precipitation reactions and the evolution of 87Sr/86Sr (Richter & DePaolo, 1987).
(Lawrence et al., 1975; Schrag et al., 1995; Stout, 1985). In Eq. 7 the 18O value of the pore fluid
(7)
In this equation Mc = mole fraction of oxygen in the carbonate, Mw = mole fraction in the pore
fluid, C = 18O [SMOW (standard mean ocean water)] of the initial sediment, w = initial 18O
value of the water, R = amount of transformation from the parent mineralogy and =
fractionation factor between calcite and water at a specific temperature (O'Neil et al., 1969).
Using this approach and assuming that the 18O values of the present pore water profile has
been similar throughout the history of the sediment, rates of recrystallization, porosity and
geothermal gradients can be varied, and the predicted 18O value of the sediment matched to
Calcium Isotopes: A more recent approach to estimating recrystallization has been through the
application of similar models as used by Richter & DePaolo (1987), but applied to the 44Ca
values of pore waters and sediments (Fantle & DePaolo, 2007). The method relies on the
difference between the equilibrium 44Ca values in Modern marine carbonates which show a
fractionation of 44Ca of 0.9987‰ compared with carbonates which are dissolved and
reprecipitated in the deep sea, a process in which there is no fractionation (= 1.000). Based
on data from ODP Site 807A from the Ontong Java Plateau, Fantle & DePaolo (2007)
determined that rates are very high in relatively young sediments (30 to 40% Myr-1), decreasing
than those calculated using the methods of Baker et al. (1982) for DSDP Site 288 (Baker et al.,
Carbonate sediments deposited on the margins of carbonate platforms, below the extent of
carbonates for several key reasons. First, the sediments contain mixtures of different forms of
calcium carbonate [aragonite, HMC and LMC; periplatform sediments (Schlager & James, 1978)]
and are therefore more reactive than pelagic LMC. Second, they frequently have higher
compared to oceanic settings. Third, they are situated significant distances above basement,
thus limiting the influence of basement-carbonate interactions upon the pore fluids and the
diagenetic minerals produced. Despite these differences, at least some of the geochemical
patterns observed in the pore fluids from oceanic carbonates are also observed in pore fluids
the pore waters causes initial undersaturation with respect to biogenic minerals. In contrast to
oceanic sediments periplatform sediments frequently have more than one carbonate phase and
therefore once the fluids are below the saturation state of aragonite and HMC, dissolution is
able to proceed until all of the metastable phases have been altered to LMC, even in the
absence of the continued oxidation of OM. This mechanism is similar to that described in
carbonate sediment exposed to freshwater saturated with respect to LMC. The dissolution and
reprecipitation of carbonates is readily visible in thin sections taken from sites such as Site
(Fig. 19A) are composed of mixtures of pelagic and platform-derived material and are
unconsolidated. With increasing depth the sediments start to dissolve and cements appear
extending into the interparticle pore space and mouldic porosity (Fig. 19B). Early cemented
portions are resistant to compaction (Fig. 19C), but susceptible to further alteration and
fracturing (Fig. 19D). During burial dolomite can be formed either replacing original sediments
or precipitating in voids (Fig. 19E). During deep burial the sediments become compacted and
The concentrations of Ca2+ and Mg2+ in the interstitial fluids in platform derived sediments
behave differently when compared to oceanic sediments. This is because, in contrast to the
continued increase of Ca2+ and the decrease in Mg2+ in pore fluids of oceanic sediments,
periplatform sediments are influenced to a greater extent by local diagenetic reactions rather
Ca2+ can decrease in zones where there is abundant carbonate precipitation, while in other
against during dolomitisation, the 26Mg of the pore waters frequently increase at sites of
dolomite formation (Higgins & Schrag, 2010). At other sites where dolomite is not forming, the
26Mg of the pore waters decreases with depth as a result of preferential incorporation of 26Mg
Plateau, Site 823 adjacent to the Great Barrier Reef (GBR) and Site 1005 adjacent to the GBB]
(Fig. 19A and B), the Ca2+ decreases initially reflecting a balance between precipitation and
dissolution and then gradually increases to values between 20 mM and 30 mM. Magnesium
Strontium
Similar to oceanic sediments, Sr2+ in the pore fluids of slope carbonates increases with depth.
However, because both aragonite and HMC initially have much higher concentrations of Sr than
coccoliths and foraminifera which dominate typical pelagic deposits, there is potentially much
more Sr available for release into the pore fluids. In addition, HMC and aragonite are meta-
stable minerals and hence are more soluble than LMC. Concentrations of Sr2+ in periplatform
sediments therefore tend to rise more quickly with burial than in oceanic sediments. Similar to
oceanic sediments, the maximum concentration of Sr2+ which can be reached in the pore fluids
is dictated by the ion activity product of celestite. This mineral is more prevalent in platform
derived sediments than oceanic sediments (Baker & Bloomer, 1987; Swart & Guzikowski, 1988).
Frequently sediments which have been partially cemented become fractured and celestite
precipitates along the fractures. In other instances the celestite forms cement which
encompasses altered and unaltered sediments. Similar mechanisms were probably also
responsible for the formation of celestite in ancient carbonates (Yan & Carlson, 2003). As rates
of deposition and concentrations of OM are higher in periplatform than oceanic sediments, the
concentration of SO42- in the pore waters can frequently become completely exhausted
allowing the concentrations of Sr2+ in the interstitial pore fluids to reach very high levels. Three
is situated off of a submerged carbonate platform (Queensland Plateau) located to the east of
the GBR. This site shows low initial concentrations of aragonite and only small decreases in the
interstitial SO42- concentrations. Consequently the Sr2+ in the interstitial pore water reaches
values similar to those seen in oceanic sediments. As a result of the low concentration of Ca2+
in the pore waters, the predicted Sr of any diagenetic calcite formed from these pore fluids
reaches values in excess of 4000 ppm (Fig. 20A). At ODP Site 823 located adjacent to Site 817,
in the trough between the Queensland Plateau and the GBR (Fig. 20B), the concentration of
sulphate in the interstitial pore fluids falls to values close to zero. As a consequence the
concentration of Sr2+ reaches very high values (3 mM). The predicted Sr concentration of
diagenetic calcite is in excess of 8000 ppm. The final example, from the Bahamas (ODP Site
1005), is also one in which SO42- decreases to zero throughout a large portion of the section
(Fig. 21A). The high amount of initial aragonite in the section and the processes of carbonate
dissolution and precipitation allow the concentration of Sr to reach up to 5 mM. The predicted
concentration of Sr in any diagenetic carbonate formed from these pore fluids would be 4% and
these concentrations are reflected in diagenetic cements at ODP sites such as Site 823 (Dix,
1995)
Although the behaviour of Sr-isotopes in the pore fluids from platform derived sediments
appear to be similar to that seen in the ocean carbonates, there is a difference which relates to
the rapid rates of dissolution and precipitation exhibited by the periplatform sediments. In the
case of Sites 1005 and 817, the pore waters always appear to have the same 87Sr/86Sr ratio as
the contemporaneous sediments. This is probably a result of the high rate of carbonate
observed at ODP Site 823 is closer to that observed in the oceanic studies.
Attempts have been made to apply the methods to calculate rates of dissolution and
precipitation outlined by Baker et al. (1982) to periplatform deposits from the Bahamian (Swart
& Guzikowski, 1988) and Maldives carbonate platforms (Swart & Burns, 1990). In both
locations the Sr diffusion method was determined to give low estimates of dissolution and
would be intuitive that these sites would show higher estimated rates of dissolution and
precipitation. It was concluded that the low rates were a result of the high initial Sr content of
the sediments without a corresponding increase in the Sr gradient in the pore waters and were
probably a result of the formation of celestite. While the Sr-diffusion method gave low rates of
sediment dissolution and precipitation in situations where there was an initially high
concentration of aragonite or the composition of the sediments was variable through time
Oxygen Isotopes
Typically the pore fluids of sediments deposited adjacent to carbonate platforms, such as the
Bahamas, show a steady increase in the 18Ow value of the pore fluids as a result of dissolution
Figure 22 shows the patterns of 18Ow values in the pore water obtained from five cores drilled
to depths of ca 1000 m adjacent to GBB. The carbonate content of the sediments at all of these
sites is typically in excess of 95% (Eberli et al., 1997). Although at the top of each of the cores
18Ow values of glacial seawater (Adkins & Schrag, 2003; Schrag, 1996), the steady increase in
18Ow of the fluids with depth reflects dissolution and precipitation at higher temperature and
can be modelled using a similar approach to that outlined for oceanic sediments (Eq. 7). The
results of the 18O values of the bulk sediment, diagenetic component and pore waters are
shown in Fig. 23. Using a rate of recrystallization of 3% per 100 m the models agree with the
While the example of the Bahamas serves to illustrate the differences between oceanic and
platform carbonates, the data only show changes in the upper ca 1000 m below the sea floor.
In fact the Bahamas rests on about 5000 m of mixed carbonates and evaporites which in turn
overlie Jurassic siliciclastics (Goodell & Garman, 1969; Schlager et al., 1988). The extent to
which these sediments are influenced by post depositional/diagenetic process is not known.
However, one process for which evidence exists at the present day is a steady increase in Cl-
with depth in the upper 1000 m (Eberli et al., 1997). Such an increase has also been observed in
other areas associated with carbonate platforms (Davies et al., 1991). Although it has been
suggested that this increase results from dissolution of an underlying NaCl unit and subsequent
continual dissolution and precipitation (Kramer et al., 2000) of carbonate. The Cl- is derived
from the release of this element during dissolution and its subsequent discrimination during
dissolution and precipitation. If the high concentration of Cl- suggests higher amounts of
dissolution and precipitation at greater depths it is also possible that the 18Ow values of pore
fluids could be elevated as high as +8 to +10‰ depending on the amount of dissolution and
without changes in salinity, hence decoupling salinity and 18Ow. Once formed such positive
18Ow fluids can be mobilized through fractures and react with shallow carbonate sediments.
Such fluids are likely to be Ca2+ rich, but Mg2+ poor as a result of dolomite formation and are
not likely, therefore, to be dolomitising, but rather leading to dedolomitising reactions and the
precipitation of LMC.
Oxidation of OM and its subsequent influence upon the pore waters and sediments is generally
principally as a result of the much higher burial rates found at such sites. With increasing depth
the SO42- can become completely exhausted leading to large increases in alkalinity. Other trace
components such as NO3-, NH4+ and PO43- derived from the degradation of OM might also be
visible in the pore waters, although these tend to be utilized by bacteria, absorbed onto
carbonate grains and, in the case of NO3-, used as an electron acceptor during oxidation of OM.
Depending on the extent of the carbonate reactions the alkalinity can remain high and conform
to the 2:1 relationship as expressed in Eq. 4 (Fig. 15), or the alkalinity can be consumed in
carbonate precipitation reactions. During the bacterial consumption of SO42- the 34S values of
the residual pore water increases and CAS incorporated into diagenetic carbonate along this
gradient will have 34S values higher than expected. The alkalinity produced from the oxidation
of OM, while tending to have a 13C value which reflects that of the OM, is usually masked by
the C released from the parent carbonate sediments. Because the absolute amount of SO42- in
the pore water within a given amount of sediment is small and the supply from diffusion
Below the zone of SO42- reduction, OM is oxidized by methanogens using CO2 and producing
CH4. The CH4 can migrate upwards through the sediment column where it can either escape to
the sea floor, be consumed by bacteria, or form CH4 hydrates, usually in association with H2S
and CO2 (Kvenvolden, 1981). The process of hydrate formation alters the salinity of the pore
fluids and produces positive 18Ow values as 16O is preferentially incorporated into the gas
hydrate. This positive 18O value can be recognised in diagenetic carbonates formed in
association with the hydrates (Hesse & Harrison, 1981). In the case of carbonate dominated
sediments, the positive 13C value of the CO2 produced during methanogenesis is not evident in
the bulk 13C value as the amount of carbon produced is small compared to the C in the parent
sediment. For example, during Leg 182 extremely high amounts of biogenic methane were
found, yet there was no noticeable influence on the 13C values of the sediments (Feary et al.,
1998; Swart et al., 2000). This process has been described in ancient carbonate poor sediments
With increasing burial depth the 18O values of diagenetic carbonates decreases, reflecting
sediments which are largely composed of carbonate, the 13C values of the altered sediments
typically will not change as the oxidation potential of pore waters in a diffusive regime is too
low to influence the 13C value of the carbonate (as in the case of the oceanic carbonates).
trend can be modified by the permeability of the sediments and the amount of siliciclastic
material. For example, in locations off of the Bahamas which show alternations in the amount
of non-carbonate material, the intervals with lower carbonate percentage show better
preservation of the original carbonate constituents than the high-carbonate intervals (Frank &
Bernet, 2000). These intervals also possessed more negative 13C and 18O values, attributes
which could reflect diagenesis or more likely the 13C and 18O values of different source
materials.
The Sr content of the diagenetic carbonates would be expected to be high, reflecting the extent
high as 3000 ppm (Dix, 1995) and dolomites with values of 2000 ppm (Swart & Melim, 2000)
have been reported. Such high Sr concentrations in ancient rocks might be misinterpreted as
reflecting precipitation from hypersaline brines. Because the favoured mineralogy of calcium
carbonate precipitated in the oceans has changed through time (Sandberg, 1983), the nature of
periplatform sediments will also have varied. Hence during the Cretaceous, a period of calcite
seas, periplatform sediments would have behaved more like oceanic sediments regarding the
buildup of Sr in the pore fluids. Consequently, it might be expected that the Sr concentration of
diagenetic calcite formed during these times will be lower than during times of aragonite seas.
During the movement of fluids through carbonate edifices, the dissolution and precipitation of
carbonates will cause elements with low distribution coefficients for the precipitation of LMC
and dolomite, such as Sr, to increase in concentration in the fluids in the down-flow direction.
concentration in the same direction. This approach has been proposed to trace the direction of
fluid flow and hence constrain the nature of diagenesis, although the results have been
inconclusive (Vahrenkamp & Swart, 1994). The 18Ow values and diagenetic carbonate formed
as a consequence can also change in the direction of fluid flow. For example, fluids which are
heated then advected upwards will gradually cool meaning that there should be a trend for the
18O values of the altered carbonates to be more positive away from the cooling. Conversely,
fluids being drawn into the margin of a carbonate platform, by a process such as Kohout
convection, will be initially cold and gradually increase in temperature. During this process the
diagenetic carbonate will be isotopically more negative and the fluids more positive, similar to
the case of sediments being buried adjacent to the margin of a carbonate platform.
An example of large scale advection is evident from the pore water and solid geochemistry of
samples from ODP site 812 drilled on the Queensland Plateau, a submerged carbonate
platform, located off of the eastern coast of Australia. Here coring of the upper ca 40 m of
sediments revealed that the pore water profiles were controlled by diffusion showing patterns
typical of sediments deposited adjacent to carbonate platforms (Fig. 24; Davies et al., 1991). At
about 40 mbsf, a hard cemented layer was encountered which was not recovered. Located
dolomite sand and silt. The concentrations of elements such as Ca2+, Mg2+, Sr2+ and SO42-within
these sediments were similar to that of seawater (Swart, 1993). The 87Sr/86Sr of the pore fluids
in the lower section also returned values similar to those in modern seawater (Elderfield et al.,
conditions from the sea floor to the well bottom, indicating that the formation was under
pressured and seawater was being sucked into the carbonate platform (Davies et al., 1991).
What was encountered at ODP Site 812 was probably part of a large circulation cell involving
modern seawater in the Queensland Plateau perhaps driven by the temperature differences
between the platform and the adjacent seaways (Kohout Convection). In this case the
circulation of seawater probably promoted the dolomitisation of the sediments. The 87Sr/86Sr
of dolomites in these sediments was measured and found to fall between the values typical for
sediments of this age and modern seawater (McKenzie et al., 1993; Fig. 24). Systems such as
this might be responsible for the formation of the precursors of marine fibrous calcites, a
common form of cement found precipitated in the pore space of carbonate accumulations
throughout the Phanerozoic (Kim & Lee, 2003; Richter et al., 2011; Tucker, 2001). Although in
most instances these are composed of LMC [fibrous dolomites have also been reported (Richter
et al., 2014)] such cements have been suggested to be either original (Kendall, 1985; Saller,
1986) or replacements of either aragonite (Kendall & Tucker, 1973) or HMC cements (Lohmann
& Meyers, 1977; Mazzullo et al., 1990; Wilson & Dickson, 1996). The presence of micro-
inclusions of dolomite or empty voids has been used as evidence of a prioir HMC mineralogy,
but these features are not always present making the interpretation controversial. Although
their apparent paucity during the Tertiary is problematic, it is possible that they represent the
replacement of either modern HMC or aragonite cements found in modern reef framework as
described by Aissaoui (1988) and others (Ginsburg & James, 1976; Grammer et al., 1993).
Differences in the original mineralogy of these cements could be induced by changes in the
the carbonate framework, changes of the Mg/Ca ratio in the oceans and/or changes relating to
the saturation state at the depth of formation. Although these cements have been widely used
as a proxy for interpreting the geochemistry of seawater, it has been suggested they may not
The principal reason why continental margins have been considered separately from oceanic
and platform sediments is because of the influence of the hydrological head on the adjacent
mainland. These fluids can be either meteoric in origin, and therefore possess low salinity, or
be remnants of hypersaline fluids. For example, ODP Site 905 drilled along the New Jersey
margin showed a decrease in Cl- with depth suggesting freshwater influence (Gieskes, 1974;
Mountain et al., 1994) probably originating from the US mainland. Similar explanations have
been proposed for ODP sites drilled off the coast of South America (Kastner et al., 1990). In
contrast, other sites show evidence of saline brines (Feary et al., 1998; Kastner et al., 1990;
Sayles & Manheim, 1975); for example, sites drilled during ODP leg 182 off of the southern
coast of Australia in the Great Australian Bight (GAB). These sites exhibited large increases in
salinity with increasing depth suggesting the presence of hypersaline lagoons on the adjacent
mainland during previous sea-level low stands (Feary et al., 1998; Fig. 25). In addition to the
role of non-marine fluids, sediments along continental margins tend to be principally siliciclastic
and are therefore relatively carbonate poor and rich in OM. For these reasons the
concentrations of Sr2+ in the interstitial pore fluids tends to be low in spite of abundant SO42-
reduction. The 87Sr/86Sr profiles at such sites are complicated, reflecting the input of Sr both
felsic minerals which typically have very elevated 87Sr/86Sr ratios (Kastner et al., 1990).
Concentrations of Ca2+ and Mg2+ tend to be low in settings unaffected by saline brines as
authigenic carbonates are precipitated consuming available Ca2+, Mg2+ and CO32-. In the
presence of saline brines the concentrations of all interstitial components can rise depending
upon the rates of reactions. For example, at Sites 681 and 680 off the Peruvian margin,
concentrations of Ca2+ and Mg2+ increased dramatically with depth and behaved more or less
conservatively (Kastner et al., 1990). In contrast, at sites off of the GAB (1127, 1129 and 1131),
concentrations of Ca2+ and Mg2+ stayed low despite a threefold increase in salinity (Feary et al.,
1998). Here both Ca2+ and Mg2+ are being consumed in carbonate precipitation reactions. An
example from Site 1127 is shown in Fig. 21B. Typically 18Ow values of pore fluids found in
continental sediments are negative reflecting mineral–rock interactions with silicates as well as
Many continental margin settings tend to have high rates of upwelling and hence high organic
productivity (Emeis & Morse, 1990; Emeis et al., 1991). In such locations sediments tend to be
siliceous, carbonate poor and rich in OM. The high organic content arises as OM is produced,
not only on adjacent shelf deposits and in the water column but also from rivers draining the
mainland. Typically, therefore, both the rates of preservation and oxidation of OM are high
and the pore fluids significantly altered with respect to the concentration of SO42-, alkalinity and
other constituents associated with the degradation of OM (Kastner et al., 1990). Organic
material frequently survives below the zone of SO42- reduction and methanogenesis produces
continental margins compared to other locations. Continental margins also tend to have lower
carbonates formed in these environments could conceivably be influenced by the negative 13C
value of the CO2 produced during SO42- reduction and methanogenesis. The 13C values of
carbonates found in such locations are therefore not related to the 13C value of the original
carbonates or the 13C value of the DIC in the ocean during deposition and the use of 13C
values from such sections to interpret changes in the global 13C cycle should be avoided. In
instances where pore fluids are influenced by adjacent saline brines an additional supply of
SO42- is provided, allowing greater amounts of OM to be oxidized and higher alkalinity pore
waters to be produced. In the example from Leg 182 off the GAB, the three-fold increase in
salinity produced a similar initial increase in SO42-. This SO42- was consumed during sulphate
reduction producing alkalinity values over 100 mM. Such high amounts of sulphate reduction
produced high concentrations of H2S which, together with CH4 formed by methanogenesis,
resulted in the possible production of CH4–H2S–CO2 hydrates at this location (Swart et al.,
2000). However, because the sediments were composed of >90% carbonate minerals, even this
extreme degree of SO42- reduction was insufficient to alter the bulk 13C value of the sediment.
quite variable because in such situations both meteoric and hypersaline fluids have been shown
to be present. In instances where there are low initial amounts of carbonate sediments, the
13C value of the pore water can become dominated by the products of the oxidation of OM
sediments such as those occurring off the GAB, even extensive OM diagenesis does not
While sediments deposited above oceanic crust experience limited burial, carbonates formed
along continental margins and on carbonate platforms can be eventually buried to great
dissolution and precipitation reactions are enhanced. Where fluids have encountered
evaporites they may be Na rich and have very high salt contents (Land & Prezbindowski, 1981).
Continued movement of such fluids results in the formation of diagenetic evaporate minerals
displaced from their original depositional sites (Machel, 1993). Carbonate recrystallization at
high temperatures can elevate the fluid 18O and in contrast to the example from the Bahamas
where 18O values reach ca 2 to 3‰ (Fig. 19), formation water 18Ow values of over +20‰ have
been reported for the Cretaceous (Land & Prezbindowski, 1981). Fluids originating at these
depths can be mobilized into shallower environments along faults, allowing hydrothermal
processes to affect rocks in much shallower burial environments. Although typical 18O values
of such hydrothermally altered rocks are very negative (ca -10 to -20‰; Katz et al., 2006),
suggesting high temperatures, as the 18O value is a product of both temperature and the
18Ow, the interpretation is often complex and even carbonates with more positive 18O values
Hydrothermal dolomites are defined as those formed at temperatures higher than the ambient
formation temperatures (Davies & Smith Jr., 2006) and are therefore associated with fault
systems controlled by structural deformation. Such faults allow higher temperature fluids to
migrate from deeper formations and interact with the host rocks precipitating and/or dissolving
carbonates along the way. For example, such a mechanism has been proposed for the
formation of the dolomites in the Arab-D Formation (Swart et al., 2005) and has recently been
supported by the use of clumped isotopes which documented that the dolomites formed at
temperatures in excess of 60 to 70oC and from fluids with positive 18Ow values (Swart et al.,
2013). Such deep basinal fluids, however, are probably not primarily dolomitising being
depleted in Ca2+ and Mg2+ and probably corrosive to carbonate minerals; they can therefore
dissolve the calcite from partially dolomitised bodies creating a highly permeable and porous
Stylolites
Stylolites are roughly planar features, present at a variety of scales from microscopic to metre
length or longer, which represent surfaces where rock material has been lost by pressure
solution. During compaction there is typically dissolution of the host carbonate, a loss of
porosity (Rittenhouse, 1971) and insoluble material (clay, OM, pyrite, etc.) is left behind leaving
a distinct planar or ragged surface. Stylolites are of particular interest in the study of reservoir
characterisation because they act as permeability barriers (Heap et al., 2014); they have been
extensively studied (Rye & Bradbury, 1988; Saller & Dickson, 2011; Swennen et al., 2012) .
Some work shows isotopically more negative 13C and 18O values associated with the stylolites
changes in the 18O values can be interpreted as dissolution of carbonate and reprecipitation at
higher temperatures, while the more negative 13C values indicate the incorporation of
isotopically negative carbon perhaps released during the thermal degradation of OM.
At high temperatures (120 to 140oC), evaporite minerals (gypsum and/or anhydrite) and/or
sulphate within the formation can react with any petroleum present to produce significant
quantities of H2S and H2O according to a simple reaction as shown in Eq. 8 (Vandeginste et al.,
2009; Worden et al., 2004; Worden & Smalley, 1996). This process is known as thermo-
chemical sulphate reduction (TSR) with a typical example shown in Fig. 26 where the reaction of
(8)
This process results in the formation of isotopically light carbonate, because the 13C value is
inherited from the CH4 and more negative 18Ow values (Worden & Smalley, 1996). The salinity
of fluids associated with TSR is substantially lower than the normal formation fluids. Typically
the diagenetic carbonate forms rinds around existing evaporite minerals with relatively
negative 13C and 18O values (Vandeginste et al., 2009). In some instances pyrite and
elemental sulphur are formed in the process. Based on the geochemical signatures and mineral
associations, some authors have invoked TSR related reactions during the formation of saddle
dolomite (Machel, 1987b), a type of dolomite known to be formed at high temperatures (Radke
profile of a geochemical parameter and then assigning ages to specific inflections in the record
or correlating the inflections between locations with similar profiles. Often the inflection points
have not actually been dated and it is only assumed that similar patterns at geographically
stratigraphic tools include 13C, 18O, 34S and the 87Sr/86Sr ratio (Veizer et al., 1999).
Strontium Isotopes
The 87Sr/86Sr of marine carbonates has changed gradually over time reflecting a balance
between the erosion of continental rocks, which typically have high 87Sr/86Sr ratios, and the
recycling of ocean water through mid-oceanic ridges where the 87Sr/86Sr ratio is low (Burke et
al., 1982). In oceanic settings as described earlier, diffusion of Sr along a concentration gradient
in the pore fluids can change the 87Sr/86Sr of the sediments if there is a significant amount of
alteration. If the rates of dissolution and precipitation reactions are high, as in the case of
platform and periplatform carbonates, then the 87Sr/86Sr of the pore fluids tends to be
indistinguishable from that of the original marine carbonates. Diagenetic carbonates formed
along such a profile have 87Sr/86Sr ratios similar to the original sediments and therefore, in spite
of diagenesis (Swart et al., 2001), this ratio can be used to obtain some idea of the depositional
age by comparing the ratio in the rock with established 87Sr/86Sr curves for the oceans (DePaolo,
1985; Koepnick et al., 1985; McArthur, 1994). In other instances, the 87Sr/86Sr ratio can be used
constrain the timing of diagenesis provided the Sr source to the system is known (Saller, 1984;
Variations in the18O values of sediments have proved to be important for dating during the
Pleistocene (Emiliani, 1955; Lisiecki & Raymo, 2005). The 18Ow values of the oceans increased
during glacial periods, reflecting the buildup of continental ice, and then subsequently
decreased as ice melted. By analyzing organisms such as foraminifera which continually track
these changes, the age of sediments with maximum or minimum 18O values can be dated by
reference to known 18O curves. By comparing the variations of 18O values in benthonic and
planktonic foraminifera, the relative influence of temperature and ice volume has been
ascertained. During the Holocene and Pleistocene, the large temperature difference between
surface and deeper waters is reflected in the 18O values of benthonic and planktonic
foraminifera. This difference decreases as one progresses back in time and has been
interpreted as indicating that deep waters were warmer in the past (Savin, 1977). An
alternative interpretation is that the smaller numbers of benthonic foraminifera have been
affected by the dissolution and precipitation reactions of the more abundant planktonic
foraminifera. With time, dissolution and precipitation in a closed system will homogenise the
18O values of benthonic and planktonic foraminifera. Any differences between the 18O values
of the benthonic and planktonic foraminifera, therefore, might in fact be a diagenetic artifact
In sediments older than the Miocene, diagenesis is too pervasive to allow the 18O value of
carbonates to be useful for detailed chronostratigraphy. While there is a tendency for the 18O
values of carbonates, marine cherts and phosphorites to become increasingly negative over the
past 600 Myr (Degens & Epstein, 1964; Keith & Weber, 1964; Knauth & Lowe, 1978; Veizer &
temperature of the early Earth, increased diagenetic alteration, or changing 18Ow values of the
oceans. The various theories have been presented in numerous papers (Knauth & Epstein,
1976; Knauth & Lowe, 1978; Perry, 1967; Perry & Tan, 1972) and new life has been injected into
the debate with the emergence of the clumped isotope technique which theoretically should be
able to distinguish between the various options, providing that solid state resetting has not
taken place.
Carbon Isotopes
The 13C values of carbonate sediments in conjunction with the 13C values of OM have been
used extensively to document changes in the global carbon cycle (Hayes et al., 1999). Increases
in 13C values have been mainly interpreted as reflecting an increase in the burial of OM, while
decreases indicate increased oxidation of OM. Correlations between the 13C value of inorganic
and OM is usually taken to indicate a 13C signal unaltered by diagenesis, and therefore one
which reflects perturbations in the global carbon cycle (Bachan et al., 2012). Many detailed
13C records have been measured, principally in bulk carbonate rocks, and these have been
correlated globally for large portions of the Phanerozoic and Proterozoic rock record (Kennedy,
1996; Koch et al., 2014; McKirdy et al., 2001; Saltzman et al., 2004). Although the best samples
for determining global changes in the carbon cycle are pelagic carbonates such as foraminifera,
or even bulk oceanic sediments, these are not available for time periods older than the
Cretaceous–Jurassic when records from carbonate platforms or epeiric seas have to be used.
There are problems with such locations and widespread disagreement as to whether these
records reflect differences in sediment types, local or global variations in the carbon cycle or
2005). Despite an almost euphoric tendency to accept that global changes in the 13C values of
carbonates are original in nature, there are several viable alternative explanations for these
global trends including variations in the origin of the sedimentary materials and diagenesis.
Origin of Sediments
The origin of carbonate sediments and their influence on 13C stratigraphy and interpretations
of the global carbon cycle can be seen in a study of sediments deposited along the slope of the
GBB (Swart & Eberli, 2005). Such a setting might be analogous to an ancient carbonate margin,
although the types of organisms and their mineralogy have undoubtedly changed through time.
At this location a series of five cores were taken during ODP Leg 166 (Eberli et al., 1997;
Shipboard Scientific Party, 1997; Fig. 12). In previous studies it was shown that the composition
of periplatform sediments respond to changes in sea level (Droxler et al., 1983; Reuning et al.,
2006; Roth & Reijmer, 2005). As the platform surface material is isotopically enriched in 13C,
the 13C signal in periplatform sediments also varies as a function of sea-level change (Swart &
Eberli, 2005). During highstands, abundant shallow-water material is produced on the platform
top and deposited along the margins. During lowstands, the principle source of carbonate is
pelagic organisms. Because these two sources have quite different 13C values, the result is a
correlation between sea level and 13C values, as well as a correlation between the 13C value
and position relative to the platform margin. Using bio-stratigraphy and seismic stratigraphy,
the 13C changes of the bulk carbonate were correlated between the five cores in a proximal to
distal transect. In an ancient setting such changes in 13C values might be interpreted as
reflecting changes in the global carbon cycle, but in the case of the Bahamas, they do not
There is an absence of correlation because changes in the Bahamas are driven by global sea
level, not the global carbon cycle, and in fact similar changes in 13C values have been observed
along carbonate platform margins in the Indian and Pacific Oceans (Swart, 2008). Another
feature of the sediments at these locations is that their 13C values become more positive
towards the present day. This trend was interpreted to reflect a gradual progradation of the
platform margins towards the present position of the cores, reflecting increased input of
The potential influence of diagenesis upon the 13C value of shallow water carbonates is well-
established. Consider the example of the Bahamas transect cores in which the 13C values of
individual sequences from ODP Sites 1004, 1005, 1006 and 1007 fall approximately on a 1:1 line
(Swart & Eberli, 2005) when plotted against data from Site 1003. The data from cores Clino and
Unda, which were cored on the surface of GBB, an area subject to repeated sea-level changes,
clearly do not plot on this line (Fig. 27). Curiously the types of trends in 13C and 18O values
associated with Pleistocene sea-level changes (Allan & Matthews, 1977; Melim et al., 2004;
Melim et al., 2001; Melim et al., 2002) are rare throughout large periods of the geological
record. In fact, where similar magnitude changes do exist such as during the Neoproterozoic or
the Permian–Triassic extinction (Krull et al., 2004), they have been mainly (but not exclusively)
interpreted as reflecting changes in the global carbon cycle (Hoffman et al., 1998; Macdonald et
al., 2010). As discussed earlier, it is difficult for samples which have experienced only marine
diagenesis to become significantly altered from their original 13C composition. Therefore
level are preferable for chronostratigraphy because they avoid most diagenetic effects. High
alternative mechanism for altering the 13C value of carbonates (Derry, 2010). Despite this
preference for deep-water deposits in chemostratigraphy, there have been numerous examples
of shallow water carbonates apparently unaltered by freshwater in that they appear to have
relatively positive 13C values which can be correlated over wide geographic areas (Saltzman,
2002; Saltzman et al., 2004; Vahrenkamp, 1996). It seems likely that there were only minor
changes in sea level during deposition of these samples, and therefore exposure to freshwater
was limited or arid climates prevailed. Such small changes in sea level still produce variation in
the 13C record (Immenhauser et al., 2003; Immenhauser et al., 2002; Rameil et al., 2012), but
the effects are more subtle than those of the Pleistocene and are often confused with changes
deposition can allow the 13C values of the carbonate to become altered, usually towards more
negative values as 12CO2 is released by respiration and sufficient oxidants are supplied through
diffusion. Such signals could conceivably be associated with well-known events in Earth history,
such as the Palaeocene–Eocene boundary, and might enhance oceanic derived 13C signals in
needed to produce global changes in the oceanic 13C value. Furthermore, marine alteration of
the 13C signal can take place in low-carbonate sediments because the pore water will no
mechanism (Schrag et al., 2013). This idea suggests that significant authigenic carbonate with
isotopically negative 13C values derived from anoxic oxidation of organic matter formed at or
near the sea floor. Hence changes in the 13C values of carbonates would not necessarily be
Because the 13C value of OM is often used to support evidence for changes in the organic
carbon cycle, Oehlert et al. (2012) investigated how the 13C value of the OM was related to the
13C value of the inorganic fraction of the ODP Leg 166 samples as measured by Swart & Eberli
(2005). Oehlert et al. (2012) showed a variable degree of correlation between the 13C values
of the inorganic and organic fraction depending upon the position of the core relative to the
platform margin. Closer to the margin of GBB (ODP Site 1005), where the majority of the OM is
derived from shallow water carbonate organisms such as sea grasses, there was no correlation
between the 13C values of the organic and inorganic fractions. Further away from the margin
the degree of correlation improved as OM with more positive 13C values originating from the
platform top mixed with material with relatively more negative 13C values derived from pelagic
sources (Site 1006). This change in the correlation is once again a feature of mixing between
platform derived materials with relatively positive 13C values in organic and inorganic
components, with isotopically lighter pelagic material. Closer to the platform the correlations
between the inorganic and organic fractions breaks down because the sediments contain a
quite a wide range of 13C values in the OM, but a relatively narrow range of 13C values in the
inorganic fraction, there is no correlation between the 13C value of inorganic material and that
of OM on the platform surface and consequently this is reflected at the proximal sites.
It has been stated that positive correlations between the 13C value of organic and inorganic
carbon in the geological record must reflect changes in the original carbon cycle. Specifically
Knoll et al. (1986) says that: “…no secondary processes are known (or for that matter
conceivable) which always shift the isotopic composition of carbonate and organic carbon in
the same direction at the same rate”. However, not only do correlations between the 13C
values of organic and inorganic carbon arise as a result of the mixing of OM in the depositional
environment, as shown by Oehlert et al. (2012), but different diagenetic processes operating at
the same time can result in processes which produce strong correlations between the 13C of
organic and inorganic carbon. For example, in altered Bahamian carbonates, a strong positive
correlation exists between the 13C value of organic and inorganic carbon (Oehlert & Swart,
2014; Fig. 28). This strong correlation clearly does not mean that the changes are related to the
global carbon cycle. In fact what takes place during sub-aerial exposure is that the carbonates
are altered by meteoric fluids in which the 13C value of the DIC becomes more negative as a
result of the addition of CO2 derived from the oxidation of OM. At the same time, OM from
terrestrial vegetation and algae is added to the semi-consolidated sediment which is in the
process of becoming lithified. In the case of exposed carbonates, such as the Bahamas, this OM
is mainly derived from C3 plants which are approximately 10 to 15‰ more negative than OM
associated with sub-aerial exposure result in both the OM and the carbonate acquiring an
isotopically more negative 13C signal than the original sediment. When a section or core is
taken through a sedimentary sequence which has been altered in this manner, the result would
be an apparent strong positive correlation between the 13C values of organic and inorganic
carbon, a trend which has been interpreted in numerous publications as being original (Knoll et
al., 1986; Rose et al., 2012). While it would not be appropriate to say that all sections which
show a positive correlation between the 13C value of organic and inorganic carbonate are a
result of diagenesis, neither would it be correct to state that there are no secondary processes
which could conceivably alter the 13C value of organic and inorganic carbon in the same
direction.
Sulphur Isotopes
The 34S value of the oceans as measured in evaporite minerals (Claypool et al., 1980), CAS (Gill
et al., 2011) and barite (Griffith & Paytan, 2012) has revealed a secular change in the 34S values
of the oceans. These changes arise as a result of a shift in sulphur between the various
reservoirs in the oceans, evaporites and inorganic minerals such as sulphide bearing minerals.
Over longer time scales the 13C values of unaltered carbonates and the 34S value of evaporites
tend to be inversely correlated (Veizer et al., 1980). Such a correlation may arise because
during periods of increased photosynthesis, leading to increased burial of OM, and hence
during which the 13C values of the carbonate record are more positive oxygen levels in the
atmosphere are higher. This leads to higher rates of weathering of sulphide bearing minerals
which normally tend to have more negative 34S values. However, positive correlations
with oceanic anoxic events (OAEs), such as those of the Cretaceous (Jenkyns, 2010). Under
such circumstances, the positive correlation between 13C and 34S values results from
widespread anoxia in the deeper water column promoted by high rates of organic production,
causing isotopic enrichment in 13C of carbonate formed in the surface oceans. The OM
descending into the deeper ocean caused partial sulphate reduction leading to a residual pool
of sulphate which is isotopically enriched in 34S. This pool is partially upwelled to the surface
oceans returning a positive source of 34S to be subsequently incorporated into the CAS.
One concern about the 34S values of CAS is its integrity during diagenesis. While the 34S
values of carbonates altered during freshwater diagenesis are retained (Gill et al., 2008), during
marine burial the concentration of SO42- in pore water is frequently completely exhausted
during oxidation of OM. Because there is significant isotopic fractionation during bacterial
sulphate reduction (BSR), pore waters become greatly enriched in 34S. Sulphate with these
elevated 34S values can become incorporated into the CAS during carbonate transformation
reactions. As a consequence, the veracity of the 34S signal in CAS from sections experiencing
questioned. The situation can be simply modelled by examining what would happen to the 34S
complete loss of the interstitial sulphate in the pore water as a consequence of BSR. Along this
same pathway the original biogenic sediments are dissolved and inorganic LMC and dolomite
precipitate. Such dissolution and precipitation is common in the majority of DSDP and ODP
sites (Baker, 1985; Gieskes, 1976; Kastner et al., 1990; Kramer et al., 2000). During BSR, 32S is
assumed that 100% of the carbonate is altered to LMC by the time the sulphate is consumed
with a value of 1.04, then each increment of carbonate which is precipitated will have a
progressively more positive 34S value reflecting the modified pore fluid. Using modern values
for oceanic SO42- concentrations (28 mM) and an initial concentration of S in the carbonate of
4000 ppm, then the 34S value of the LMC at the bottom of the sulphate reduction zone would
increase to +29‰, 9‰ more positive than the present oceanic 34S value (Fig. 29). If the SO42-
concentration of the initial carbonate was lower or the value >1.04, then the diagenetic effect
would be greater.
Although the above model indicates that there might be significant potential for alteration of
the 34S value of the CAS signal as a consequence of marine diagenesis, the scenario outlined
above would be a worse case situation, as the model predicts continual carbonate dissolution
and precipitation through the zone of sulphate reduction. If the carbonate was completely
altered to LMC early before BSR took place, then the original 34S value of the CAS would be
preserved. Conversely, if the carbonate survived unaltered below the BSR zone, then the 34S
value of the CAS would also be unaffected. Intermediate situations would also result in a
Although the present concentration of sulphate in the oceans is probably as high as it has ever
been (Demicco et al., 2005), lower initial concentrations would not result in a reduced
diagenetic effect because presumably the initial amount of sulphate in the original carbonate
would also have been reduced. Consider the Carboniferous where the SO42 concentration of
initial SO42 the CAS content of the carbonates precipitated in these oceans would also be 30% of
modern values and therefore the model diagenetic effect would be approximately the same.
The study of diagenesis in carbonate rocks using geochemistry has reached an exciting point.
Not only have the classic tools of C and O isotopes and trace elements continued to improve in
terms of their accuracy, sensitivity and microsampling ability, but there are a number of
additional geochemical proxies and methods of applying these proxies potentially available.
Some of these proxies need to be refined and there are still other proxies waiting to be
discovered. However, one concern is the rush to apply all geochemical techniques to the
interpretation of ancient rocks without adequately understanding the factors controlling the
proxy. The extensive literature documenting diagenetic changes in modern and recent
sediments and rocks would also repay critical perusal. The geochemical signatures of
carbonates are undoubtedly related to the temperature and geochemical conditions of the
environment in which they formed or in which they were altered. This has been confirmed
in the laboratory under controlled conditions. The next logical step of such experiments is to
studies, even to very recently formed carbonates, shows that it is difficult to obtain a clean
environmental signal. This is because, in spite of the large amount of research already
performed, there are still many aspects of carbonate formation (skeletal and non-skeletal)
there be much faith in interpreting the same proxy in 600 Myr old samples? The literature is
replete with cases where geochemical interpretations have been used to support changes in
the original depositional environment rather than diagenetic changes. For example, it was
recognised very early in the application of stable C and O isotopes to carbonates that diagenesis
plays a critical role in governing the ultimate stable isotope composition of the rock. However,
C and O isotopes are frequently applied to rocks which are hundreds to thousands of millions of
years old without recognizing the basic principles affecting the alteration of sediments to
lithified rocks. The same criticism can be levelled at the numerous other geochemical proxies
applied to ancient rocks. Unfortunately it is more ‘news worthy’ to publish studies which
interpret large changes in the geochemical record as extreme events in Earth’s early history,
rather than as a result of more mundane geochemical process such as diagenesis. If the
present or the Neogene is any kind of key to the past, geochemical studies of ancient rocks
combining petrographic and geochemical approaches. So that we do not: “Throw the Baby out
with the Bath Water” (Marshall, 1992), such proxies should be cautiously applied to older
materials. Only then can the potential of geochemical indices be made available to correctly
study both the palaeoenvironment and the diagenesis of carbonates in older time periods.
I would like to thank my graduate students and post-doctoral associates who have worked on
the meaning of trace elements and stable isotopes in carbonates. These include Monica
Arienzo, Steve Burns, Gong Chung, Mike Guzikowski, Genny Healy, Yula Hernawati, Phil Kramer,
Mike McClain, Sevag Mehterian, Leslie Melim, Sean Murray, Amanda Oehlert, Brad Rosenheim,
Victor Rossinsky, Philip Staudigel, Volker Vahrenkamp and Amanda Waite. Discussions with
David Budd, Tracy Frank, Robert Ginsburg, Kacey Lohmann, Martin Kennedy and Gene Shinn
helped with the preparation of this paper. Victor Rossinsky and Leslie Melim are thanked for
the images in Figs 10 and 19. The paper benefited from reviews by Art Saller, Jim Hendry,
Maurice Tucker and Isabel Montanez. Ali Pourmand commented on the section of REEs. Greta
Mackenzie is thanked for drafting Fig. 9 and final reading of the manuscript.
References
Adkins, J.F., Boyle, E.A., Curry, W.B. and Lutringer, A. (2003) Stable isotopes in deep-sea corals and a
new mechanism for "vital effects". Geochim. Cosmochim. Acta, 67, 1129-1143.
Adkins, J.F. and Schrag, D.P. (2003) Reconstructing Last Glacial Maximum bottom water salinities from
deep-sea sediment pore fluid profiles. Earth Planet. Sci. Lett., 216, 109-123.
Affek, H.P., Bar-Matthews, M., Ayalon, A., Matthews, A. and Eiler, J.M. (2008a) Glacial/interglacial
temperature variations in Soreq cave speleothems as recorded by 'clumped isotope' thermometry.
Geochim. Cosmochim. Acta, 72, 5351-5360.
Affek, H.P., Guo, W., Daeron, M. and Eiler, J.M. (2008b) 'Clumped isotopes' in speleothem carbonate
and atmospheric CO2 - Is there a kinetic isotope effect? Geochim. Cosmochim. Acta, 72, A6-A6.
Aissaoui, D.M. (1988) Magnesian calcite cements and their diagenesis- DIssolution and dolomitization,
Mururoa Atoll. Sedimentology, 35, 821-841.
Alibert, C., Kinsley, L., Fallon, S.J., McCulloch, M.T., Berkelmans, R. and McAllister, F. (2003) Source of
trace element variability in Great Barrier Reef corals affected by the Burdekin flood plumes. Geochim.
Cosmochim. Acta, 67, 231-246.
Allan, J.R. and Matthews, R.K. (1977) Carbon and oxygen isotopes as diagenetic and stratigraphic tools:
Surface and subsurface data, Barbados, West Indies. Geology, 5, 16-20.
1- The distribution of 18O values in precipitation as calculated using the IAEA database
(Yurtsever, 1975; Yurtsever & Gat, 1981) and using known relationships between 18O
values and temperature elevation, and distance from coastlines (Bowen & Wilkinson,
2002).
2- (A) Changes in the abundance of carbon species comprising the DIC. (B) Changes in the
13C values of the different species contributing to the DIC assuming that the system
maintains a constant 13C value of the DIC of -5‰. (C) Changes of the 18O value of the
sum of the DIC species as a function of pH as calculated using the approach of Zeebe
(Zeebe, 2007; Zeebe & Wolf-Gladrow, 2001). Fractionation factors are taken from Beck
et al. (2005).
3- Solutions for a range of 18O values for fluids used in the precipitation of LMC (dashed
lines; Kim & O'Neil, 1997) and dolomite (solid lines; Sheppard & Schwarcz, 1970). A LMC
with a measured 18O value of approximately -5‰ could form from fluids with a 18O
value of 0‰ at ca 40oC or +4‰ at ca 65oC. A dolomite with the same 18O value would
need a fluid with 18O values between -4‰ and 0‰ to form at the same temperatures.
4- The Meteoric Water Line (MWL) and the pathway taken during the evaporation of fluids
changes in the activities of H and O with increasing ionic strength, the 18O and 2H
5- Compilation of 18O and 13C values in aragonite cements from Bermuda and Belize,
(Gonzalez & Lohmann, 1985)❶, the Bahamas (Grammer et al., 1993)❷, and HMC
cements from the Pacific (Aissaoui, 1988)❸, HMC cement from Bermuda, and Belize
(Gonzalez & Lohmann, 1985)❹, Bahamian sediments (Swart et al., 2009)❺, Heron
Island sediment (Weber & Woodhead, 1969)❻, Enewetak sediment (Weber & Schmalz,
1968)❼ and zooxanthellate corals (Swart et al., 1996)❽. The boxes represent
6- Compilation of the 18O and 13C values of various types of sediments and corals (from
Fig. 5), deep-sea corals (Emiliani et al., 1978; Land et al., 1977), brachiopods (Carpenter
& Lohmann, 1995), calcareous algae (Lowenstam & Epstein, 1957), brachiopods and
equilibrium of aragonite, LMC and HMC deposited at the surface of Great Bahama Bank.
7- (A) Photomicrograph of boring algae and fungi using fluorescence microscopy in the
skeleton of a recently deceased coral (scale bar = 500 m). (B) Etch transverse section
through the septa of a coral showing boring organisms penetrating the centres of the
trabecular axis (scale bar = 100 m). (C) SEM of an etched thin section of a recently
deceased coral showing the abundance of boring endolithic organisms (scale bar = 100
m).
2005), Pliocene San Salvador (Dawans & Swart, 1988) and Clino (Swart & Melim, 2000)
9- Schematic figure showing the distribution of the different diagenetic zones associated
with early freshwater diagenesis. The pattern of 13C and 18O values is generalized
from the work of Allan & Matthews (1982) as well as this paper.
10- (A) Host Pleistocene aeolinite from Caicos (Bahamas) capped by a thin calcrete
(Rossinsky, 1986). (B) Close up of calcrete in (A) (Image from Rossinsky, 1986). (C)
Peloids from Caicos in the Bahamas cemented by vadose cements and penetrated by
roots from overlying vegetation (Image from Rossinsky, 1986) (scale bar = 500 m). (D)
SEM of vadose cement from Ocean Bight (Bahamas) (Image from McClain et al., 1992).
Clino core at a depth of 131 mbmp. Sample shows dissolution of precursor grains and
preservations of micritic envelopes (scale bar = 500 m). (F) Sample showing the
influence of freshwater alteration of a mudstone in the phreatic zone of the Clino core
(95 mbmp). Although the sample does not necessarily show classic petrographic
indicators of freshwater diagenesis, the 13C and 18O values indicate that this sample
11- Data from diagenetically altered carbonate rocks from the Bahamas show the inverted
‘J’ pattern (Lohmann, 1987). Data are from Rossinsky & Swart (1993) and Rossinsky et
al. (1992).
(Ginsburg, 2001) and Florida Bay (location of Jimmy and Crane Keys).
13- The 18O and 13C values of carbonates (Melim et al., 2001) from a core drilled in the
Bahamas together with the interpretation from the work of Allan & Matthews (1982). In
this core the upper portion was unconsolidated and lost during the drilling. The 13C
values of surface sediments average +4.5‰, but can be as positive as +6‰. Below this
the following zones can be recognised: (1) the vadose zone: The 13C values are highly
variable, while 18O values are negative but more homogenous. Exposure surfaces have
very negative 13C values; (2) the freshwater phreatic zone is characterized by less
variable 13C but similar 18O values to those in the vadose zone; (3) the mixing-zone is a
sediments. More dissolution than precipitation occurs in this zone and 13C and 18O
values co-vary; (4) the marine phreatic zone is characterized by limited changes in both
13C and 18O values. Variations which do exist such as in the example here relate to the
input of different types of sediment. In this case bank top non-skeletal sediments form
the majority of the sediment in this interval. The two inserts (367 m and 536 m) show
two non-depositional surfaces where sea floor diagenesis can alter the signal of the
14- Excesses and deficits of non-conservative elements relative to Cl- at ODP Site 1005 (Fig.
9). Deficits are caused by utilization of the element in reactions such as the oxidation of
organic material which utilizes SO42- and produces alkalinity (Fig. 13).
16- Relationship between Sr2+/Ca2+ and Ca2+/Cl- from ODP Site 1005 (solid circles; Kramer et
al., 2000) and in a core taken from Florida Bay (red diamonds; Swart et al., 1987a; see
Fig. 9 for location). At Site 1005 the change in ratios down core starting in the circle is
represented by the dashed line. The circle represents the approximate starting
LMC and dolomite drives the Ca2+/Cl- ratio down and the Sr2+/Ca2+ ratio up. Calcium ions
diffuse downward from the seawater–sediment interface into the Ca2+ minimum zone
as well upwards from the underlying pore waters where there is dissolution of CaCO3.
17- Relationship between Sr2+/Ca2+ and Mg2+/Cl- from ODP Site 1005 (solid circles; Kramer et
al., 2000) and in a core taken from Florida Bay (red diamonds; Swart et al., 1987a; see
Fig. 4 for location). The circle represents the approximate position of seawater. In Crane
Key dissolution of aragonite occurs as there is an increase in Mg2+/Cl- and Ca2+/Cl- ratio
without a change in the Sr2+/Ca2+ (as the distribution coefficient for aragonite is close to
combined with the precipitation of LMC. At Site 1005 the change in ratios is represented
by the dashed line. In the upper portion of the sediment column, precipitation of LMC
and dolomite drives the Mg2+/Cl- ratio down and the Sr2+/Ca2+ ratio up.
sediments associated with carbonate platforms at DSDP Site 541 (A) and Site 504 (B).
19- Examples of marine diagenesis in platform derived sediments from Leg 166 (Fig. 10).
Scale bar = 500 m. (A) Sample of unconsolidated sediment near the top of Site 1006
(10.57 mbsf). The sediment is composed of pelagic material (foraminifera), molluscs and
shows no evidence of dissolution and cementation. (B) Sample from Site 1003 and a
depth of 140 mbsf, showing some dissolution and precipitation. (C) Samples from Site
1003 (483 mbsf) showing extensive infilling and dissolution. (D) Sample from Site 1003
and a depth of 601 mbsf. (E) Sample from Site 1003 (679 mbsf) showing extensive
dissolution and precipitation and a fracture which has been infilled with celestite. (F)
Sample from Site 1003 (734 mbsf) showing extensive dissolution and precipitation of
20- Changes in pore water chemistry (Ca2+, Mg2+, Sr, 18O and 87Sr/86Sr) in sediments
associated with carbonate platforms at ODP Site 817 (A) and Site 823 (B). The predicted
Sr concentration in diagenetic calcite is based on the Sr/Ca ratio in the pore fluids and a
21- (A) Changes in pore water chemistry (Ca2+, Mg2+, Sr, 18O and 87Sr/86Sr) in pore fluids
associated with carbonate platforms at ODP Sites 1005. The predicted Sr concentration
in diagenetic calcite is based on the Sr/Ca ratio in the pore fluids and a distribution
22- Contour map of the 18Owvalues of the pore fluids from ODP Leg 166 (Swart, 2000). The
18O values of pore waters become more positive with depth as a result of alteration of
23- Modelled (blue line) and measured changes in the 18O values of the pore fluids
(diamonds; Swart, 2000), bulk sediments (black line and black symbols) and diagenetic
carbonate (red line) from ODP Site 1006. Model 18O data were calculated using a model
adapted from Lawrence (1989) by Stout (1985). The pore water data agrees well with
the model with the exception of the increase in the 18Owvalues, interpreted to be a
result of increases in the oceanic 18Owassociated with the last glacial period which has
diffused downward into the sediments (Schrag et al., 2002). Variations in the 18O values
of the sediments falling away from the modelled line are interpreted to be a result of
24- The change in the concentration of Sr (black circles) of the pore fluids (red circles) and
the 87Sr/86Sr ratios of pore fluids and dolomites (diamonds) from ODP Site 812 drilled off
the Queensland plateau (Elderfield et al., 1993; McKenzie et al., 1993). See Fig. 18 for
location of sites.
25- Concentration of Cl- in pore waters in cores drilled adjacent to a continental margin
showing the influence of saline fluids derived from the continent in the pore waters
(Feary et al., 1998). The Cl- concentrations cross cut the sedimentary layers as revealed
The sample has been stained with Alizarin red (Friedman, 1959) to show the presence of
calcite (red). Dolomite is brown. The hydrocarbons react with anhydrite (white) to
produce pyrite (black blobs) and calcite. Typically fringes of calcite are found
surrounding anhydrite. In this example the 13C value of the calcite could not be
measured and the impact upon the bulk 13C value was insignificant.
27- Relationship of the mean 13C values of the sediments in seismic sequences from ODP
Sites 1007, 1006, 1005, 1004, Clino and Unda relative to data from the same sequence in
Site 1003 (Swart & Eberli, 2005). Error bars represent the standard deviation of the data
within each sequence. The data from the ODP sites on a 1:1 line indicating the 13C
values correlate from site to site, the data from Clino and Unda fall off the line because
they have been influenced by meteoric diagenesis, rendering them useless for
chronostratigraphy.
28- The changes in the 13C values of organic and inorganic carbon from Clino (See Fig. 9 for
location). Data from Oehlert & Swart (2014). Note the strong covariation between the
two signals.
29- The potential effect of diagenesis on the 34S values of CAS in a bulk carbonate (solid
circles), assuming that BSR reduces the SO42- concentration in the interstitial pore water
(diamonds) from 28 mM (initial 34S = +22‰) to 0 mM at a depth of 1000 mbsf and that
the carbonate is completely dissolved and reprecipitated over the same interval. It is