You are on page 1of 10

Geochimica et Cosmochimica Acta, Vol. 68, No. 8, pp.

1691–1700, 2004
Copyright © 2004 Elsevier Ltd
Pergamon Printed in the USA. All rights reserved
0016-7037/04 $30.00 ⫹ .00
doi:10.1016/j.gca.2003.10.023

Isotopic order, biogeochemical processes, and earth history


Goldschmidt Lecture, Davos, Switzerland, August 2002
JOHN M. HAYES*
Department of Geology and Geophysics, Woods Hole Oceanographic Institution Woods Hole, MA 02543, USA

Abstract—The impetus to interpret carbon isotopic signals comes from an understanding of isotopic
fractionations imposed by living organisms. That understanding rests in turn on studies of enzymatic isotope
effects, on fruitful concepts of isotopic order, and on studies of the distribution of 13C both between and within
biosynthetic products. In sum, these studies have shown that the isotopic compositions of biological products
are governed by reaction kinetics and by pathways of carbon flow.
Isotopic compositions of individual compounds can indicate specific processes or environments. Examples
include biomarkers which record the isotopic compositions of primary products in aquatic communities, which
indicate that certain bacteria have used methane as a carbon source, and which show that some portions of
marine photic zones have been anaerobic. In such studies, the combination of structural and isotopic lines of
evidence reveals relationships between compounds and leads to process-related thinking. These are large
steps. Reconstruction of the sources and histories of molecular fossils redeems much of the early promise of
organic geochemistry by resolving and clarifying paleoenviron-mental signals. In turn, contemplation of this
new information is driving geochemists to study microbial ecology and evolution, oceanography, and
sedimentology. Copyright © 2004 Elsevier Ltd

1. INTRODUCTION specific organisms, and particular sources of signals (e.g., a


paleoclimatic variation).
My aim is to trace the transformation of organic geochem- There are multiple approaches to decoding. By resolving the
istry into biogeochemistry. It has been an interesting revolution structures of sedimentary molecules in great detail, many work-
with many leaders. I am honored to appear as a representative. ers have been able to constrain their histories. Stratigraphic
Forty years ago, organic geochemists thought mainly about variations also provide useful evidence: molecules whose abun-
petroleum and its sources. In that milieu, Philip Abelson, dances vary together might have a common source. The present
Melvin Calvin, and Geoffrey Eglinton were revolutionaries. discussion will focus on isotopic variations.
Abelson called his research “paleobiochemistry” (Abelson, An isotopic variation does not constitute an interpretable
1954) and then “biogeochemistry” (Abelson and Hoering, signal unless the mechanism controlling it is known. For inor-
1960). Eglinton and Calvin (1967) took organisms and bio- ganic processes, these are usually equilibrium isotope effects
chemistry as the starting points for their interpretations of (Chacko et al., 2001). For organic molecules, the mechanisms
sedimentary molecules. The impact of this new thinking was prominently include kinetic as well as equilibrium isotope
great and the future looked very promising. The development effects (Hayes, 2001). Large and diverse isotopic variations are
of the field was discussed using terms from information theory common. The questions are what causes them and whether the
(Shannon, 1948; Hamming, 1986). Sedimentary organic mol- mechanisms are consistent enough that observed fractionations
ecules were viewed as channels linking modern observers to can be interpreted reliably.
ancient events. The great diversity of molecular structures
meant that the number of channels, and thus the flow of 2. INTRAMOLECULAR PATTERNS OF ISOTOPIC ORDER
information, was practically unlimited. The implied system is
summarized in Figure 1, which is both conceptual and caution- A key contribution was made by Galimov (1973), who
ary. As it illustrates the hoped-for flow of information, it also developed accurate techniques for calculating equilibrium dis-
reminds us that the molecular channels are coupled to remark- tributions of isotopes within organic molecules. Because vibra-
ably diverse sources (in the information-theoretical sense) and tional environments differ significantly between carbon posi-
encoders. They are also affected by an important source of tions, these intramolecular distributions are not expected to be
noise. uniform. In fact, Galimov’s results indicated an astonishing
The tools of modern organic chemical analysis were brought world of isotopic order. Figure 2 provides an arresting example.
to bear and data were immediately plentiful. But data can be Each circle in the diagram at right represents a carbon position
turned into information only by accurate decoding. For organic in a molecule of chlorophyll, with each diameter proportional
geochemists, the essence of decoding is to know not only how to the abundance of 13C expected at isotopic equilibrium. Is
the atoms within the molecule are arranged now, but also to natural chlorophyll really like this? Even today, no analytical
know about their history. Decoding establishes relationships technique can either confirm or refute the prediction. Moreover,
between sedimentary molecules, initial biosynthetic products, that issue is secondary. The shock is that an image this exotic
is required to describe the idea. Conventional depictions of
molecular structure (shown at left in the figure) have no isoto-
* Author to whom correspondence should be addressed pic dimension.
(jhayes@whoi.edu). Until then, conventional thinking about molecular structures
1691
1692 J. M. Hayes

Fig. 1. Schematic view of the role of sedimentary organic molecules


as carriers of information. The elements portrayed here (“Message
source,” etc.) are those specified by information theory (Shannon,
1948). In this rigorous and formal system, the source of a signal is
distinct from the chemical or biological sources of molecules them-
selves.

Fig. 3. Factors affecting the distribution of 13C in n-alkyl chains. (a)


had no isotopic dimension. But suddenly, such isotopic patterns Biosynthetic sources of the C atoms in a typical n-alkanoic acid
(⫽ “fatty acid”). (b) The distribution of 13C expected if isotopic
seemed likely to offer a means of constraining molecular his- equilibrium prevails throughout the biosynthetic process (Galimov,
tories. The task would be to learn what isotopic order was built 1973). (c) The distribution expected if isotopic equilibrium prevails
into molecules during their biosynthesis and then to determine during the biosynthesis of acetate but not during the polymerization of
whether and how those patterns were preserved in the geo- acetate to produce the hydrocarbon chain. (d) The distribution expected
if kinetic factors govern the distribution of 13C in both acetate and the
sphere. final biosynthetic product.
The impact of Galimov’s work was amplified because Abel-
son and Hoering (1961) had already demonstrated that measur-
able isotopic order existed within amino acids. They couldn’t
determine the abundance of 13C at each carbon position, but carbon chains in nature contain two subsets of carbon atoms.
they did show that the carboxyl groups were commonly en- One is derived from the methyl position of the parent acetate,
riched relative to the rest of the molecule. In both sign and the other from the carboxyl position. Since all but the termi-
magnitude, the internal isotopic contrasts were for many amino nal-C positions are equivalent CH2 groups, equilibration calls
acids similar to those predicted by Galimov’s calculations. Was for isotopic homogeneity within the chain (Fig. 3b). Several
this coincidental, or did it mean that thermodynamics, rather alternatives can be considered. The first (Fig. 3c) shows what
than kinetics and biosynthetic pathways, actually controlled the would be expected if the carboxyl position in the acetate
distribution of 13C within—and thus between— organic mole- monomer were enriched in 13C and if that enrichment, rather
cules? than equilibria established during biosynthesis of the final
The clearest answer would come from examination of struc- product, controlled the distribution of isotopes.
turally equivalent positions in a biosynthetic product. Lipids, The second alternative (Fig. 3d) is interesting because it is
particularly the common fatty acids, provide the best example. expected if the distribution of 13C is controlled by kinetic
Biosynthetically, these abundant products derive from the po- factors. DeNiro and Epstein (1977) showed that pyruvate de-
lymerization of acetate. As shown in Figure 3 all unbranched carboxylase, an enzyme similar to that which produces the
acetate biomonomer, has an isotope effect that discriminates
against 13C at the C-2 position in pyruvate. That isotope effect
would act like a dam. Carbon-13 would pile up behind it and be
diverted to alternate fates (Hayes, 2001). The C2 units produced
by pyruvate dehydrogenase would be depleted in 13C at the
carboxyl position. As shown in Figure 4, the magnitudes of the
enrichment and depletion would depend on the branching ratio.

3. DIRECT MEASUREMENTS OF INTRAMOLECULAR


ISOTOPIC ORDER

As a first step in searching for such effects, we learned how


to drive the Schmidt decarboxylation

* O H ⫹ HN ™™™™3 RNH ⫹ C
*O ⫹ N
RC 2 3 2 2 2 (1)
H2SO4
Fig. 2. Two representations of the chemical structure of chlorophyll absolutely to completion so that we could analyze carboxyl
a. On the left, the conventional structural formula. On the right, a
scheme that was invented by Galimov (1974) in order to describe
groups (Vogler and Hayes, 1979, 1980). We also synthesized
position-specific variations in the abundance of 13C. The isotopic carboxylic acids in which the isotopic composition of the
distribution indicated here is that expected at isotopic equilibrium. carboxyl group was known absolutely (Vogler et al., 1978;
Isotopic order, biogeochemical processes, and earth history 1693

Fig. 4. Branching flow of carbon at pyruvate dehydrogenase and its


effect on 13C abundances (DeNiro and Epstein, 1977; Monson and
Hayes, 1982). The scheme on the left summarizes the reactants and
products. A carbon kinetic isotope effect is associated with the action
of pyruvate dehydrogenase. As a result, the carboxyl position in acetyl
coenzyme A is depleted in 13C relative to the carbonyl position in
pyruvate. The graph at right depicts the relationship between the Fig. 5. Measurement of the isotopic compositions at specific posi-
magnitude of that depletion and the branching ratio, expressed in terms tions within hydrocarbon chains (Monson and Hayes, 1982). (a) Chem-
of the fraction of carbon flowing to acetate. ical reactions used to convert positions to carboxyl groups. (b) Mea-
sured isotopic compositions. Additionally, combustion of the parent
acids yielded overall ␦ values. The average result, ⫺12.8‰, does not
differ significantly from the average of the internal methyl (⫺9.5) and
Vogler and Hayes, 1978). Specifically, conventional Grignard internal carboxyl (⫺16.0). The terminal carboxyl positions have been
syntheses (Eqns. 2 and 3) were started with a known excess of fractionated by an additional process (Monson and Hayes, 1980).
CO2,

xRMgX ⫹ CO2 3 xRCO2MgBr ⫹ 共1 ⫺ x兲 CO2 (2) bling these chains and, in the process, quantitatively producing
new carboxyl groups. As shown in Figure 5a, this allowed
multiple, spatially resolved analyses of a parent fatty acid and
xRCO2MgBr ™™™™3 xRCO2H (3) its cleavage products. Carbon positions 10 and 9, highlighted in
H⫹,H2O
the figure and structurally equivalent in the parent acid, derive
the unconsumed CO2 was recovered and analyzed isotopically, respectively from the methyl and carboxyl carbons of an ace-
a mass balance showed that all the starting CO2 was accounted tate group.
for, and the ␦ value of the carboxylic acid was calculated by The results (Fig. 5b) showed that these structurally equiva-
difference. The acids served as standards that demonstrated the lent positions had different isotopic compositions and, there-
adequacy of the optimized Schmidt decarboxylation (Table 1). fore, that isotopic equilibrium did not prevail. As expected on
To analyze carbon positions within hydrocarbon chains, the basis of the isotope effect identified by DeNiro and Epstein
Monson and Hayes (1982) developed methods for disassem- (1977), the carbon position derived from the carboxyl group of
acetate was isotopically depleted. These and all other results
(Monson and Hayes, 1982) were quantitatively consistent with
Table 1. Tests of Schmidt decarboxylation. control of isotopic compositions by kinetic isotope effects at
two sites of fractionation. The first of these was at pyruvate
␦13CPDB, ‰ dehydrogenase, where a 23‰ kinetic isotope effect accounted
Substrate Predicted for carboxyl Found for CO2 Difference for the alternating pattern of isotopic order within hydrocarbon
chains. The second was at fatty acid synthetase, were a kinetic
Acetic acid ⫺44.4 ⫾ 1.6 a
⫺46.1 ⫾ 0.1 ⫺1.7 isotope effect of 13‰ at the branch point between liberation of
Na salt ⫺44.7 ⫾ 0.1 ⫺0.3 a free acid vs. continued elongation accounted for varying
Hexanoic acid ⫺57.6b ⫺57.4 ⫾ 0.5 0.2
Na salt ⫺58.7 ⫾ 0.2 ⫺1.1
isotopic abundances of terminal carboxyl groups (Monson and
Me ester ⫺57.3 ⫾ 0.2 0.3 Hayes, 1980).
Octanoic acid ⫺50.3b ⫺50.7 ⫾ 0.0 ⫺0.4 Particularly in retrospect, this dissection of controlling mech-
Na salt ⫺51.8 ⫾ 0.0 ⫺1.5 anisms has had multiple lines of significance. It provides direct,
Me ester ⫺50.6 ⫾ 0.2 ⫺0.3 quantitative support for the existence of intramolecular isotopic
Na salt ⫺58.2 ⫾ 1.0c ⫺59.1 ⫾ 0.1 ⫺0.9
Nonanoic acid ⫺25.0b ⫺25.2 ⫾ 0.2 ⫺0.2 order and for the concept that isotopic differences between
Na salt ⫺65.2 ⫾ 1.8c ⫺64.6 ⫾ 0.3 0.7 molecules are in fact the attenuated reflections of isotopic
differences within molecules. Together with the demonstration
a
Determined by pyrolysis of the sodium salt (Meinschein et al., that kinetic isotope effects and flows of carbon within reaction
1974).
b
Independently analyzed by the method of von Unruh and Hayes
networks govern isotopic fractionations, it provides a straight-
(1976). forward framework for the interpretation of molecular-isotopic
c
Determined by synthesis (Vogler and Hayes, 1978). signals.
1694 J. M. Hayes

would be expected to fall along a straight line with an intercept


near 25‰ on the ␧P axis. On the right, the data have been
transformed and a strong correlation observed. The transfor-
mation is one in which the variable, b, serves as a proxy for
growth rate. For each point, b is estimated from (25 ⫺ ␧P)ce.
That is, given ␧f ⬇ 25‰, b is equal to the product ␤␮(V/S) in
Eqn. 4. Since ␤ and (V/S) are practically constant for popula-
tions of E. huxleyii and G. oceanica, variations in b must reflect
variations in ␮, the specific growth rate. The observed corre-
lation with concentrations of soluble reactive phosphate need
not indicate that phosphate itself is the growth-limiting nutrient.
Instead, the specific growth rate may be controlled by some
Fig. 6. Main processes by which phytoplankton can assimilate inor- trace micronutrient that covaries with phosphate (Morel et al.,
ganic carbon (Wolf-Gladrow and Riebesell, 1997; Laws et al., 1997; 1991).
Laws, 1998; Badger et al., 1998; Reinfelder et al., 2000; Morel et al., Over long periods of geologic time, where the globally
2002).
averaged fractionation between dissolved inorganic carbon and
algal biomass has varied widely, we have a choice between
4. FRACTIONATION OF CARBON ISOTOPES DURING believing that all cells grew more slowly, or that the globally
PRODUCTION OF ORGANIC MATTER averaged ratio of surface area to volume was higher, or that
concentrations of CO2 were higher. On those scales of time and
The most sophisticated dissection of biochemical reaction averaging, the CO2-related interpretations, in which the in-
networks thus far has been that dealing with the assimilation of creased depletion of 13C in marine organic matter prior to the
carbon dioxide by photosynthetic or chemoautotrophic organ- Miocene is attributed mainly to higher concentrations of CO2,
isms. Studies of land plants, summarized by Farquhar et al. may still be largely correct (Arthur et al., 1985; Hayes et al.,
(1982), showed that isotopic fractionation was largely governed 1999).
by two processes which operate in series. The first, transport of
CO2 from outside the plant to the active site of a carbon-fixing 5. ISOTOPIC SIGNALS OF BIOGEOCHEMICAL
enzyme, has a small isotope effect. The second, fixation (the PROCESSES
bonding of carbon from CO2 to a larger organic molecule),
usually has a much larger isotope effect. The net fractionation Consideration of molecular-isotopic signals leads immedi-
is either large or small, depending which of these processes ately to a focus on biogeochemical processes. These include not
controls the overall rate. More recently, the research team at only production but also a wide variety of aerobic and anaer-
Hawaii has beautifully elucidated the processes governing the obic secondary processes in the water column and sediments.
fractionation of carbon isotopes by algae (Laws et al., 1995, All of these interact to control the molecular and isotopic
2002). The relevant systems are summarized in Figure 6. The composition of total organic carbon (TOC) in sediments. It’s
prominent role of bicarbonate in some of these shows that easy to focus on that end product. Throughout the Phanerozoic,
equilibrium isotope effects can play a role in determining isotopic mass balances allow estimation of the fraction of
isotopic compositions of organic molecules. If there is a re- carbon that’s been buried in organic form (Holser et al., 1988).
versible reaction, if it has an isotope effect, and if there is time For example, from the isotopic compositions of carbonate
for equilibration to occur, it will be a controlling factor. A carbon and organic carbon deposited 30 million years ago, we
treatment of carbonate equilibria incorporating considerations
of mass transport as well as reaction kinetics has recently been
published by Zeebe and Wolf-Gladrow (2001).
For algae which obtain CO2 by passive diffusion, overall
fractionations observed to date can be summarized by this
expression (Popp et al., 1998):

␮ 共V/S兲
␧P ⫽ ␧f ⫺ ␤ (4)
ce
where ␧P is the overall fractionation, ␧f is the isotope effect
associated with carbon fixation, ␤ is a constant, ␮ is the specific
growth rate, V and S are the volume and surface area of the
algal cells, and ce is the concentration of dissolved CO2 exter-
nal to the algal cell. In the modern ocean, the most important
controlling factor seems to be ␮. Its power is seen in Figure 7, Fig. 7. Two plots summarizing the same observations of the frac-
which pertains specifically to the alkenone-producing hapto- tionation of 13C by alkenone-producing algae (Bidigare et al., 1997;
phytes, Emiliana huxleyii and Gephyrocapsa oceanica, thus Laws et al., 2001). On the left, measured values of ␧P are plotted as a
function of 1/ce, where ce is the concentration of dissolved CO2
eliminating size and shape as factors. On the left, the data have external to the algal cells. On the right, the same data have been
been plotted simply in terms of ␧P and ce. If concentrations of transformed so that b, a proxy for growth rate (defined in the figure and
dissolved CO2 were the dominant controlling factor, the results explained in the text) is plotted as a function of nutrient concentrations.
Isotopic order, biogeochemical processes, and earth history 1695

can estimate that 23% of the carbon being sequestered in


sediments was organic and that the rest was accounted for by
carbonates. In such calculations, the surface carbon cycle is a
black box that puts out carbonates and organic materials. Pro-
cesses within the box are not considered. Determining why
23% of the buried carbon was organic, or understanding dif-
ferences between sedimentary basins 30 million years ago,
requires consideration of processes.
Molecular-isotopic techniques make some key details acces-
sible. We can, for example, estimate the isotopic difference
between primary products and the TOC that was finally buried.
In our first investigations of this kind, we used porphyrins—
degradation products of chlorophyll—rather than algal lipids as
the isotopic proxy for primary products (Hayes et al., 1989).
Our aim was to partition the isotopic fractionation between
carbonates and organic carbon into two components. The first, Fig. 9. Isotopic profiles, Greenhorn Limestone (Hayes et al., 1989).
The upward triangles and solid line refer to the rising portion of the
measured by the difference between carbonates and porphyrins,
isotopic excursion as seen in carbonates. The downward triangles and
would reflect fractionations associated with primary production dotted line refer to the falling portion.
and with the formation of carbonate minerals. The second,
measured by the difference between porphyrins and TOC,
would reflect the sum of fractionations associated with second-
1983). Under the guidance of Brian Popp, such influences were
ary processes. The approach is summarized schematically in
stringently minimized in this work by analyzing only non
Figure 8.
luminescent, micritic carbonates (Hayes et al., 1989). Values of
The section investigated was from the Western Interior Sea-
␦roc tend to follow those of ␦carbonate because the dissolved
way in North America. It includes the boundary between the
inorganic carbon is the reactant from which the TOC has been
Cenomanian and Turonian ages and its well-known carbon-
isotopic excursion (Arthur et al., 1988). Depth profiles for produced. They also include the effects of variations in ␧P, the
carbonates, porphyrins, and TOC are shown in Figure 9. As in isotopic fractionation associated with primary production, and
many sedimentary sequences, the 4.1‰ range of variations for ⌬2, the isotopic shift associated with secondary processes,
the organic materials significantly exceeds the 2.3‰ range in namely all of those between primary products and diageneti-
carbonates. cally stabilized TOC. These include fractionations imposed by
The differing ranges indicate that the isotopic composition of the food chain in the water column; by the benthic community;
the organic material is indeed a separate signal that encodes by the sedimentary microbial community; and, depending on
information beyond that carried by the isotopic composition of the sediment’s history, by thermal processes.
carbonates. Values of ␦carbonate vary as a result of changes in It’s useful to think in terms of relationships between precur-
the globally averaged inputs and outputs to the carbon cycle. sors and products. Those associated with primary production
They can also be affected by diagenetic processes (Veizer, are summarized in Figure 10. In the upper graph, the deviations
from slope ⫽ 1 show that the isotopic composition of primary
products did not precisely follow that of dissolved inorganic
carbon. In the lower graph, the same data are recast in terms of
␧1. As values of ␦carbonate increased during the late Cenoma-
nian, values of ␧1 and thus of ␧P dropped and then returned to
(nearly) their starting value. Thanks to the Hawaiians, we know
that the variations must represent some combination of changes
in concentrations of CO2, rates of growth, and algal shape and
size. Of course all may have varied, with partially offsetting
effects. Broadly, however, the signal on the rising side of the
isotopic excursion represents a perturbation in which concen-
trations of CO2 were drawn down and then recovered, in which
rates of growth increased and then returned to their starting
levels, or in which the algal community temporarily was dom-
inated by fatter cells (larger V/S).
A second precursor-product linkage exists between primary
Fig. 8. Schematic representation of the partitioning of the isotopic products (the biomass produced by photosynthetic organisms,
difference between carbonate minerals and total organic carbon into here represented isotopically by the porphyrins) and sedimen-
two components (Hayes et al., 1989). The first, denoted by ␧1, includes tary TOC. The resulting isotopic shift is symbolized by ⌬2 (Fig.
␧P and the relatively constant offsets between carbonate minerals and 11). Remarkably, the slope relating reactants and products is
dissolved CO2 and between biomass and the carbon in porphyrin ring
systems (the heteroaromatic portion of a chlorophyll molecule). The
not one but instead 1.5. As porphyrins are enriched in 13C,
second, ⌬2, is the sum of fractionations associated with food webs, something is enriching the TOC even more strongly. This
diagenesis, and other secondary processes. corresponds to an increase in ⌬2. In this case, the perturbation
1696 J. M. Hayes

values of ⌬2. Increased heterotrophy could result from rising


levels of O2 in the atmosphere and surface waters. But the
isotopic excursion marks an oceanic anoxic event! With only
slightly more confidence than perplexity, we concluded that
“oxygenation of environments other than deep ocean basins
must have increased” (Hayes et al., 1989). Now, stratigraphic,
sedimentological, and geochemical studies by Sageman and
coworkers (Meyers et al., 2001, 2004) have elegantly con-
firmed this interpretation.
Other isotopically distinct biomarkers can serve as indicators
of significant paleoenvironmental conditions or processes. The
first example was provided by Summons and Powell (1986).
They initially recognized that the aromatic-hydrocarbon frac-

Fig. 10. Two views of primary processes, based on the data in Figure
9. Lines and symbols as in Figure 9. (a) Isotopic composition of
porphyrins, a proxy for primary biomass, plotted as a function of the
isotopic composition of micritic carbonates, a proxy for dissolved
inorganic carbon. The nonunit slope shows that the isotopic composi-
tion of the organic product does not simply follow that of the inorganic
precursor. (b) The fractionation between porphyrins and micritic car-
bonates plotted as a function of the isotopic composition of the latter.
Decreases in the fractionation can be attributed to the causes noted at
lower left in the graph. Increases in ␦carbonate indicate increases in the
fraction of carbon buried in organic form.

maximizes near the top of the isotopic excursion and returns to


its preexcursion value at the end of the event.
Most of the carbon loss on the pathway between primary
products and TOC is due to the activities of respiring hetero-
trophs in the marine food chain (Longhurst and Harrison, Fig. 11. Two views of secondary processes, based on the data in
1989). In modern systems, each trophic step produces a 13C Figure 9. Lines and symbols as in Figure 9. (a) Isotopic composition of
enrichment of 0.5–1.5‰ (Yoshii et al., 1999). The magnitude TOC as a function of that of porphyrins, a proxy for primary biomass.
probably depends on the composition of the input (e. g., the As in the relationship summarized in Figure 10, the isotopic composi-
tion of the product (TOC) again does not precisely follow that of its
relative abundances of lipids, carbohydrates, and proteins). precursor (primary biomass). (b) The same variations summarized in
Cretaceous effects may have differed. In any case, increases in terms of ⌬2, which is seen to rise and fall over the course of the isotopic
the intensity of heterotrophic processes are expected to increase excursion.
Isotopic order, biogeochemical processes, and earth history 1697

The isotopic depletion in methanotrophs is doubly interest-


ing. It results in part from isotope effects associated with the
assimilation of methane (Reeburgh et al., 1997). As kinetic
effects, these are unsurprising: the product (biomass) is de-
pleted in the heavy isotope. If the fractionation derived instead
from equilibrium isotope effects, its sign would be reversed,
with the biomass being enriched relative to the methane (Gali-
mov, 1985). For the question of organic 13C control, kinetic or
equilibrium, the relationship between methane and meth-
anotrophs is, therefore, decisively informative: the control is
kinetic (if equilibrium effects prevailed, the methanotrophic
Fig. 12. (a) Isorenieratane, an aromatic polyisoprenoid and (b) isore- products would be enriched in 13C relative to the methane). For
nieratene (Schaeflé et al., 1977), a carotenoid pigment unique to the the other relationship between a C1 carbon source and biomass,
green photosynthetic bacteria, the Chlorobiaceae. namely that between CO2 and plants, the kinetic and equilib-
rium isotope effects both predict depletion, which is of course
tions of several North American Paleozoic oils contained a observed. Indecisive though it is, proponents of equilibrium
prominent series of compounds related to isorenieratane (Fig. and kinetic controls have both pointed to the “fit,” with out-
12a). In turn, it was apparently derived from isorenieratene siders and students being frustrated by the confusion.
(Fig. 12b), for which the only known source was and is the Isotopically depleted biomarkers signaling the recycling of
green photosynthetic bacteria, the Chlorobiaceae. These organ- methane turned up in some of the first sedimentary extracts
isms are obligate anaerobes. They fix carbon and produce examined using continuous-flow techniques (Freeman et al.,
biomass by using energy from photons, but they rely on sulfide as 1990). It was no great surprise; the samples came from the
an electron donor and are killed by exposure to O2. Accordingly, Messel Shale, which formed from the mud of an Eocene
the depositional environments for the source rocks of these oils swamp. The molecular structures were characteristic of aerobic
must have had waters that were both anaerobic (and sulfidic) and methanotrophic bacteria, which are known to thrive in such
sunlit. In large bodies of water, this is a rare condition. settings.
For confirmation, Summons and Powell (1986, 1987) iso- The challenge has been to find biomarkers characteristic of
lated individual compounds, converted their carbon to CO2, the microorganisms catalyzing the anaerobic oxidation of
and measured the abundance of 13C. The aromatic polyisopre- methane at the expense of sulfate. Depth profiles of the con-
noids were enriched in 13C relative to saturated hydrocarbons centrations of methane and sulfate in sedimentary pore waters
in the same oils by as much as eight permil. That isotopic signal made it perfectly clear that this process was occurring (Iversen
firmly associates these compounds with the Chlorobiaceae, and Jørgensen, 1985). Success eventually came to Bian (1994),
which are unique in their reliance on the reversed tricarboxylic who found isotopically depleted crocetane (Fig. 13) in samples
acid cycle as a means of fixing carbon. Isotope effects within from transition-zone muds provided by Niels Iversen and Bo
that cycle are much smaller than those in other pathways of
carbon fixation (Hayes, 2001). Moreover, lipids can be en-
riched in 13C relative to biomass (van der Meer et al., 1998).
In subsequent work, the group led by Jaap Sinninghe Damsté
and Stefan Schouten at the Royal Netherlands Institute for Sea
Research has developed improved methods for the detection of
these important indicators of water-column anoxia (Koopmans
et al., 1996). They are notably abundant, for example, in North
Atlantic sediments from the Cretaceous anoxic events (Sin-
ninghe Damsté and Köster, 1998). The presence of products
from fastidious anaerobes in open-marine sediments is notable.
Extraordinary levels of stratification are required. These may
have been stabilized—at least well enough to allow transient
growth of the Chlorobiaceae and generation of the molecular-
isotopic signal— by the estimated sea-surface temperatures of
35 to 38°C (which derive from a new, molecular-organic tem-
perature indicator; Schouten et al., 2002).
Molecular-isotopic techniques have also been useful in stud-
ies of methane cycling. This is because, as a result of isotope
effects associated with its production by either biological or Fig. 13. Depth profiles of porewater sulfate and methane, the rate of
thermal processes, methane is commonly depleted in 13C oxidation of methane, and the 13C content of the isoprenoid hydrocar-
(Whiticar, 1994). The offset is so large that, for practical bon, crocetane in cores raised from the seafloor of the Kattegat (Bian
purposes, methane carbon is labeled almost as effectively as et al., 2001). Complementary depth profiles of sulfate, methane, and
methane oxidation were first observed by Iversen and Jørgensen
many artificial tracers. Values of ␦ for lipids from meth- (1985). The role of crocetane as biomarker for methane-consuming
anotrophs frequently drop below ⫺100‰ (Hinrichs et al., anaerobes was discovered independently by Bian (1994) and by Thiel
2000). et al. (1999).
1698 J. M. Hayes

carbon cycle over the course of earth history. The young


scientists pursuing these quests are progressing rapidly. Some
are certain to be recognized as fitting representatives of V. M.
Goldschmidt’s scientific legacy. I do not claim to reach their
level, still less Goldschmidt’s. I have, however, helped to build
the interdisciplinary connections that now characterize biogeo-
chemistry. Soon to be surpassed, I nevertheless have confi-
dence in the value of that role.

Acknowledgments—My work has been supported by the programs in


Exobiology and Astrobiology of the National Aeronautics and Space
Administration and by the National Science Foundation. Most of the
work described was completed while I was a member of the faculties
in Chemistry and in Geological Sciences at Indiana University, Bloom-
ington. I am grateful to that institution and to my colleagues there,
particularly Stephen A. Studley, who worked with me on the develop-
Fig. 14. A gas chromatogram of lipid biomarkers extracted from ment of mass spectrometric techniques from 1972 until I moved to
sediments at gas seeps in the Santa Barbara Basin (Hinrichs et al., Woods Hole in 1996. This published version of my lecture has bene-
2000). The peaks marked in light gray represent products derived fited from constructive comments provided by Jaap Sinninghe Damsté,
exclusively from producers and consumers in the oceanic water col- David Des Marais, and an anonymous reviewer.
umn. Those marked in darker gray represent monoalkyl-(M) and dial-
kyl-(D) glycerol ethers produced by sulfate-reducing bacteria within REFERENCES
the methane-consuming consortia. The peaks marked with a heavy
black line are products of methanotrophic archaea. Abelson P. H. (1954) Paleobiochemistry. Carnegie Inst. Washington
Yrbk. 53, 97–101.
Abelson P. H. and Hoering T. C. (1960) The biogeochemistry of the
Jørgensen. Crocetane and other products of anaerobic methan- stable isotopes of carbon. Carnegie Inst. Washington Yrbk 59, 158 –
otrophy were discovered independently by Michaelis, Thiel, 165.
and coworkers (Thiel et al., 1999). Ether-linked lipids identical Abelson P. H. and Hoering T. C. (1961) Carbon isotope fractionation
in formation of amino acids by photosynthetic organisms. Proc. Nat.
to those in some methanogens, but with isotopic compositions Acad. Sci. USA 47, 623– 632.
indicating derivation from methanotrophs, were discovered by Arthur M. A., Dean W. E., and Claypool G. E. (1985) Anomalous 13C
Hinrichs, DeLong, and coworkers (Hinrichs et al., 1999). enrichment in modern marine organic carbon. Nature 315, 216 –218.
Much additional information about lipids from methanotro- Arthur M. A., Dean W. E., and Pratt L. M. (1988) Geochemical and
climatic effects of increased marine organic carbon burial at the
phic communities has now accumulated. The Hamburg group Cenomanian/Turonian boundary. Nature 335, 714 –717.
was the first to show that 13C-depleted products probably from Badger M. R., Andrews T. J., Whitney S. M., Ludwig M., Yellowlees
sulfate-reducing bacteria were associated with those from the C., Leggat W., and Price G. D. (1998) The diversity and coevolution
methane-consuming archaea (Thiel et al., 1999). A more spe- of Rubisco, plastids, pyrenoids and chloroplast-based CO2 concen-
cific connection to sulfate reducers was established by Pancost trations. Limnol. Oceanogr. 46, 1378 –1391.
Bian L. (1994) Isotopic biogeochemistry of individual compounds in a
et al. (2000). Additional results from Hinrichs et al. (2000), modern coastal marine sediment (Kattegat, Denmark and Sweden).
which serve to summarize these findings, are shown in Figure M.S. thesis, Department of Geological Sciences, Indiana University.
14. Important further reports have followed (Pancost et al., Bian L., Hinrichs K.-U., Xie T., Brassell S. C., Iversen N., Fossing H.,
2001a, b; Thiel et al., 2001a, b). Jørgensen B. B., and Hayes J. M. (2001) Algal and archaeal poly-
isoprenoids in a recent marine sediment: Molecular-isotopic evi-
At the time of its publication, the initial report from Hinrichs dence for anaerobic oxidation of methane. Geochem. Geophys. Geo-
et al. (2000) was unique in its combination of molecular- syst. 2, 2000GC000112.
isotopic and genomic lines of evidence. The genomic approach Bidigare R. R., Fluegge A., Freeman K. H., Hanson K. L., Hayes J. M.,
was promptly extended by Boetius et al. (2000) at the Max Hollander D., Jasper J. P., King L. L., Laws E. A., Millero F. J.,
Planck Institute for Marine Microbiology, who used fluores- Pancost R., Popp B. N., Steinberg P. A., and Wakeham S. G. (1997)
Consistent fractionation of 13C in nature and in the laboratory:
cently labeled oligonucleotide probes to show that archaea Growth-rate effects in some haptophyte algae. Glob. Biogeochem.
closely related to those found by Hinrichs et al. (1999) were Cycl. 11, 279 –292.
closely associated with bacteria closely related to known sul- Boetius A., Ravenschlag K., Schubert C. J., Rickert D., Widdel F.,
fate reducers. This provided support for key aspects of the Gieseke A., Amann R., Jørgensen B. B., Witte U., and Pfannkuche
O. (2000) A marine microbial consortium apparently mediating
“consortium hypothesis” introduced by Hoehler et al. (1994). anaerobic oxidation of methane. Nature 407, 623– 626.
Further multidisciplinary investigations have confirmed that the Chacko T., Cole D. R., and Horita J. (2001) Equilibrium oxygen,
aggregates observed by Boetius et al. (2000) are indeed meth- hydrogen and carbon isotope fractionation factors applicable to
anotrophic (Orphan et al., 2001b) and revealed considerable geologic systems. In Reviews in Mineralogy and Geochemistry:
diversity in anaerobic methanotrophic communities (Orphan et Stable Isotope Geochemistry, Vol. 43 (eds. J. W. Valley and D. R.
Cole). Mineralogical Society of America.
al., 2001a, 2002). DeNiro M. J. and Epstein S. (1977) Mechanism of carbon isotope
These most recent studies illustrate the convergence of three fractionation associated with lipid synthesis. Science 197, 261–263.
lines of inquiry: studies of lipid biomarkers (isotopic and struc- Eglinton G. and Calvin M. (1967) Chemical fossils. Sci. Am. 216,
tural), microbial physiology and ecology, and microbial 32– 43.
Farquhar G. D., O’Leary M. H., and Berry J. A. (1982) On the
genomics. This synergy promises to revolutionize, first, our relationship between carbon isotope discrimination and the intercel-
understanding of terminal processes in the modern carbon cycle lular carbon dioxide concentration in leaves. Aust. J. Plant Physiol.
and, ultimately, our understanding of the evolution of the 9, 121–137.
Isotopic order, biogeochemical processes, and earth history 1699

Freeman K. H., Hayes J. M., Trendel J.-M., and Albrecht P. (1990) CO2 mechanisms, and implications for paleoreconstructions. Funct.
Evidence from carbon isotope measurements for diverse origins of Plant Biol. 29, 323–333.
sedimentary hydrocarbons. Nature 343, 254 –256. Longhurst A. R. and Harrison W. G. (1989) The biological pump:
Galimov E. M. (1973) Izotopy ugleroda v neftegazovoy geologii. Nedra Profiles of plankton production and consumption in the upper ocean.
Press, Moscow. Translated and reprinted in English as TT F-682, Prog. Oceanog 22, 47–123.
Carbon Isotopes in Oil-Gas Geology, National Aeronautics and Meinschein W. G., Rinaldi G. G. L., Hayes J. M., and Schoeller D. A.
Space Administration, Washington, 1974. (1974) Intramolecular isotopic order in biologically produced acetic
Galimov E. M. (1974) Organic geochemistry of carbon isotopes. In acid. Biomed. Mass Spec. 1, 172–175.
Advances in Organic Geochemistry 1973 (eds. B. Tissot and F. Meyers S. R., Sageman B. B., and Hinnov L. A. (2001) Integrated
Bienner), pp. 439 – 452. Editions Technip, Paris. quantitative stratigraphy of the Cenomanian-Turonian Bridge Creek
Galimov E. M. (1985) The Biological Fractionation of Isotopes. Aca- Limestone Member using evolutive harmonic analysis and strati-
demic Press. English translation by D. B. Vitaliano of Priroda graphic modeling. J. Sediment. Res. 71, 628 – 644.
biologicheskogo fraktsionirovaniia izotopov, Nauka, Moscow, 1981. Meyers S. R., Sageman B. B. and Lyons T. W. (2004) The role of
Hamming R. W. (1986) Coding and Information Theory 2nd ed. sulfate reduction in organic matter degradation and molybdenum
Prentice Hall. accumulation: Theoretical framework and application to a Creta-
Hayes J. M. (2001) Fractionation of the isotopes of carbon and hydro- ceous organic matter burial event. Paleoceanography, in press.
gen in biosynthetic processes. In Stable Isotope Geochemistry, Re- Monson K. D. and Hayes J. M. (1980) Biosynthetic control of the
views in Mineralogy and Geochemistry (eds. J. W. Valley and D. R. natural abundance of carbon 13 at specific positions within fatty
Cole), 43, 225–278. Mineralogical Society of America. [Note: this acids in Escherichia coli, evidence regarding the coupling of fatty
pdf is made available with the explicit permission of the publisher–it acid and phospholipid synthesis. J. Biol. Chem. 255, 11435–11441.
is not an illegal posting]: pdf available at http://www.nosams. Monson K. D. and Hayes J. M. (1982) Carbon isotopic fractionation in
whoi.edu/general/jmh/index.html the biosynthesis of bacterial fatty acids: Ozonolysis of unsaturated
Hayes J. M., Popp B. N., Takigiku R., and Johnson M. W. (1989) An fatty acids as a means of determining the intramolecular distribution
isotopic study of biogeochemical relationships between carbonates of carbon isotopes. Geochim. Cosmochim. Acta 46, 139 –149.
and organic carbon in the Greenhorn Formation. Geochim. Cosmo- Morel F. M. M., Hudson R. J. M., and Price N. M. (1991) Limitation
chim. Acta 53, 2961–2972. of productivity by trace metals in the sea. Limnol. Oceanogr. 36,
Hayes J. M., Strauss H., and Kaufman A. J. (1999) The abundance of 1742–1755.
13
C in marine organic matter and isotopic fractionation in the global Morel F. M. M., Cox E. H., Kraepiel A. M. L., Lane T. W., Milligan
biogeochemical cycle of carbon during the past 800 Ma. Chem. Geol. A. J., Schaperdoth I., Reinfelder J. R., and Tortell P. D. (2002)
161, 103–125. Acquisition of inorganic carbon by the marine diatom Thalassiosira
Hinrichs K.-U., Hayes J. M., Sylva S. P., Brewer P. G., and DeLong weissflogii. Funct. Plant Biol. 29, 301–308.
E. F. (1999) Methane-consuming archaebacteria in marine sedi- Orphan V. J., Hinrichs K.-U., Ussler W., III, Paull C. K., Taylor L. T.,
Sylva S., Hayes J. M., and Delong E. F. (2001a) Comparative
ments. Nature 398, 802– 805.
analysis of methane-oxidizing archaea and sulfate-reducing bacteria
Hinrichs K.-U., Summons R. E., Orphan V., Sylva S. P., and Hayes
in anoxic marine sediments. Appl. Environ. Microbiol. 67, 1922–
J. M. (2000) Molecular and isotopic analyses of anaerobic methane-
1934.
oxidizing communities in marine sediments. Org. Geochem. 31,
Orphan V. J., House C. H., Hinrichs K.-U., and McKeegan K. D.
1685–1701.
(2001b) Methane-consuming archaea revealed by directly coupled
Hoehler T. M., Alperin M. J., Albert D. B., and Martens C. S. (1994)
isotopic and phylogenetic analysis. Science 293, 484 – 487.
Field and laboratory studies of methane oxidation in an anoxic
Orphan V. J., House C. H., Hinrichs K.-U., McKeegan K. D., and
marine sediment: Evidence for a methanogen-sulfate reducer con-
DeLong E. F. (2002) Multiple archael groups mediate methane
sortium. Glob. Biogeochem. Cycl. 8, 451– 463. oxidation in anoxic cold seep sediments. Proc. Natl. Acad. Sci. USA
Holser W. T., Schidlowski M., Mackenzie F. T., and Maynard J. B. 99, 7663–7668.
(1988) Geochemical cycles of carbon and sulfur. In Chemical Cycles Pancost R. D., Sinninghe Damsté J. S., de Lint S., van der Maarel
in the Evolution of the Earth (eds. C. B. Gregor, R. M. Garrels, F. T. M. J. E. C., Gottschal J. C., and the Medinaut Shipboard Scientific
Mackenzie, and J. B. Maynard), pp. 105–173. J. Wiley. Party. (2000) Biomarker evidence for widespread anaerobic methane
Iversen N. and Jørgensen B. B. (1985) Anaerobic methane oxidation oxidation in Mediterranean sediments by a consortium of methano-
rate at the sulfate-methane transition in marine sediments from genic archaea and bacteria. Appl. Environ. Microbiol. 66, 1126 –
Kattegat and Skagerrak (Denmark). Limnol. Oceanogr. 30, 944 – 1132.
955. Pancost R. D., Bouloubassi I., Aloisi G., Sinninghe Damsté J. S., and
Koopmans M. P., Köster J., van Kaam-Peters H. M. E., Kenig F., the Medinaut Shipboard Scientific Party. (2001a) Three series of
Schouten S., Hartgers W. A., de Leeuw J. W., and Sinninghe Damsté non-isoprenoidal dialkyl glycerol diethers in cold-seep carbonate
J. S. (1996) Diagenetic and catagenetic products of isorenieratene: crusts. Org. Geochem. 32, 695–707.
Molecular indicators for photic zone anoxia. Geochim. Cosmochim. Pancost R. D., Hopmans E. C., Sinninghe Damsté J. S., and the
Acta 60, 4467– 4496. Medinaut Shipboard Scientific Party. (2001b) Archaeal lipids in
Laws E. A. (1998) Sources of inorganic carbon for marine microalgal Mediterranean cold seeps: Molecular proxies for anaerobic methane
photosynthesis: A reassessment of ␦13C data from batch culture oxidation. Geochim. Cosmochim. Acta 65, 1611–1627.
studies of Thalassiosira pseudonana and Emiliania huxleyi. Limnol. Popp B. N., Laws E. A., Bidigare R. R., Dore J. E., Hanson K. L., and
Oceanogn. 43, 136 –142. Wakeham S. G. (1998) Effect of phytoplankton cell geometry on
Laws E. A., Popp B. N., Bidigare R. R., Kennicutt M. C., and Macko carbon isotopic fractionation. Geochim. Cosmochim. Acta 62, 69 –
S. A. (1995) Dependence of phytoplankton carbon isotopic compo- 77.
sition on growth rate and [CO2]aq: Theoretical considerations and Reeburgh W. S., Hirsch A. I., Sansone F. J., Popp B. N., and Rust T. M.
experimental results. Geochim. Cosmochim. Acta 59, 1131–1138. (1997) Carbon kinetic isotope effect accompanying microbial oxi-
Laws E. A., Bidigare R. R., and Popp B. N. (1997) Effect of growth dation of methane in boreal forest soils. Geochim. Cosmochim. Acta
rate and CO2 concentration on carbon isotopic fractionation by the 61, 4761– 4767.
marine diatom Phaeodactylum tricomutum. Limnol. Oceanogr. 42, Reinfelder J. R., Kraepiel A. M. L., and Morel F. M. M. (2000)
1552–1560. Unicellular C4 photosynthesis in a marine diatom. Nature 407,
Laws E. A., Popp B. N., Bidigare R. R., Riebesell U., Burkhardt S., and 996 –999.
Wakeham S. G. (2001) Controls on the molecular distribution and Schaeflé J., Ludwig B., Albrecht P., and Ourisson G. (1977) Hydro-
carbon isotopic composition of alkenones in certain haptophyte carbures aromatiques d’origine geologique. II. Nouveaux carote-
algae. Geochem. Geophy. Geosyst. 2, 2000GC000057. noı̈des aromatiques fossiles. Tetrahedron Lett. 41, 3673–3676.
Laws E. A., Popp B. N., Cassar N., and Tanimoto J. (2002) 13C Schouten S., Hopmans E. C., Schefuss E., and Sinninghe Damsté J. S.
discrimination patterns in oceanic phytoplankton: Likely influence of (2002) Distributional variations in marine crenarchaeotal membrane
1700 J. M. Hayes

lipids: a new tool for reconstructing ancient sea water temperatures? Vogler E. A. and Hayes J. M. (1978) The synthesis of carboxylic acids
Earth Planet. Sci. Lett. 204, 265–274. with carboxyl carbons of precisely known stable isotopic composi-
Shannon C. E. (1948) A mathematical theory of communication. Bell tion. Int. J. Appl. Radiat. Isotopes 29, 297–300.
System Technical J. 27, 379 – 423 623– 656. Vogler E. A., Stein R. L., and Hayes J. M. (1978) Mechanism of
Sinninghe Damsté J. S. and Köster J. (1998) A euxinic southern North formation of Grignard reagents. J. Am. Chem. Soc. 100, 3163–
Atlantic Ocean during the Cenomanian/Turonian oceanic anoxic 3166.
event. Earth Planet. Sci. Lett. 158, 165–173. Vogler E. A. and Hayes J. M. (1979) Carbon isotopic fractionation in
Summons R. E. and Powell T. G. (1986) Chlorobiaceae in Palaeozoic the Schmidt decarboxylation: Evidence for two pathways to prod-
seas revealed by biological markers, isotopes, and geology. Nature ucts. J. Org. Chem. 44, 3682–3686.
319, 763–765. Vogler E. A. and Hayes J. M. (1980) Carbon isotopic compositions of
Summons R. E. and Powell T. G. (1987) Identification of aryl isopre- carboxyl groups of biosynthesized fatty acids. In Progress in Or-
noids in source rocks and crude oils: Biological markers for the ganic Geochemistry 1979 (eds. A. G. Douglas and J. R. Maxwell),
green sulfur bacteria. Geochim. Cosmochim. Acta 51, 557–566. pp. 597–704. Pergamon.
Thiel V., Peckmann J., Seifert R., Wehrung P., Reitner J., and Michae-
von Unruh G. E. and Hayes J. M. (1976) Intramolecular distribution of
lis W. (1999) Highly isotopically depleted isoprenoids: Molecular
carbon isotopes in fatty acids. In Proceedings of the Second Inter-
markers for ancient methane venting. Geochim. Cosmochim. Acta
national Conference on Stable Isotopes (eds. P. D. Klein and E. R.
63, 3959 –3966.
Thiel V., Peckmann J., Richnow H. H., Luth U., Reitner J., and Klein), pp. 561–568. US ERDA, CONF-751027.
Michaelis W. (2001a) Molecular signals for anaerobic methane Whiticar M. J. (1994) Correlation of natural gases with their sources. In
oxidation in Black Sea seep carbonates and a microbial mat. Mar. The Petroleum System—From Source to Trap, pp. 261–283. AAPG
Chem. 73, 97–112. Memoir 60.
Thiel V., Peckmann J., Schmale O., Reitner J., and Michaelis W. Wolf-Gladrow D. and Riebesell U. (1997) Diffusion and reactions in
(2001b) A new straight-chain hydrocarbon biomarker associated the vicinity of plankton: A refined model for inorganic carbon
with anaerobic methane cycling. Org. Geochem. 32, 1019 –1023. transport. Mar. Chem. 59, 17–34.
van der Meer M. T. J., Schouten S., and Sinninghe Damsté J. S. (1998) Yoshii K., Melnik N. G., Timoshkin O. A., Bondarenko N. A.,
The effect of the reversed tricarboxylic acid cycle on the 13C con- Anoshko P. N., Yoshioka T., and Wada E. (1999) Stable isotope
tents of bacterial lipids. Org. Geochem. 28, 527–533. analyses of the pelagic food web in Lake Baikal. Limnol. Oceanogr.
Veizer J. (1983) Trace elements and isotopes in sedimentary carbon- 44, 502–511.
ates. In Carbonates: Mineralogy and Geochemistry (ed. R. J. Zeebe R. E., Wolf-Gladrow D. (2001) CO2 in Seawater: Equilibrium,
Reeder), pp. 265–393. Mineralogical Society of America. Kinetics, Isotopes. Elsevier.

You might also like