You are on page 1of 11

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/265088122

Sediments, Diagenesis, and Sedimentary Rocks

Book January 2004

CITATIONS READS

33 233

1 author:

F. T. Mackenzie
University of Hawaii at Mnoa
357 PUBLICATIONS 15,944 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Carbon dioxide-carbonic acid system in seawater View project

All content following this page was uploaded by F. T. Mackenzie on 24 January 2017.

The user has requested enhancement of the downloaded file.


VOLUME EDITORS INTRODUCTION

This volume is dedicated to Robert M. Garrels in recognition of his contributions to our understanding of the
fundamental geochemical processes governing the formation and transformation of sedimentary rocks through
geologic time and their use in deducing the evolution of the Earths exogenic system. In retrospect, Bobs
research career can probably be divided into two main phases. The first dealt with theoretical and experimental
aspects of geochemistry that led to the publication of his book Mineral Equilibria at Low Temperature and Pressure
in 1960 and later to the books successor Solutions, Minerals, and Equilibria coauthored with Charles Christ in
1965. These books have become classics in the field of aqueous and sedimentary geochemistry, as they were
one of the first efforts to apply chemical thermodynamics to geological processes with an emphasis on the
construction of stability diagrams from thermodynamic data. The second phase of Bobs career led him to
consider the importance of the chemical cycling of sediments and the processes involved of weathering,
erosion, transport, sedimentation, burial, diagenesis, and uplift restarting the sedimentary cycle. This recycling
concept had lain dormant for 150 years before the publication of Evolution of Sedimentary Rocks with Fred
Mackenzie in 1971. The chapters in this volume largely address this phase of Bobs career.
Biogeochemical processes and the products of those processes, the sedimentary rocks and their geochemical
features are discussed in this second edition of the original Volume 7 of the Treatise on Geochemistry, as is the use
of various geochemical and isotopic proxies found in sedimentary rocks to constrain the chemical evolution of
the Earths surface environment. Most important sedimentary rock facies are considered in separate chapters,
particularly in terms of the evolution of sedimentary rocks and their usefulness in deciphering Earth history.
Sediments at their origin have measurable properties such as mineralogy, chemistry, isotopic composition of
organic and inorganic phases, biological make-up, and texture. The diagenetic processes of compaction,
cementation, and solution and reprecipitation leading to mineral stabilization can modify all of these prop-
erties. Thus as sediments pass through the diagenetic fence (Chave, 1960), they lose information necessary for
environmental interpretation. In light of this, both the geochemistry of modern sediment types and their pore
waters and the diagenetic products of these sediments, the sedimentary rocks and their associated interstitial
fluids, are discussed in this volume. The discussion of processes and products is tied together in most chapters
to the chronological and historical aspects of sediments and sedimentary rocks and the use of biogeochemical
and isotopic proxies within them in interpreting the history of Earths surface environment.
In this Volume 9 of the Treatise, there are 19 chapters, 5 of which are new and 8 significantly revised. The
remaining eight chapters are reprints of material developed for the first edition and for which the authors in
general did not feel required revision. In Chapter 9.1 (reprint), Chemical Composition and Mineralogy of Marine
Sediments, Telu Li and Jane Schoonmaker review the chemical composition of modern marine sediments,
building on the classical works of Tizzard et al. (1885), Murray and Menard (1891), and of more recent
authors (e.g., Arrhenius, 1963). Li and Schoonmaker address the mineralogy and geochemistry of pelagic
sediments in general and consider several specific sediment types, particularly in terms of their hydrogenous
components. Sediment elemental enrichment factors and factor analysis are used to determine the major
components and mineralogical phases in which the elements of marine sediments reside. The major compo-
nents of marine pelagic sediments are shown to be aluminosilicates derived mainly from the weathering of
shales and eolian and riverine transport of the weathered products to the ocean, and ferromanganese oxides,
carbonate fluorapatite, zeolites, biogenic carbonate, silica, and barite derived from the products of organic
productivity and chemical reactions involving elements directly derived from the general background of
elements in seawater or resulting from diagenetic reactions in the oxic, sub-oxic, and anaerobic realms of
early diagenesis (Chester, 1990).

xxiii
xxiv Volume Editors Introduction

Bill Martin and Fred Sayles in Chapter 9.2 (revised), The Recycling of Biogenic Material at the Sea Floor,
consider the biogenic component of the rain of particles to the sea floor, consisting of organic matter, calcium
carbonate, and biogenic silica, and show they are efficiently recycled at the surface of marine sediments. The
chapter deals with the early diagenetic reactions occurring in sediments, including those involving organic
matter degradation, calcium carbonate dissolution, and silica dissolution/precipitation, and the role of bio-
turbation in mixing particles downward from the sedimentwater interface. The reactions these materials
undergo affect Earth-surface biogeochemical cycles on both short and long time scales. Furthermore, they
alter the composition of the particulate flux, so that their effects must be considered in the interpretation of the
sedimentary record. Martin and Sayles primarily use the wealth of concentration versus depth profiles of pore
water solutes that they themselves (e.g., Martin and Sayles, 2003, 2006; Martin et al., 2000; Sayles, 1979, 1981),
as well as other investigators (e.g., Archer et al., 1989; Bender et al., 1989) have obtained during the past several
decades to interpret diagenetic processes. Mass balances provided by the authors show that only a small fraction
of the organic matter and the opal that falls to the sea floor survives early diagenesis; thus, only a small portion
accumulates in the sediment and is available for downcore interpretations of accumulation rates. The incor-
poration of aluminum into opal and the hypothesized reprecipitation of H4SiO4 from dissolving opal into
neoformed minerals highlight the significant influence of the silica cycle on the cycling behavior of the major
constituents of pore water and the likelihood that reverse weathering (e.g., Mackenzie and Garrels, 1966;
Mackenzie and Kump, 1995; Michalopoulos and Aller, 1995) is an important process affecting ocean chem-
istry. For calcium carbonate, the role of dissolution in sediments above the calcite lysocline driven by acids
generated by oxic respiration is highlighted and brings into question present interpretations of the temporal
record of CaCO3 accumulation rates. The dissolution of reactive CaCO3 phases in shallow waters, like reef
environments, is also discussed. Martin and Sayles furthermore demonstrate that about half of the O2
consumption in the ocean below 1000 m occurs in deep-sea sediments, and that sedimentary denitrification
is a dominant term in the marine fixed nitrogen budget. As an aside, the authors point out that pore water
studies probably began with Murray and Irvines work in 1885 (Murray and Irvine, 1885).
The discussion of the early diagenetic processes occurring in carbonate sediments begun in Chapter 9.2 is
continued and expanded on in Chapter 9.3, Formation and Diagenesis of Carbonate Sediments, by Rolf Arvidson
and John Morse. This chapter is based on John Morses previous chapter in Volume 7 of the first edition of the
Treatise but due to Johns untimely death, the revision and expansion of the chapter are largely due to Rolf
Arvidson. Aside from discussion of the sources and early diagenesis of deep-sea and shoal-water (shallow-
water) carbonate sediments discussed in Johns initial chapter, there is greater emphasis on carbonate mineral
surface reactions, including crystal growth, dissolution, and reaction rates in this revised chapter. Atomic force
microscopy, interferometry, and thermodynamic and kinetic modeling are highlighted in terms of bringing
new insights into the behavior of carbonate phases in the natural environment. One of the long-standing
controversial areas of carbonate geochemistry has been the relationship between calcium carbonate accumu-
lation in deep-sea sediments and the saturation state of the overlying waters. Hypotheses considered have
ranged from a nearly thermodynamic ocean where the calcite and aragonite compensation depths are at the
calcite and aragonite saturation horizons (e.g., Li et al., 1969) to a strongly kinetically controlled ocean system
where major differences exist between the carbonate compensation depths, lysoclines, and saturation depths
(e.g., Morse and Berner, 1972). Arvidson and Morse point out that if recent calculations related to these features
are correct, the long cherished idea of a tight coupling between seawater chemistry and carbonate depositional
facies requires reexamination. That such a coupling may not exist has been suggested by Milliman et al. (1999)
who evaluated several lines of evidence and concluded that considerable dissolution of calcium carbonate,
perhaps as much as 6080%, occurs in the upper 5001000 m of the ocean, well above the chemical lysocline.
The role of organic matter in this water-column dissolution process and in microbially driven organic oxidation
processes in both shoal-water and deep-sea sediments is now well recognized. Arvidson and Morse also
consider what is known about early dolomite formation. They point out that most modern dolomite is forming
from relatively high ionic strength seawater solutions in which the Mg2 to Ca2 solution ratios are consider-
ably higher than in normal seawater. However, they also recognize that factors other than the Mg2 to Ca2
ratio may be involved in dolomite formation. Sulfate depletion and alkalinity increase (increase in saturation
state) in pore waters resulting from anaerobic oxidation of organic matter by bacteria using SO42 as a
substrate, as well as direct participation by bacteria, may play a role in the dolomite formation process.
Dolomite is not found in abundance in modern sediments but is an important component of older sedimen-
tary rocks, and its origin has important implications for interpreting the history of seawater and the atmosphere
through geologic time (see Veizer and Mackenzie, Chapter 9.15).
Volume Editors Introduction xxv

Except for the sponges of the latest Precambrian Ediacaran fauna, most of the organisms that make their
skeletons of amorphous silica are of Cambrian age or younger. The appearance of radiolarians and diatoms in
the Phanerozoic changed the manner of deposition of amorphous silica on the sea floor from a primarily
abiotic process to one in which biochemical processes play a strong role, and led to a shift in the locus of silica
deposition from the shallow shelf to the deep sea. Prior to the evolution and spread of siliceous biota in the
Cambrian, the oceans contained significant concentrations of dissolved silica (Garrels and Perry, 1974; Siever,
1992), reaching saturation levels with respect to amorphous silica in Precambrian aquatic systems. As siliceous
organisms became more prevalent in the Phanerozoic, the surface waters of the ocean became depleted in
dissolved silica, and the average concentration of dissolved silica in seawater fell from a level of about 1000 to
100 mM. In the context of these significant events in biotic evolution and seawater chemistry, Dave DeMaster in
Chapter 9.4 (reprint) and Ed Perry and Liliana Lefticariu in Chapter 9.5 (revised) discuss the Diagenesis of
Biogenic Silica: Chemical Transformations Occurring in the Water Column, Seabed, and Crust and the Formation and
Geochemistry of Precambrian Cherts, respectively. DeMaster examines the chemical and structural transforma-
tions that take place following the death of siliceous organisms in the oceans and that lead to changes in the
solubility and dissolution kinetics of the original biogenic amorphous silica phase. He documents the
stabilization pathways of the transformation of biogenic silica from amorphous opal-A to opal-CT and finally
in many cases to microcrystalline quartz (the rock chert) as the biogenic silica is buried beneath meters to
kilometers of sediment and is exposed to higher temperatures and pressures (e.g., Williams and Crerar, 1985).
DeMaster also recognizes the importance of the tie between the Al and Si cycles and reverse weathering
diagenetic processes as a sink for elements in the ocean.
In Chapter 9.5, Perry and Lefticariu highlight the fact that the texture and mode of formation of Precambrian
cherts make them a uniquely suitable medium for preservation of early life forms. They point out that one
fruitful approach to reconstructing the sedimentary and diagenetic history of Precambrian cherts (and chert-
bearing rocks such as iron formations; see Bekker, Chapter 9.18) is to study the isotopic composition of their
constituent elements: oxygen, silicon, iron, and hydrogen. Advances in precision and availability of isotopic
analyses together with rare earth element chemistry have produced a large and rapidly increasing body of
geochemical data that provide constraints on mechanisms of chert formation and on geochemical processes
and environmental conditions on the Precambrian Earth. Cherts precipitated from aqueous solution have the
potential to preserve a record of the oxygen isotopic composition of contemporaneous sea or lake water and
hence the temperature of the environmental medium. The problems of course are that the oxygen isotopic
signal of cherts is a function of both temperature and solution composition and that siliceous precipitates
undergo diagenesis and may encounter temperatures and pressures of metamorphism in their passage through
the sedimentary cycle leading to isotopic exchange. Perry and Lefticariu conclude that the low d18O of many
Precambrian cherts is a consequence either of Precambrian Earth-surface temperatures higher than the average
Phanerozoic temperature or of a lower 18O/16O ratio in the Precambrian ocean (or of both). At present, both
the high seawater temperature and the low d18O hypotheses have evidence to support them. Each also has
evidence for rejection. However, suggestions that one of the earliest progenitors of life was an anaerobic
thermophilic microorganism evolving in a hydrothermal or volcanic environment (e.g., Kasting and Chang,
1992) would be compatible with an interpretation of high early Earth-surface temperatures.
Chapters 9.6 and 9.7 focus mainly on the geochemistry and sedimentology of detrital and organic carbon-
rich sedimentary facies. These sedimentary materials preserve records of multiple biogeochemical proxies that
reveal the processes and conditions of sediment formation, transport, deposition, burial, and diagenesis and
provide information regarding the evolution of Earths surface environment. The proxies include sedimentary
mineralogical suites and their distribution in time and space, major, minor, and trace-element concentration
data, the stable isotopic composition of carbon, oxygen, sulfur, and nitrogen, radiogenic isotopes such as 40K,
and the abundance and isotopic composition of biomarker compounds. In Chapter 9.6 (revised and signif-
icantly refocused), Geochemistry of Fine-Grained Organic Carbon-rich Facies, Brad Sageman, Tim Lyons, and
Young Ji Joo deal mainly with the fine-grained, organic carbon-rich black shale and mixed siliciclasticbiogenic
sedimentary facies. The chapter provides an overview of the fundamental concepts and tools that have been
applied to study organic carbon-rich facies and the history of changes in marine redox conditions that they
uniquely record. It first develops a conceptual model for the processes and feedbacks involved during the
deposition of fine-grained organic carbon-rich black shales and illustrates how geochemical proxies are
developed from detrital, biogenic, and authigenic fluxes and processes. The proxies reviewed include some
recently introduced metal isotope proxies (isotope systematics of Mo, Fe, and ReOs), in addition to elemental
and compound concentration/accumulation proxies (trace metal, biomarker, CaCO3, etc.) and stable isotopes
xxvi Volume Editors Introduction

(C and S) that have been in use for many years. Highlights from recent publications on Precambrian strata,
along with selected studies of Phanerozoic and modern anoxic deposits (i.e., Ocean Anoxic Events and
sapropels), are presented in this chapter to provide a broad review of the concepts and methods used to
study organic carbon-rich deposits commonly referred to as hemipelagic. The authors conclude that fine-
grained, siliciclasticbiogenic sediments perhaps provide the most fully integrated record of past Earth-surface
processes of any sedimentary facies and thus offer the opportunity to open a clearer window into interpretation
of the environment of Earths surface through geologic time.
In Chapter 9.7 (reprint), Late Diagenesis and Mass Transfer in Sandstone-Shale Sequences, Kitty Milliken
discusses what is known about the chemical and physical processes that lead to the transformation of detrital
siliciclastic sediments into rocks. These processes take place during burial diagenesis and low-grade metamo-
rphism and certainly involve basin-scale mass transfer of helium, water, and petroleum, and significant
transport of carbon out of shales into the oceanatmosphere system. Large-scale silica transfer from shales to
sandstones is also a persistent feature of burial diagenetic sequences, as the processes of smectite dissolution
and illite precipitation (see also Lerman and Clauer, Chapter 9.16) and quartz pressure dissolution in shales
and quartz precipitation in sandstones act in concert, and silica solubility increases with increasing burial and
temperature. Calcium is lost from shales, as is carbonate, due to the dissolution of CaCO3 and, although
controversial, the probable loss of Ca from smectite as it is transformed to illite (Garrels and Mackenzie, 1974).
Potassium progressively increases in some shale sequences undergoing burial diagenesis but decreases in
others. However, the global effect appears to be the transport of potassium from other lithologies into shales
(Garrels and Mackenzie, 1972). The fate of magnesium and sodium is not well constrained, although Garrels
and Mackenzie (1971, 1974) argue that both elements are lost from shales with diagenesis and time, and that
some magnesium is transferred via late dolomitization processes to the carbonate reservoir with increasing age
of the rock mass (see Holland, 2003 and Veizer and Mackenzie, Chapter 9.15). Milliken concludes her chapter
with emphasis on the conclusion that most precipitation reactions occurring during the diagenesis of sand-
stones and shales are acid-releasing reverse weathering reactions and can influence the crustal CO2 cycle (e.g.,
Kastner, 1974; Kerrick et al., 1995).
Chapters 9.8, 9.9, and 9.10 deal with the formation and fate of materials derived from living organic matter,
its degradation products, and elements generally associated with organic-rich sediments deposited in aquatic
environments, especially the biogeochemistry of sedimentary sulfur. In Chapter 9.8 (revised), Coal Formation
and Geochemistry, Bill Orem and Bob Finkelman point out that because coal is an organic rock largely
composed of an assemblage of amorphous, degraded plant material, an admixture of syngenetic, diagenetic,
epigenetic, and detrital mineral grains, and contains within its structure water, oils, and gases, it is one of the
most complex and challenging sedimentary rocks to analyze and understand. Coal is the most abundant of the
fossil fuels and the second largest source of energy in the world. In 2010, it supplied about 30% of the worlds
commercial energy (Mackenzie, 2011). In the USA, more than half of the electricity generated is by coal-fired
power plants.
Coal formation begins during the process of peatification in which microorganisms preferentially degrade
plant biomolecules such as carbohydrates and leave other materials such as lignin and waxes selectively
preserved. In the early stages of coal formation (coalification), numerous chemical changes to biomolecules
occur such as loss of O functionality and changes in lignin composition but the overall structure of plant
materials of low rank coals remains relatively intact. Later stage coalification producing bituminous and higher
rank coals involves more substantial changes in the structure of the coal. These include the transformation of
coal to a hard, black, lustrous organic rock depleted in both O and H relative to the precursor material and with
aromatic structures that are more condensed.
In contrast to coal formation in which the most important starting materials are biopolymers found in
vascular plants with some admixture of resistant algae polymers as well as bacterial polymers, kerogens are the
insoluble organic materials that are the immediate precursors to the formation of most oil and gas. Kerogen is a
very heterogeneous and complex aggregate of macerals, discrete particles of insoluble organic material identi-
fiable under the microscope and representing residual detritus from various sources of organic matter (e.g.,
Tissot et al., 1974). The types and maturity levels of kerogen play an important role in determining the type of
oil or gas generated. In 2010 oil provided 40%, and gas provided 20% of the commercial global energy
consumption (Mackenzie, 2011).
In Chapter 9.9 (revised), Formation and Geochemistry of Oil and Gas, Paul Philp describes the early steps of oil
and gas formation starting with the process ultimately responsible for most oil, gas, and coal formation, that of
photosynthesis. Today only about 0.1% of global net primary production of organic matter accumulates in
Volume Editors Introduction xxvii

sediments (Mackenzie et al., 1993; Tissot and Welte, 1978) and is available for the formation of fossil fuels;
thus, these commercial deposits of the stored energy of photosynthesis actually represent but a very small
portion of the gross production of organic matter and of total organic matter present in the sedimentary mass.
Philp discusses the nature of the insoluble organic materials, kerogen, and the source, depositional environ-
ment, maturity, biodegradation, age dating, and migration of the soluble components of kerogen. It is now
generally accepted that the major mechanism for the formation of most oil (and gas) is through thermal
degradation of kerogen:
Kerogen ! bitumen ! oil gas residue

This chapter is further concerned with the utilization of geochemical techniques to investigate the origin and
history of crude oils and natural gas and the source rocks from which they were formed. Once the oils are
formed and migrate to the reservoirs, molecular geochemistry is used to characterize the oils in terms of their
source, maturity, depositional environments, level of biodegradation, geologic age, and relationship with
specific source rocks. In a similar manner, natural gas and relationship to source rocks are investigated through
the utilization of stable isotopes and compositional analyses. The ultimate goal of the chapter is to demonstrate
the added value of organic geochemistry in terms of improving the success of petroleum exploration efforts and
understanding the origin and history of crude oils and natural gas. In conclusion, Philp argues that the
integration of geochemical parameters, such as biomarkers, with sequence stratigraphy models will become
more important in the future in petroleum exploration in terms of predicting how the processes of oil and gas
generation, expulsion from source rocks, migration to reservoir rocks and subsequent trapping, and preserva-
tion control the volume, quality, and distribution of petroleum.
Sulfur is one of the important elements commonly associated with organic-rich sediments and a principal
element found in living organisms. Since the chapter by Marty Goldhaber on sulfur-rich sediments appeared in
volume 7 of the first edition of the TOG, these sediments have proved to be one of the most exciting areas of
focus of contemporary Earth and environmental science research. In Chapter 9.10, The Sedimentary Sulfur
System: Biogeochemistry and Evolution through Geologic Time, David Rickard discusses the geochemistry and
geomicrobiology of recent and ancient sedimentary sulfur systems, both in the context of processes and the
evolution of the sulfur system. Rickard also points out that the sulfur system is important to environmental
scientists considering the present and future effects of pollution and anoxia in that the development of the
sulfur system, particularly the characteristics of ocean anoxia over the last 200 Ma, can be considered to infer
the future fate of the Earths surface system. Rickard shows that the shale lithofacies is by far the largest sulfur
reservoir in sedimentary rocks and nearly all marine sediments with more than a few tenths of a percent of
organic carbon and those nonmarine sediments with significant concentrations of SO42 in depositional
waters contain the mineral pyrite (FeS2).
Pyrite, along with sulfur-bearing organic compounds and other mineral phases of sulfur such as mack-
inawite, greigite, and elemental sulfur, forms at Earth-surface temperatures and pressures primarily through the
metabolic activities of sulfate reducing microorganisms. In the global context, the evolution of the cycling
behavior of sulfur through geologic time is closely tied to the carbon cycle and the evolution of atmospheric
oxygen (e.g., Garrels et al., 1976; Petsch and Berner, 1998). Variations in the sulfur isotopic composition of
sedimentary sulfate minerals and pyrite provide constraints on the isotopic composition of seawater through
geologic time and suggest that both atmospheric O2 and oceanic SO42 were very low prior to 2.4 Ga. The
presence of very significant mass-independent sulfur fractionations in sulfate and sulfide minerals older than
2.45 Ga (Farquhar et al., 2000) supports this conclusion. Rickard goes on to demonstrate that the pyrite burial
data are consistent with the secular variations in the average d34S pyrite, which show that this variable becomes
negative in the Phanerozoic, whereas it was around zero (or slightly positive) during the Precambrian. He
concludes that the Neoarchean sulfur data are presently seriously imbalanced with a large mass of 34S missing
from the system as it is currently understood. Richard further points out that although the biochemical
apparatus for sulfate reduction evolved at an early stage in the history of the Earth, the general dearth of
marine sulfate limited its ubiquity. To conclude his chapter, David Rickard states that If the old sailing masters
were cruising the pre-Phanerozoic oceans, it is doubtful that they could have used the presence of blue muds as
a location indicator.
Chapters 9.11 and 9.12 are concerned with the formation of manganiferous phases in sediments, rocks, and
ores and the sedimentary green clay minerals. Manganese is of significant interest in geochemistry because its
minerals are both tracers of redox processes and accumulators of other elements of great geochemical signif-
icance (Maynard, 1983). Manganese is the ninth or tenth most abundant element in the crust and is very
xxviii Volume Editors Introduction

similar to iron in its chemical properties; however, higher valence states distinguish their behavior from iron
and give rise to a plethora of Mn oxide minerals. In Chapter 9.11 (revised), Manganiferous Sediments, Rocks, and
Ores, Barry Maynard discusses the basic geochemical properties of Mn and the common Mn minerals, the
distribution of Mn in rocks and natural waters, and the composition of significant accumulations of Mn. He
addresses the behavior of Mn especially in mid-ocean ridge vent systems and during sedimentation. Through-
out the chapter, Maynard emphasizes the difference between the geochemistry of Mn and Fe, and points out
that under reducing and mildly oxidizing conditions because of the solubility of Mn compared to Fe, Mn is
preferentially exported from low oxygen environments, such as basalt-hydrothermal systems and euxinic
sedimentary basins. The Mn subsequently accumulates in oxidizing environments of the shallow ocean or in
low-productivity areas of the deep sea. Maynard further shows that tracking the Mn/Fe ratios of sediments
provides a means of reconstructing the oxidation structure of ocean basins, soils, or groundwater systems. In
addition, the strong affinity of Mn for certain transition elements and the rare earth elements is an important
property of the element. In the case of Ce and Eu, these elements when preserved in Mn accumulations can
define the relative importance of hydrothermal and diagenetic processes in the sedimentary record.
In Chapter 9.12 (reprint), Green Clay Minerals, Bruce Velde demonstrates that the basic green color of
glauconite, celadonite, berthierine, verdine, chamosite, nontronite, and talc is due to the presence of iron in the
structures of these minerals. This is not a startling revelation to mineralogists, but it is the key to understanding
the origin and stability of the green clay mineral phases in nature. Velde concludes that green clay mineral
names can reflect compositional differences, mineral structural differences, or differences in the geological
occurrence of these phases. This in itself leads to much confusion in the literature concerning the origin and
diagenesis of the green clay minerals. Regardless, the mineralogy and chemistry of the green clay minerals
provide valuable clues to the temperature and solution composition of the environment in which they formed.
For example, both glauconite and berthierineverdine minerals are low-temperature phases that form in
shallow-water sediments at low sedimentation rates. The formation of glauconite probably requires the more
oxidizing conditions of organic-poor sediments rather than the lower redox conditions in organic-rich sedi-
ments. The original glauconite precursor evolves toward a potassic, ferric clay mineral whereas that of
berthierine evolves toward an alkali-free, ferrous mineral. Thus the two minerals can be used to infer their
environmental conditions of deposition and diagenesis. Many of Veldes conclusions are based on work started
more than 25 years ago (e.g., Velde, 1985).
The next four chapters of Volume 9 concentrate on the longer term historical aspects of sediments and
sedimentary rocks, their cycling behavior, and the biogeochemical and isotopic proxies found within them that
permit the interpretation of the evolution of Earths surface environment through geologic time. Interpretations
of this nature first require a temporal framework in which to place information and events, and this is the
subject of Bill Berrys Chapter 9.13 (reprint), Chronometry of Sediments and Sedimentary Rocks. Sadly Bill passed
on before he could complete revision of his chapter for this edition of the Treatise. Berry points out in the
reprint of his chapter that it is economically imperative to develop a chronometry of sedimentary rocks in order
to find and recover larger volumes of natures resources of coal, ores, building stone, sand, gravel, and
eventually petroleum. The basic principle of chronometry based on the fossil record is embodied in the
Table of Strata that William Smith dictated to his ecclesiastic associates in June 1799 (see Berry, 1987) and
still holds true today, that of the succession of fauna and flora found in the rock record. Berrys description of
the development of chronometry based on the fossil record is essentially that of the early development of the
science of geology itself. Chorology, the science that deals with the geographical distribution of living
organisms, is discussed in terms of the constraints it imposes on chronometry. As Berry points out, chronom-
etry cannot provide the absolute ages of rocks. Fossil-based chronometry has been enhanced through studies of
radiochronometry and magnetobiochronometry, as well as orbital chronometry. The chronometry of sedi-
mentary rocks supplies the necessary framework for interpretation of events in Earth history and the long-term
secular evolution and cycling of the sediment-oceanatmosphere-biota system, the subject matter of the
following two chapters.
In Chapter 9.14 (reprint), The Geochemistry of Mass Extinction, Lee Kump points out that the course of
biological evolution is inextricably linked to that of the environment in a complicated network of feedbacks at
all time and space scales. Perturbations of the environment and disruptions within have biological conse-
quences and vice versa. Kump considers mainly the elemental and carbon, sulfur, strontium, and oxygen
isotopic evidence for the environmental nature of the big five extinctions of the Phanerozoic (Sepkoski, 1993):
the Late Ordovician ( 440 Ma), the Late Devonian ( 367 Ma), the PermianTriassic ( 251 Ma), the Triassic
Jurassic ( 200 Ma), and the CretaceousTertiary ( 65 Ma). His conclusion is that there are no universal
Volume Editors Introduction xxix

geochemical precursors or responses to extinction events in the Phanerozoic. However, the development of
widespread ocean anoxia appears to be associated with all three Paleozoic mass extinctions. Such anoxic
conditions can certainly lead to mass mortality of extant organisms, and potential rapid turnover of anoxic
ocean waters could produce elevated concentrations of CO2 and H2S in surface waters resulting in the death of
aerobic organisms. Rapid ocean turnover could be the result of an oceanic impactor during the Late Devonian
and PermianTriassic events. However, a purely terrestrial origin of the Late Ordovician extinction event is
probable due to a glacioeustatic sea-level fall causing a loss of shallow-water marine habitats. In contrast, it
appears that both the TriassicJurassic and CretaceousTertiary mass extinctions occurred during periods of
well-oxygenated oceans. An asteroid or cometary impact appears to be the proximal cause of these events. At
least for the K-T event, the consequent release of huge quantities of CO2 and SO2 to the atmosphere changed
the climate and caused mass mortality of organisms.
In Chapter 9.15 (revised), Evolution of Sedimentary Rocks, Jan Veizer and Fred Mackenzie show that the
progressive increase with decreasing age of the observed extant thickness and areal extent of Phanerozoic
sedimentary strata and the secular variations in the relative proportions of lithological types and their chemistry
and mineralogy are a function of both evolution and recycling of the sedimentary mass. They also conclude that
the global sedimentary mass of 2.7  1024 g is largely cannibalistic and has a half-mass age of about 600 Ma,
resulting in total sediment deposition and destruction equivalent to about five sedimentary masses over the
course of geologic time (e.g., Garrels and Mackenzie, 1971). The differential recycling of various components of
the sedimentary rock mass is primarily controlled by the probability of preservation of different tectonic
settings in which the sediments are found and the recycling rates of the tectonic realms (e.g., Veizer, 1988).
Because of differential recycling, the preservation of the sedimentary mass is inherently biased. For example, the
half-life of the post-Devonian sedimentary mass is about 130 Ma, whereas the half-life of post-Permian
carbonates is only about 86 Ma; thus, the carbonates are recycled at a rate roughly 1.5 times faster than the
total sedimentary mass. This difference is due to the fact that since post-Permian time, the locus of carbonate
deposition has progressively shifted to the deep sea where post-Permian carbonates are more susceptible to
destruction because of subduction. The absence of Paleozoic and earlier preserved deep-sea carbonates implies
that continental carbonates are the only record of carbonate deposition and chemical, mineralogical, and
isotopic data for much of geologic time. However, this lack of preservation of deep-sea carbonates during this
time interval does not preclude their deposition abiotically in the Paleozoic and earlier deep ocean basins
(Berner and Mackenzie, 2011). Using the full range of mineralogical, chemical, and Sr, Os, S, C, and O isotopic
evidence, Veizer and Mackenzie also consider the chemical evolution of the oceanatmosphere system from the
anoxic conditions of the early Precambrian to the modern oxic global environment.
In Chapter 9.16 (new in print and revised), Stable isotopes in the Sedimentary Record, Abraham Lerman and
Norbert Clauer address the problems inherent in the interpretation of the sedimentary records of several
common stable isotopes and their fractionation in natural systems. There are 258 stable isotopes of the
chemical elements; those with the atomic numbers 120, hydrogen through calcium, are elements of the
main chemical constituents of the atmosphere, continental and oceanic waters, sediments, and the upper part
of the crystalline continental and oceanic crust. Lerman and Clauer consider the isotopes of hydrogen (D/H)
and oxygen (18O/16O) that are important in terms of interpretation of the behavior of the global water cycle
and are fractionated by a number of physical and chemical processes in the gasliquidsolid H2O system, as
well as in many mineralwater reactions in sediments. The isotopes of calcium (44Ca/40Ca) in biogenic
carbonates are shown to be an important tool in assessing the riverine flux of Ca to the ocean over geologic
time. Carbon, nitrogen, and sulfur are three of the major six elements essential to life and the sedimentary
record of their isotopes are variably extensive during Earths history. The sedimentary records of the isotopic
fractionation of carbon (d13C) and sulfur (d34S) are among the most well developed and shed light on ocean
atmospherebiosphere evolution over geologic time. Lerman and Clauer emphasize that the boron isotopes
(11B/10B) are invaluable in determining the paleo-pH of ocean waters. Finally, in addition to the nonradiogenic
stable isotopes mentioned above, the authors devote a section in this chapter to the radiogenic stable 40Ar
isotope because of its importance as a tracer of certain diagenetic processes, including the conversion of
montmorillonite to illite in sand-shale sequences during burial diagenesis (see also Milliken, Chapter 9.7).
Chapters 9.17, 9.18, and 9.19 conclude Volume 9 with discussions of three more sedimentary lithofacies:
evaporite mineral accumulations, iron formations, and bedded barite deposits. In Chapter 9.17, Geochemistry of
Evaporites and Evolution of Seawater, Maciej Babel and Charlotte Schreiber demonstrate that marine evaporites
are chemical deposits which represent the direct record of the chemistry of ancient oceans and the soluble ionic
load present in their waters. The chemistry of modern and ancient marine evaporites and their parent waters is
xxx Volume Editors Introduction

reflected in their mineralogy and facies distribution in geological space and time. Babel and Schreider show that
the evaporation of modern seawater and the order of precipitated evaporite salts, as well as the environment in
which marine evaporites are deposited, is the key for understanding the fossil record. Crucial to our under-
standing of the origin of these evaporites is the order of salts precipitated from evaporated modern seawater:
Ca-carbonates ! Ca-sulfates ! Na-chlorides ! Mg-, K-sulfates and chlorides:

This mineralogical order is usually preserved in the Phanerozoic evaporites, although Mg- and K-salts are rarely
present. This order was the cornerstone of the now discredited idea that the chemistry of the ocean was stable
and remained constant throughout Phanerozoic time and presumably was similar to todays ocean even earlier,
in the Precambrian. Studies over the past few decades have led to the emergence of a new and evolving picture
of the ocean in part based on the chemical composition of fluid inclusions in evaporite minerals (see e.g.,
Horita et al., 2002); that is, ocean chemistry has varied significantly over geologic time. Babel and Schreiber
show that the pair of ions Mg2 and Ca2 was especially crucial in the evolution of the sedimentary record, and
during the Phanerozoic, the seawater Mg2/Ca2 molar ratio changed significantly in a broadly cyclic pattern
with time.
Andrey Bekker, N. Planavsky, B. Krapez, B. Rasmussen, A. Hofmann, J. F. Slack, O. J. Rouxel, and K. O.
Konhauser in Chapter 9.18, Iron Formations: Their Origins and Implications for Ancient Seawater Chemistry, point
out that iron formations are economically significant, iron- and silica-rich sedimentary rocks that are restricted
to Precambrian successions. Modern and Phanerozoic analogues for these deposits that are comparable in
terms of areal extent and thickness are unknown. The authors show that secular changes in the style of iron
formation deposition are genetically linked to plate tectonic processes, mantle plume events, and evolution of
the Earths surface environment. They recognize two types of Precambrian iron formations: (1) those formed
proximal to volcanic centers and interlayered with or linked laterally with submarine volcanic rocks and, in
some cases, with volcanogenic massive sulfide deposits, and (2) larger sedimentary rock-hosted iron forma-
tions developed in passive-margin settings and typically lacking any direct association with volcanic rocks.
Banded iron formations, the BIFs, are predominant in Archean to earliest Paleoproterozoic successions,
whereas granular iron formations, the GIFs, are more common in middle to late Paleoproterozoic successions
deposited after the rise of atmospheric oxygen at about 2.4 Ga.
Bekker et al. further emphasize that giant sedimentary rock-hosted iron formations first appeared about
2.6 Ga. From about 2.6 to 2.4 Ga, global mafic-to-ultramafic magmatism culminated in the deposition of giant
sedimentary rock-hosted iron formations in South Africa, Australia, Brazil, Russia, and Ukraine. The younger
BIFs in this age range were deposited immediately before a shift from reducing to oxidizing conditions in the
oceanatmosphere system. It is postulated by the authors that enhanced magmatism at 2.502.45 Ga, which
likely delivered large amounts of reductants to surface environments, may have triggered atmospheric oxida-
tion. After the rise of atmospheric oxygen at about 2.4 Ga, GIFs became more abundant in the rock record than
BIFs, and iron formations largely disappeared from the rock record at about 1.85 Ga, but reappeared at the end
of the Neoproterozoic tied to periods of intense magmatic activity and also to global-scale glaciations during
Snowball Earth events. In the Phanerozoic, marine iron deposition was restricted to local environments of
closed to semi-closed basins, where volcanic and hydrothermal activity was important, such as in back-arc
basins. Basin-scale and episodically deposited Phanerozoic oolitic and pisolitic ironstones are linked to periods
of intense magmatic activity and ocean anoxia.
In Chapter 9.19, Bedded Barite Deposits: Environments of Deposition, Styles of Mineralization, and Tectonic
Settings, Erika Elswick and Barry Maynard show that modern accumulations of barite can be classified as
pelagic, diagenetic, cold seep, or hydrothermal. The isotopic composition of the Sr incorporated in the barite
can discriminate some, but not all, of these regional settings. The bulk chemistry of the barites themselves,
which dominantly reflects fluid sources and fluid mixing, and the chemistry of associated volcanic rocks, which
reflects the tectonic setting, can further discriminate between these settings. Ancient examples of massive
bedded barite have marine host rocks, deposited in relatively deep water, and share a black shalechert
turbidite sandstone facies. Geochemical discriminators for these types of deposits are the chemistry of the
mudstone hosts, which dominantly reflects the environment of deposition; the chemistry of the barites
themselves, which dominantly reflects fluid sources and fluid mixing; and the chemistry of associated volcanic
rocks, which reflects tectonic setting. The authors conclude that the best discriminator of tectonic setting is the
composition of associated volcanic rocks. Interestingly, the authors suggest that for the commercially important
Nevada bedded barite deposits, there are two competing tectonic models: (1) an island arc underlain by an
east-dipping subduction zone, which would make the area a back-arc basin; or (2) a west-dipping subduction
Volume Editors Introduction xxxi

zone, which would produce a remnant ocean basin with a passive continental margin to the east and the fore-
arc side of an island arc to the west with an associated accretionary prism eventually obducted onto the eastern
platform. However, it appears that the California Borderland classical model of barite deposition is the most
likely analog for the Nevada bedded barite deposits (see e.g., Maynard and Okita, 1991).

Fred T. Mackenzie
University of Hawaii at Manoa, Honolulu, HI, USA

References

Archer D, Emerson S, and Smith CR (1989) Dissolution of calcite in deep-sea sediments: pH and O2 microelectrode results. Geochimica
et Cosmochimica Acta 53: 28312845.
Arrhenius G (1963) Pelagic sediments. In: Hill MN (ed.) The Earth Beneath the Sea: History. The Sea: Ideas and Observations on
Progress in the Study of the Seas, vol. 3, pp. 655727. New York: Interscience Publishers.
Bender M, Jahnke R, Weiss R, et al. (1989) Organic carbon oxidation and benthic nitrogen and silica dynamics in San Clemente Basin, a
continental borderland site. Geochimica et Cosmochimica Acta 53: 685697.
Berner RA and Mackenzie FT (2011) Burial and preservation of carbonate rocks over Phanerozic time. Aquatic Geochemistry
17: 727733.
Berry WBN (1987) Growth of the Prehistoric Time Scale, Revised Edition. Oxford: Blackwell.
Chave KE (1960) Carbonate skeletons to limestones: Problems. Transactions of the New York Academy of Sciences 23: 1424.
Chester R (1990) Marine Geochemistry. London: Unwin Hyman 698 p.
Farquhar J, Huiming B, and Thiemens MH (2000) Atmospheric influence of earths earliest sulfur cycle. Science 289: 756758.
Garrels RM, Lerman A, and Mackenzie FT (1976) Controls of atmospheric O2 and CO2: Past, present, and future. American Scientist
63: 306315.
Garrels FT and Mackenzie FT (1971) Evolution of Sedimentary Rocks. New York: W. W. Norton & Company 397 p.
Garrels RM and Mackenzie FT (1972) A quantitative model for the sedimentary rock cycle. Marine Chemistry 1: 2741.
Garrels RM and Mackenzie FT (1974) Chemical history of the oceans deduced from post-depositional changes in sedimentary rocks.
In: Hay WH (ed.) Studies in Paleo-Oceanography, pp. 193204. Tusla, OK: Society of Economic Paleontologists and Mineralogists
Special Publication No. 20.
Garrels RM and Perry EA Jr. (1974) Cycling of carbon, sulfur, and oxygen through geologic time. In: Goldberg ED (ed.) The Sea, vol. 5,
pp. 303357. New York: John Wiley.
Holland HD (2003) The geologic history of seawater. In: Elderfield H (ed.) The Oceans and Marine Geochemistry. 1st ed., Treatise on
Geochemistry, 1st ed., vol. 6, pp. 146. Amsterdam: Elsevier.
Horita J, Zimmermann H, and Holland HD (2002) Chemical evolution of seawater during the Phanerozoic: Implications from the record of
marine evaporites. Geochimica et Cosmochimica Acta 66: 37333756.
Kasting JF and Chang S (1992) Formation of the earth and the origin of life. In: Schopf JW and Klein C (eds.) The Proterozoic Biosphere:
An Interdisciplinary Study, pp. 912. New York: Cambridge University Press.
Kastner M (1974) The contribution of authigenic feldspars to the geochemical balance of the alkali metals. Geochimica et Cosmochimica
Acta 38: 650653.
Kerrick DM, McKibben MA, Seward TM, and Caldeira K (1995) Convective hydrothermal CO2 emission from high heat flow regions.
Chemical Geology 121: 1727.
Li Y-H, Takahashi T, and Broecker WS (1969) Degree of saturation of CaCO3 in the oceans. Journal of Geophysical Research
74: 55075525.
Mackenzie FT (2011) Our Changing Planet: An Introduction to Earth System Science and Global Environmental Change, 4th ed.
New York: Prentice Hall.
Mackenzie FT and Garrels RM (1966) Chemical mass balance between rivers and oceans. American Journal of Science 284: 507525.
Mackenzie FT and Kump LR (1995) Reverse weathering, clay mineral formation, and oceanic element cycles. Science 270: 586587.
Mackenzie FT, Ver LM, Sabine C, and Lane M (1993) C, N, P, S global biogeochemical cycles and modeling of global change.
In: Wollast R, Mackenzie FT, and Chou L (eds.) Interactions of C, N, P and S Biogeochemical Cycles and Global Change, pp. 161.
Berlin: Springer.
Martin WR, McNichol AP, and McCorkle DC (2000) The radiocarbon age of calcite dissolving at the sea floor: Estimates from pore water
data. Geochimica et Cosmochimica Acta 64: 13911404.
Martin WR and Sayles FL (2003) The recycling of biogenic material at the sea floor. In: Mackenzie FT (ed.) Sediments, Diagenesis, and
Sedimentary Rocks. 1st ed., Treatise on Geochemistry, 1st ed., vol. 7, pp. 3765. Amsterdam: Elsevier.
xxxii Volume Editors Introduction

Martin WR and Sayles FL (2006) Organic matter oxidation in deep-sea sediments: Distribution in the sediment column and implications
for calcite dissolution. Deep-Sea Research II 53: 771792.
Maynard JB (1983) Geochemistry of Sedimentary Ore Deposits. New York: Springer.
Maynard JB and Okita PM (1991) Bedded barite deposits of the U.S., Canada, Germany, and China: Two major types based on tectonic
setting. Economic Geology 86: 364376.
Michalopoulos P and Aller RC (1995) Rapid clay mineral formation in Amazon Delta sediments: Reverse weathering and oceanic
elemental cycles. Science 270: 614617.
Milliman JD, Troy PJ, Balch WM, et al. (1999) Biologically mediated dissolution of calcium carbonate above the chemical lysocline?
Deep Sea Research 46: 16531669.
Morse JW and Berner RA (1972) Dissolution kinetics of calcium carbonate in seawater. II. A kinetic origin for the lysocline. American
Journal of Science 274: 638647.
Murray J and Irvine R (1885) On the chemical changes which take place in the composition of seawater associated with blue muds on the
floor of the ocean. Transactions of the Royal Society of Edinburgh 37: 481507.
Murray J and Menard A (1891) Deep sea deposits. In: Challenger Reports. London: Longmans, 525 p.
Petsch ST and Berner RA (1998) Coupling the geochemical cycles of C, P, Fe, and S: The effect on atmospheric O2 and the isotopic
records of carbon and sulfur. American Journal of Science 298: 246262.
Sayles FL (1979) The composition and diagenesis of interstitial solutions-I. Fluxes across the seawater-sediment interface in the Atlantic
Ocean. Geochimica et Cosmochimica Acta 43: 527545.
Sayles FL (1981) The composition and diagenesis of interstitial solutions-II. Fluxes and diagenesis at the sediment-water interface in the
high latitude North and South Atlantic. Geochimica et Cosmochimica Acta 45: 10611086.
Sepkoski J Jr. (1993) Ten years in the library: New data confirm paleontological patterns. Paleobiology 19: 4351.
Siever R (1992) The silica cycle in the Precambrian. Geochimica et Cosmochimica Acta 56: 32653272.
Tissot B, Durand B, Espitalie J, and Combaz A (1974) Influence of nature and diagenesis of organic matter in formation of petroleum.
American Association of Petroleum Geologists Bulletin 58: 499506.
Tissot B and Welte DH (1978) Petroleum Formation and Occurrence. Berlin: Springer.
Tizzard TH, Moseley HN, Buchanan MA, and Murray J (1885) Report of the Scientific Results of the Voyage of H. M. S. Challenger during
the Years 18731876, vol. 1, pt. 1. London: Her Majestys Stationery Office, 509 p.
Veizer J (1988) The evolving exogenic cycle. In: Gregor CB, Garrels RM, Mackenzie FT, and Maynard JB (eds.) Chemical Cycles in the
Evolution of the Earth, pp. 175220. New York: Wiley.
Velde B (1985) Clay Minerals: A Physio-Chemical Explanation of Their Occurrence. Amsterdam: Elsevier 427 p.
Williams LA and Crerar DA (1985) Silica diagenesis, II. General mechanisms. Journal of Sedimentary Petrology 55: 312321.

View publication stats

You might also like