4.1 The Sediment Cycle
4.1.1 The Concept
Marine sediments consist of deposits
accumulating below the sea. They show great variety. There is the
debris from the wearing down of continents and volcanic mountains,
the shells derived from organisms, organic matter, minerals
precipitated from seawater, and there are volcanic products such as
ash and pumice. All such matter is transported in various ways, is
deposited, and subsequently suffers diagenesis and redeposition to
various degrees (Fig. 4.1). Terrigenous muds (with their tremendous
fraction of continental erosion products) make up well over one
half of the total volume of marine deposits. By area it is deep-sea
sediments that are dominant, with roughly one half of the deep
seafloor covered by calcareous ooze – largely shell material from
coccolithophores and foraminifers, along with some reef debris
along many tropical margins. In contrast, extraterrestrial matter,
while interesting for its content of information, is negligible
when considering abundance.
Fig.
4.1
Schematic summary of sources, transport,
and destination of marine sediments
4.1.2 Notes on Geochemistry
Concerning
seafloor geochemistry, reactions of seawater with the hot basalt of
the oceanic crust are fundamental. Many of the reactions are
conspicuous at the Ridge Crest (hot vents). Such reactions may
contribute considerable amounts of matter to seawater. Also, they
may largely relieve seawater of certain elements, such as
magnesium. The types and amounts of elements added or subtracted
are the subject of intensive studies, with the origin and evolution
of seawater at focus.
As concerns the input to the
terrigenous sediment reservoir by erosion, the sea itself takes a
toll from the continents: waves and tides can eat into the land,
taking debris offshore. Such material, along with the sediment
delivered by rivers, by ice, or by winds, may stay on the shelf, or
it may bypass the shelf to accumulate on the slope or below that on
the rise or the abyssal plain. Sediments that stay on the shelf, of
course, essentially stay on the continent. The sediment deposited
beyond the shelf eventually is carried back to a continent (e.g.,
by mountain building) or disappears into the mantle by subduction,
as discussed earlier.
4.2 Sources of Sediment
4.2.1 River Input
What rivers carry (the dissolved and
particulate load) constitutes the main source of marine sediment
making the coalescing fans that build the continental slope off
California and elsewhere. The fans are largely riverine mud with an
admixture of shell and organic matter. The admixed shell material
(largely made of calcium carbonate and opal) also derives to a
great extent from river input, that is, from the dissolved load
brought by rivers to the sea. We can make a rough estimate of how
much material is involved. Slope sediments typically accumulate at
a rate of some 100 mm per millennium and deep-sea oozes at rates at
one tenth of that, with one half of the deep seafloor ooze-free and
accumulating clay at 1 mm/1000 years.
The sediment supply from rivers is
reported to be near 12 cubic km per year. If this amount is
distributed on the area of seafloor on the planet, we obtain an
average sedimentation rate of about 30 mm/1000 years. The erosion
rate for continents, at twice that value, would be 60 mm per
millennium. The rates of input from reactions of seawater with the
basalt of the oceanic crust and from leaching of soils and rocks on
land are neglected in this calculation, an input that if considered
would decrease the estimate of continental erosion rates
somewhat.
In general, mechanical erosion
dominates in high latitudes, where much of the water is ice, and in
deserts, where water transport is linked to flash floods and other
sporadic action unimpeded by vegetation. Chemical weathering
(leaching) is favored by rainfall and high temperatures and
dominates in tropical areas. In extrapolating these and other
present-day patterns into the past, we must remember that we live
in a highly unusual period. The growth of mountain ranges and the
powerful abrasive action of glaciers have greatly increased
mechanical erosion for the last several million years.
4.2.2 Input from Ice
The great importance of ice in
delivering sediment to the sea in high latitudes is readily
appreciated when contemplating the immense masses of outwash
material glaciers bring to the shores, material that is reworked on
the shelves by wave action and by nearshore currents.
Less important as concerns sediment
mass, but very useful in the reconstruction of paleoclimate, is the
fact that icebergs from calving glaciers can transport both fine
and very coarse materials far out to the sea. Upon melting, an
iceberg drops its load (as ice-rafted debris or IRD). Around Antarctica iceberg
transport reaches to about 40°S, well off the continent. At
present, about 20% of the seafloor receives at least some
ice-transported sediment. During the last glacial maximum (LGM), as expected, the dropstone limit
extended much farther toward the equator than today (Fig.
4.2).
Fig.
4.2
IRD (ice-rafted debris) on the seafloor of
the North Atlantic, as recorded by H. R. Kudrass, Bundesanstalt,
Hannover. Present limits of drift ice shown schematically (broken
lines). Samples of IRD (black
dots) well toward the south presumably indicate conditions
of cold phases of the ice age (suggested southern limit here added:
shaded zone) (after H.R. Kudrass, 1973. Meteor
Forschungs-ergebnisse Reihe C 13:1, modified). Note the assumed
great extent of winter icebergs in the glacial North Atlantic,
based on IRD (blue). Also
note the addition of land as sea level dropped from ice buildup on
land
4.2.3 Input from Wind
In striking contrast to ice, wind only
moves fine material. Medieval Arabian scientists already noted the
dust coming out of the Sahara into the “dark sea” of the Atlantic
(Fig. 4.3). In
the nineteenth century, Charles Darwin assumed (correctly) that the
dust blown offshore off NW Africa must end up on the seafloor.
Particles in a large Saharan dust storm in 1901 had an average size
of about 0.012 mm in Palermo. Sizes of particles were about half of
that in Hamburg (i.e., Hamburg was reached by extremely fine silt,
the rest having settled out on the way). During this same storm up
to 11 g of dust per cubic meter were measured over the
Mediterranean. Some Saharan dust ends up in the Caribbean, as noted
in satellite images. In the Pacific, we have the example of quartz
on and around Hawaii, fine grains apparently blown in from sources
in China, thousands of km away.
Fig.
4.3
Abundance of haze from dust over the
Atlantic (G.O.S. Arrhenius, 1963. In M.N. Hill (ed.) The Sea vol.
3: 695; color here added)
4.2.4 Input from Volcanism
Volcanoes supply a substantial amount
of marine sediment, especially in the vicinity of active margins,
such as the ones around much of the Pacific basin. While much of
the “volcanic ash” carried
by winds is widely dispersed and makes thin layers, layers several
cm thick are not uncommon in deep-sea sediments. Some of these,
marking periods of major eruptions, correlate over large distances (i.e.,
they originate from the same event and are of equal age,
therefore). A well-known example is the Toba super-eruption in
Sumatra (Indonesia) 73,500 years ago. Volcanic products that reach
the stratosphere result in strong cooling. In the Toba case, the
effect presumably was severe and lasted for several years. It has
been suggested that such cooling may have contributed to the fast
re-glaciation at the end of the last interglacial. While we may
indeed be looking at tectonic feedback in climate change
(re-glaciation may have changed the distribution of gravitational
forces on the planet and thereby helped trigger eruptions),
suggestions of a link between volcanism and re-glaciation are
somewhat ad hoc and superfluous in a Milankovitch setting.
Near island arcs, volcanic ash layers
(tephras) can build
sediment aprons several thousand meters thick from tens of
thousands of eruptions. The composition of deep-sea clay (Chap.
10) suggests that in the time before
ten million years ago, before vigorous mountain building and
glaciation changed the planetary environment in drastic fashion,
the main source of deep-sea clay in the Pacific was the
decomposition of volcanic ash. Not all volcanogenic sediment is
brought by wind. Pumice,
highly porous volcanogenic rock with plenty of air within, can
float and move with ocean currents for long distances, even
carrying gooseneck barnacles and other organisms. Of course,
volcanic material also is eroded on land and is then brought into
the sea as terrigenous sediment. Differences in color, mineral
composition, glass properties, as well as subtle differences in
chemistry hold the clues as to the source of a given volcanogenic
deposit.
Geologically, volcanoes have a short
life, with single eruptions marking flash-like events. Ash layers,
therefore, are very useful in much of regional stratigraphy (as
tephrochronology), for
example, in the Mediterranean realm (Fig. 4.4) or in the vicinity of
Iceland. Volcanic activity also adds gases and hydrothermal
solutions to the sea. These fluxes have an important bearing on the
evolution of the chemistry of seawater and of the atmosphere. As
mentioned earlier, they are the subject of intense study.
Fig.
4.4
Distribution of volcanic ash on the
seafloor of the Mediterranean near Crete; ash is produced by two
large explosions in the Aegean Sea (presumably Thera and
Santorini). The lower ash layer marks a prehistoric event
(>25,000 years). The upper layer is less than 5000 years old and
possibly reflects an eruption that ended the Minoan culture (After
D. Ninkovich and B.C. Heezen, from K.K. Turekian, 1968. Oceans.
Prentice-Hall, New Jersey. Simplified topography; color here
added)
4.3 Sediments and Seawater Chemistry
4.3.1 Acid-Base Titration
To a first approximation, seawater is
a solution of sodium chloride (with a bit of Epsom salt thrown in).
Sodium and chloride make up 86% of the ions present by weight. The
other major ions are magnesium, calcium, and potassium (alkaline
and earth alkaline metals) and the acid radicals sulfate and
bicarbonate. The major cations form strong bases that are balanced,
on the whole, by the acid radicals. However, bicarbonate forms a
weak acid, and seawater is slightly alkaline, therefore, with a
“pH” near 8, slightly basic (neutral: pH = 7). On the whole, the
salty ocean may be understood as the end product of emission of
acid gases from volcanoes (hydrochloric, sulfuric, and carbonic
acid) and the leaching of common silicate rocks of oceanic and
continental crust, the rocks having minerals of the form [Me
Sia Alb Oc], where Me stands for
the metals Na, K, Mg, and Ca and the remainder makes insoluble
silica-aluminum oxides, that is, clay minerals.
How stable was the composition of
seawater through geologic time? If we use the above concept of
acid-base titration and assume that seawater is a solution in
equilibrium with the sediments on the seafloor, the result is that
the composition was rather stable. Paleontologic evidence certainly
agrees with this assessment. Already in the early Paleozoic, there
are organisms in various groups whose closest modern relatives have
rather narrow salt tolerances: radiolarians, corals, brachiopods,
cephalopods, and echinoderms. Naturally, we cannot rule out
adaptation of the organisms in question to a changing salt content
of the sea. There is no guarantee that the ratios of leached rocks
did not change considerably through time. Thus, it is quite likely
that the composition of seawater did change, contrary to the
equilibrium argument, which implicitly assumes that present
conditions are a permanent feature of the sea.
Comparison of solutes in river water
and in seawater suggests that siliceous acid is removed from river
water as it enters the sea. Diatom production is especially high at
river mouths. This is one aspect of the observation that the ratios
in the river influx of dissolved matter to the sea tend to be
irrelevant to the makeup of sea salt. In the simplest terms, very
soluble salts are abundant in seawater and the others are not. For
example, iron is removed rapidly and efficiently, keeping
concentrations very low. The implication is that the common iron
compounds (hydroxides and sulfides) are not very soluble.
4.3.2 Interstitial Water and Diagenesis
Fine-grained
sediments (clays and silts) have porosities (i.e., water content)
of 70–90% by volume when first deposited on the seafloor, while
sands have around 50%. As the sediments are buried, pore space is
reduced by compression. The water expelled does not necessarily
have the same composition as the water trapped originally. Instead,
the expelled water has contents reflecting reactions within the
sediment, reactions that correspondingly change the chemistry of
the solids accumulating. Compaction and chemical reactions
involving pore fluids (or solids only, as in recrystallization)
constitute diagenesis, the
process that ultimately transforms sediment into rocks.
Typically,
diagenesis is most active in the uppermost meter or so of freshly
deposited sediments, and the chemical reactions within the sediment
are commonly driven by the reactions of organic matter. In fact,
redox reactions usually
dominate the process (called “early diagenesis”). They depend
greatly on the organic carbon present. The oxidation of organic
carbon leads to removal of dissolved oxygen from pore waters.
Additional oxygen demand, after such removal, is satisfied by
stripping oxygen from dissolved nitrate and from solid iron oxides
and hydroxides (e.g., coatings on grains). If the demand is strong
enough, sulfate also is stripped of its oxygen, resulting in an
abundance of hydrogen sulfide (which produces the foul-smelling
hydrogen sulfide, as well as the ubiquitous gold-colored iron
sulfide or “fool’s gold”). The redox reactions within sediments are
mediated by bacteria and archaea, which are in evidence even deeply
below the surface (Fig. 1.11), especially in the organic-rich
deposits below the coastal ocean.
Once the sediment is buried,
concentrations within the interstitial waters can increase to the
point where re-precipitation must occur. This leads to cementation of remaining grains by
carbonate and silica cements (most commonly calcium carbonate and
also quartz with microscopic grains). These various processes,
including recrystallization (crystal growth within preexisting
solids), can be reconstructed by studying the distribution of the
elements and compounds involved, within the solids and the
interstitial waters. The distribution of certain isotopes (oxygen,
carbon, strontium, and others) is of special interest in this
context, because dissolution, migration, and re-precipitation under
various conditions can result in altered ratios of the isotopes.
Diagenetic processes are extremely important for making hydrocarbon
source rocks, of course, and for migration and for reservoir rock
porosity and permeability (see Chap. 14).
4.3.3 Methane
Escaping pore waters carry information
from the redox reactions: they are enriched in gases such as
methane (from fermentation of organic compounds), carbon dioxide
(from plain oxidation of organic matter), ammonia (from reaction of
water with oxygen-stripped nitrate). Methane (Earth gas) can react with water (given
low temperature and high pressure) to make methane clathrates (Fig. 4.5).
Fig.
4.5
Stability field of methane ice
(left panel, a) and application to sedimentary
deposits (right panel,
b). A bottom-simulating
reflector (BSR) can occur at the bottom of the layer bearing
methane hydrate. At the BSR sound is strongly reflected. The
distribution of clathrate is restricted; its presence requires
appropriate temperature and pressure to ensure survival if
clathrate is present, in addition to a source of methane (stability
field and example of occurrence after Erwin Suess (Kiel and
Corvallis) and Gerhard Bohrmann (erstwhile Kiel, now Bremen) in G.
Wefer and F. Schmieder (eds.) 2010. Expedition Erde (3rd ed.),
MARUM, Bremen Univ. Here modified for clarity). A high supply of
organic matter is necessary to make methane. For clarity, the
clathrate presumably forms within slope sediments, not in
water
The stability of methane clathrate
depends on temperature and pressure; hence the temperature below
the seafloor is of enormous importance in defining the depth at
which the hydrate becomes unstable. Correspondingly, the relevant
graph (Fig. 4.5, right panel) emphasizes that hydrates will
not be found at great depth within certain sediment stacks, owing
to the rising temperature below the seafloor. Since methane implies
the presence of large amounts of organic matter for fermentation,
the marine methane is likely to be found fairly close to continents
(in the coastal ocean); that is, much of the seafloor likely does
not have it. Nevertheless, the amounts of carbon fixed in methane
ice are thought to be enormous, exceeding the carbon in the world’s
estimated coal reserves.
The evidence for methane storage
within sediments includes pieces of methane ice (burning when lit, in spite
of its icy nature; Fig. 1.10), cold seeps with its unusual fauna and
microbes, bottom-simulating
reflectors (BSRs),
and chlorine-poor water from melting, occasionally obtained during
drilling. In addition, escaping methane can produce mud volcanoes on the seafloor (Fig.
4.6), features
that are identified in seismic profiles or by acoustic side
scanning. Sizes of the gas-produced mud volcanoes vary; smaller
ones are known as pock
marks.
Fig.
4.6
Mud volcano in an acoustic side-scan image.
Side-scan record is from the bottom of the Black Sea (Courtesy Dr.
Glunow, Moscow; see UNESCO-IMS Newsletter 61, Paris)
Besides redox
reactions, the dissolution (and also the re-precipitation) of
carbonate and of opal is of prime importance in diagenesis. During
early diagenesis much of
the dissolved matter can leave with the escaping pore water or
depart simply by diffusion out of the sediment into the overlying
water.
4.3.4 Residence Time
For steady state conditions, output
must equal input of a geochemical system. Thus, if seawater
composition is to stay constant, seawater has to rid itself of all
new salts coming in, in a “sink.” Where then are the sinks for this material? We might first
look to the sediments for maintaining the balance, but this is not
the whole story. In fact, the quantitative assessment of sinks is a
major geochemical problem. For calcium carbonate the sink indeed is
largely calcareous shells and skeletons built by organisms, and for
silica it apparently is opaline skeletons. Metals presumably leave
the ocean largely in newly formed minerals such as authigenic clay
(produced in place), oxides, sulfides, and zeolites, as well as in
products resulting from reactions between hot basalt and seawater
near the crest of the mid-ocean ridge. Sulfur is precipitated in
heavy metal sulfides in anaerobic sediments near the land and as
gypsum in restricted basins. Some salt leaves with pore waters in
sediment. Under the assumption that seawater maintains its
composition, we can calculate the average time a seawater component
remains in the sea before going out. This time is called
residence time.
Calculating this time is analogous to
figuring out how long people are staying in a museum from counting
the people present and the number of people entering per unit time.
The ratio is the average viewing time (commonly between half an
hour and somewhat greater than an hour, within a museum):
(4.1)
where A is the number present and
r is the rate of input. A
successful exhibit has a large t. In the sea, the residence time is,
in essence, a measure of geochemical reactivity. Sodium and
chloride have a very long residence time, while silica has a short
one. Sodium, having found its partner chloride, tends to stay in
solution, while silica is readily precipitated and becomes close to
inert as a consequence of diagenesis. Equation (4.1) was once used to
calculate a salt age for
the ocean, assuming that the ocean started out as a freshwater body
and retained the sodium added. It was then a useful concept as an
estimate for the scale of geologic time. The result came out near
100 million years, much longer than many other estimates and closer
to the truth, therefore. However, it was still very short of
reality: the calculation was flawed. (Earth is more than 40 times
older than 100 million years; the age of the ocean remains unknown,
but presumably exceeds a factor of 30 over the salt-age guess,
based on various clues.)
4.4 Major Sediment Types
4.4.1 General
There are essentially three types of
marine sediments: those that come into the ocean as particles, are
dispersed, and settle to the seafloor, those that are precipitated
out of solution inorganically, and those that are precipitated by
organisms. We call the first type lithogenous, the second hydrogenous,
and the third biogenous (see Box 4.1: Classification of Marine Sediments).
Around the ocean margins, lithogenous
sediments are predominant, their source being largely mechanical
weathering of continental rocks. Also, there are salt deposits here
(hydrogenous) and biogenous sediments, but in relatively low
abundance. On the deep seafloor, on the other hand, biogenous
sediments (“oozes”) dominate (Fig. 4.7), with calcareous ooze being much more common
than siliceous ooze. In
fact, calcareous ooze covers roughly one half of the seafloor.
Defining components are (silt-sized) coccoliths and (sand-sized)
foraminifers and very small mollusks for the calcareous ooze and
(silt- to sand-sized) diatoms and radiolarians for the siliceous
ooze. The organisms involved are eukaryotic microbes, with
coccoliths and diatoms photosynthesizing (i.e., these are microbial
algae).
Fig.
4.7
Components of “ooze.” (Left side; calcareous ooze;
upper panel:
coccolithophores and coccoliths (The former courtesy G. Wefer, and
R. Norris; the latter mainly after A. McIntyre, Lamont);
lower panel: foraminifers
(SEM by M. Yasuda, S.I.O.). Right: components of siliceous ooze
(diatoms and radiolarians; from E. Haeckel, 1904). Colors: calc.
particles buff to very light
gray; diatoms greenish
brown; radiolarians glassy. Sizes: silt, except the
foraminifers, many of which are sand. Coccoliths commonly are
extremely fine silt
4.4.2 Sand
Marine sediment particles mainly come
in three sizes: sand, silt, and clay (Box 4.1, Fig. 5.4; Appendix
A5). To geologists “sand” denotes solid particles between 0.064 and
2 mm in diameter, no matter what it is. Under the microscope, we
can see that beach sand in Southern California (Fig. 4.8) and in most places
elsewhere mostly consists of mineral grains, small pieces of rock,
and shell fragments. We have known for more than half a century
that sand grains are readily transported far offshore into the
realm of continental fans and continental rises by turbidity
currents running down submarine canyons, as discussed in the
previous chapter. The same process works with sediments of the deep
sea, of course, as long as the seafloor is sloped, as along the
flanks of the MOR, except that we largely deal with oozes here that
are being displaced. Foraminifers are largely sand (the smallest
ones are silt size).
Fig.
4.8
Beach off S.I.O., in La Jolla. Left: pebbles. Right: sand grains. The beach sand has
an abundance of (glassy) quartz grains (proportionally many more
than produced by erosion). Quartz is resistant to chemical attack
and to abrasion, other common minerals less so. The pebbles are
igneous rocks from the cliffs (Photo of pebbles W.H.B.; microphoto
courtesy of P.A. Anderson, S.I. O)
Beach sediments are mainly mixed
lithogenous and biogenous.
4.4.3 Classification
Box 4.1 Classification of Marine Sediment
Types
Lithogenous sediments. Particles
derived from preexisting rocks and volcanic ejecta. Nomenclature
based on grain size and various properties including composition,
structure, and color. Typical examples, with most common
environment in parentheses:
-
Organic-rich clayey silt with root fragments (marsh)
-
Finely laminated sandy silt with small shells (delta-top)
-
Laminated quartzose sand, well sorted (beach)
-
Olive-green homogeneous mud rich in diatom debris (upper continental slope)
(Mud is the same as terrigenous clayey silt or silty clay, commonly with some
sand.)
Fine-grained lithogenous sediments are
the most abundant by volume of all marine sediment types (about two
thirds), largely because of the great thickness of sediment in
continental margins.
Biogenous sediments. Remains of
organisms, mainly skeletal parts (calcium carbonate from mollusks,
calcareous algae, coral, foraminifers, coccolithophorids, etc.),
hydrated silica from diatoms and radiolarians, and calcium
phosphate from arthropods and vertebrates. Organic sediments, while strictly
speaking biogenous, are commonly considered separately. Arrival is
as particles (some precipitated in situ) or in aggregates,
dispersal by waves and currents. Redissolution is common, both on
the seafloor and within the sediment. Appellation is by organism
source and by chemical composition and by other properties.
Examples:
-
Oyster bank (lagoon or embayment)
-
Shell sand (tropical beach)
-
Coral reef breccia (fragments, reefal debris)
-
Oolite sand, well sorted (strand zone, Bahamas)
-
Light gray to buff calcareous ooze, bioturbated (deep seafloor)
-
Greenish-gray siliceous ooze (deep seafloor)
Biogenous sediments are widespread on
the seafloor, covering about one half of the shelves and more than
one half of the deep ocean bottom, for a total of around 55%. About
30% of the volume of marine sediments being deposited at the
present time may be labeled biogenous, although there may be
considerable admixture of lithogenous material, in part
volcanogenic.
Hydrogenous sediments. Precipitates
from seawater or from interstitial water, especially within freshly
deposited sediments (early
diagenesis). Redissolution is common. Nomenclature is based
on origin and chemical composition, as well as some other obvious
properties. Examples:
-
Laminated translucent halite (salt flat)
-
Finely bedded anhydrite {i.e., calcium sulfate} (Mediterranean basin, subsurface)
-
Nodular grayish white anhydrite (ditto)
-
Manganese nodule, black, mammilated, 5-cm diameter (deep seafloor, Pacific)
-
Phosphatic concretion, irregular slab, 5-cm thick, 15-cm diam., light brown to greenish, and granular (upwelling area, upper continental slope)
Hydrogenous sediments are widespread
but not important by volume at present.
4.5 Lithogenous Sediments
4.5.1 Grain Size
The bulk of sediment around the
continents consists of debris washed off the elevated areas of the
continents. The debris results from mechanical breakup with or
without chemical attack on continental igneous and sedimentary
rocks. Transportation of the resulting rock fragments and minerals
is by rivers. Thus, the process is part of the hydrological cycle
(Fig. 4.9).
Fig.
4.9
Origin of lithogenous sediments. Weathering
of source rock and river transport as part of the hydrologic cycle.
The situation schematically depicted is typical for the western
coast in North America, as well as for many other places where
erosion of mountains delivers sediments to the sea (W.H.B., 2013.
San Elijo Lagoon, UCSD, modified)
Grain size is extremely important for
assessing source and transport processes. One distinguishes
gravel, sand, silt, and
clay (see Hjulström
Diagram; Fig. 4.3). Sand, as mentioned, is all solid
material of a size between 0.063 and 2 mm, regardless of
composition or origin. Silt is the next smaller size category, at
sizes between 0.063 and 0.004 mm (see Appendix A5). The next
smaller category, clay, has particles smaller than 0.004 mm (or 4
μm). In some scales 2 μm is taken as the upper clay limit. The term
“clay” is somewhat confusing, clay
minerals being sheetlike minerals of a certain type
regardless of size (albeit being common within the size category of
“clay”). For clarity, “pebble,” “sand,” “silt,” and “clay” are
size categories to a
geologist. The material needs naming if it is to be
specified.
Overall there is a gradation of grain
size from source to place of deposition, with coarser particles
(including boulders) closer to the source and clay-sized material
far away, commonly carried off by water and wind. Gravel-size (2–256 mm) and boulder-size (>256 mm) material does
not commonly travel far, except when taken along by ice (which is
hardly sensitive to the size of the load and thus gives rise to
erratics, that is, boulders
without an obvious source in strange surroundings. Except around
reefs and in high latitudes, boulders and gravel are not common in
marine sediments.
4.5.2 Lithic Sand
Lithogenous sands are typical of
certain beach and shelf deposits, as in Southern California, for
example (Fig. 4.8). The sand may consist of minerals (commonly
dominated by the resistant quartz) or of rock fragments, as is the
case for volcanic material. A striking example of the latter is
provided by black beach sands in the Hawaiian Islands (Fig.
4.10). Also,
green mineral sand of volcanic derivation can be found on occasion.
Other volcanic islands such as Iceland and the Galapagos Islands
also have dark gray lithogenous beaches – light brown beaches are
largely of continental origin, while white beaches are typical for
carbonate environments.
Fig.
4.10
Lithogenic volcanic dark gray beach sand on
Hawaii. Note volcanogenic gravel and boulders in background. The
young coconut tree presumably brought in by waves as seedling
(foreground) is unlikely to survive on the shifting sand (Photo
W.H. B)
The source areas and dispersal history
of sands can be explored by noting the compositional types of
heavy minerals (densities
>2.8 g/cm3; examples are the silicates hornblende,
pyroxene, and olivine; also the iron-rich compounds magnetite and
hematite; and the titanium-bearing ilmenite and rutile (see
Appendix A4)). The heavy mineral association allows the mapping of
depositional provinces. In turn, such mapping provides clues to the
action of shelf current wave climate and other factors.
The shape of sand grains has been used to
obtain clues about their origins. For example, sharp edges have
been linked to recent mechanical action, while rounding has been
ascribed to reworking. Problems arise with recycled sand grains
(i.e., polycyclic sand) and
with post-depositional etching of grains within sediments.
4.5.3 Lithic Silt
Lithogenous silt is common on the
continental slope and rise, although largely mixed with coccoliths
and other biogenous particles (Fig. 4.11). The composition of
the silt is easily rationalized by comparing with adjacent sand
deposits and with associated clay. Mica, a platy terrestrial
mineral delivered both by igneous and ancient sedimentary rocks, is
commonly especially well represented.
Fig.
4.11
Mixed terrigenous and biogenous sediment
makes up the fine silt fraction (2–6 μm) on the continental rise
off Cape Verde (NW Africa). The medium-size silt (right panel) likewise consists of a
mixture. c coccoliths,
f foram shell, m mica, q quartz (SEM photos courtesy D.
Fϋtterer, then Kiel)
Sand is commonly studied with a
binocular microscope, while the composition of clay is investigated
by X-ray diffraction and other sophisticated methods. The study of
silt used to fall into the crack between the methods. It has had a
rather low popularity rating. For the last several decades, the
scanning electron microscope (SEM) has made it possible and
attractive to investigate this size fraction in some detail. Not
surprisingly, it turned out that the composition of the silts is
commonly closely related to that of the associated fine sand
fraction. Biogenous contributions (plankton remains of the coastal
ocean) can be dominant in places (including diatoms and
radiolarians), rather than continental debris (Fig. 4.11).
4.5.4 Clay-Sized Sediment
Clay-sized particles are ubiquitous
both on the continental margins and on the deep seafloor. Much like
fine silt, the presence of clay, where abundant, indicates
low-energy environments.
Clay is easily transported, although, in places, pickup is hindered
by a bacterial mat covering the sediment. In well-oxygenated
environments, however, such mats are commonly disturbed by the
churning of sediment (bioturbation) by larger organisms, if
they form at all.
Common clay minerals are montmorillonite (or
smectite), illite, chlorite, and kaolinite. When mapping their
distributions on the deep seafloor, patterns emerge that are
readily interpreted in terms of origin and paleoclimate, with links
to volcanic activity, metamorphic processes, and physical erosion
as well as deep chemical weathering on land (Chap. 10).
The high surface areas of clay
particles that come with their minute size give clayey sediments
special chemical properties. For example, clays readily absorb a
large variety of substances, and many react easily with the ions in
seawater and in interstitial waters, establishing chemical
equilibrium with their surroundings. During diagenesis within
buried sediments, new clay minerals can form at elevated
temperature and pressure. Ultimately such reactions have important
implications for the chemistry of seawater and for geochemical
processes in general.
Clayey sediments
in regions of high sedimentation rate are commonly rich in organic
matter, partly because organics cling to the clay during deposition
and partly because where conditions are quiet enough for the
deposition of clay, they also are favorable for the deposition of
fluffy organic particles. Much of the clayey matter actually may be
brought to the seafloor within fecal pellets of organisms filtering
the water. Such clay is associated with food particles. Experience
with sediment-trapping equipment (in the Baltic and off California,
among other places in the sea) suggests that the fecal pellet transport mechanism is a
significant agent of sedimentation.
4.6 Biogenous Sediments
4.6.1 Types of Components
Organisms produce sediments in the
form of shells and other skeletal materials and organic matter. The
label “biogenous” is generally applied only to the hard parts, that
is, to calcareous, siliceous, and phosphatic matter. The organisms
involved actually include bacteria and archaea (e.g., in producing
metal oxides, hydroxides, and sulfides). Easily recognized as
fossils are diatoms, radiolarians, silicoflagellates, and primitive
multicellular forms such as sponges (hydrated silicon oxide).
Coccolithophores, foraminifers, various kinds of algae, mollusks,
corals, bryozoans, certain brachiopods, arthropods, echinoderms,
annelid worms, calcareous sponges, and vertebrate remains (i.e.,
solids made of calcite, aragonite, Mg-calcite, and those made of
calcium phosphate minerals) are especially common.
Perhaps the most abundant and
certainly the most conspicuous biogenous types of sediment are
delivered by calcareous skeletal parts of organisms. In shallow
tropical waters, there are abundant remains of coral and associated
materials, depending on location (Fig. 4.12). On the deep
seafloor, there are the remains of coccolithophores and of pelagic
foraminifers (Fig. 4.7). The shells of benthic foraminifers and
coccoliths are ubiquitous and are especially abundant in shallow
water. The benthic “forams” show great variety (Fig. 4.13) and are useful
therefore in stratigraphy and in environmental studies. We have
already pointed out the strong biogenic aspect of low-latitude silt
in slope sediments and on the continental rise. In places, the
biogenic component can be truly striking, as in certain shell
pavement deposits in the “wadden” of the North Sea and in similar
tidal flats.
Fig.
4.12
Biogenous sediments on the beaches in the
Hawaiian Islands. Left,
pebble-size material; middle, coarse sand; right, beach, dark gray: volcanogenic
pebbles among coral-derived pebbles thrown up by storm waves
(Photos W.H. B)
Fig.
4.13
Shells from benthic foraminifers,
illustrating the great variety in these forms (From E. Haeckel,
1904. Kunstformen der Natur. Leipzig)
One source of carbonate that is
ubiquitous, involving both shelves and the deep seafloor, is a
tremendous variety of foraminifers, with hundreds of common species
in the benthic types (Fig. 4.13). Pelagic forms, with relatively modest
diversity (tens of species), may have arisen several times during
geologic history from benthic stock.
For siliceous (opaline) hard parts,
the gradient pattern from shallow to deep is similar to the
carbonate pattern: remains of benthic organisms dominate in shallow
water (sponges, benthic diatoms) and pelagic materials dominate in
sediments of deep water (diatoms, radiolarians). Pelagic diatoms
are typical for the coastal ocean (overlying shelf and upper slope)
and especially for upwelling areas. Likewise, river mouths are
prime areas of diatom deposition, especially benthic ones.
In addition to carbonate and silica,
phosphatic particles are produced by organisms (including
phosphatic fecal pellets). The sedimentation of such particles
plays an important role in the phosphorus budget of the biosphere
and is of great interest geochemically, therefore. The various
other types of hard parts – strontium sulfate, manganese, iron, and
aluminum compounds – are intriguing in the context of evolutionary
studies but do not materially contribute to marine sediments.
Organogenic sediment is important in the context of the
productivity of the sea (Chap. 7) and of hydrocarbons (Chap.
14).
4.6.2 Sediment Contributions by Benthic Organisms
The great importance of
carbonate-secreting benthos in producing sediment is apparent all
through the Phanerozoic record, that is, more than the last half
billion years. There is some indication that the less stable
carbonate minerals aragonite and Mg-calcite are more abundant
during cool and cold conditions (latest Precambrian, latest
Paleozoic, Neogene) than during warm time periods such as much of
the Cretaceous. At present, the most conspicuous biogenic edifice
in the sea is the Great Barrier Reef. It is largely composed of the
remains of benthic organisms: corals, calcareous algae, mollusks,
and benthic foraminifers. Among the calcareous matter, both calcite
and aragonite (same chemical composition, different arrangement of
elements) are found. Aragonite is much more soluble than calcite.
Also, there is a tendency for higher Mg content in the calcite
precipitated in warm water, compared with pure calcite.
Mesozoic and later Cenozoic platform
carbonates once were deposited all around the Tethys, the ancient tropical seaway
that once linked the western Pacific with the central Atlantic. The
shallow-water limestone rocks of the geologic record, which
originated in the shelf environments of ancient oceans, commonly
contain an admixture of siliceous rocks, as layers of chert arranged along horizons parallel
to the bedding. The “chert” usually appears as microcrystalline
quartz lumps and beds and originated from recrystallization after
expulsion of the water from the opal. Any of the silica-producing
organisms can be responsible for the origin of a given piece of
chert, even though some fossils (e.g., sponge remains,
radiolarians) may be more conspicuous within it than other
fossils.
Why is there little or no evidence for
incipient chert formation in modern shelf carbonates?
The answer is not clear. The reason
may be that silicate concentrations are comparatively low in
present tropical waters. Seawater (especially shallow seawater
making shelf carbonates) commonly is highly undersaturated,
presumably stripped of its silicate by diatoms in upwelling
regions. In fact, a general rarity of chert (as is evident after
the Eocene) may be an inevitable consequence of planetary cooling
and associated upwelling. In any case, the once popular idea that
waning volcanism is to blame for the post-Eocene rarity of chert
has been largely abandoned. Eocene chert in deep-sea sediments
attracted much attention by deep-sea geologists when deep-sea
drilling started, its presence being rather inimical to any
drilling.
4.6.3 The Remains of Planktonic Organisms
Remains of pelagic species are much in
evidence in slope sediments, as we pointed out earlier when
discussing silt and Fig. 4.11. The bulk of deep-sea sediments, to be
sure, is rich in planktonic skeletal matter. Shelf sediments as
well can contain considerable amounts of planktonic remains. Thus,
for example, the English chalk of Cretaceous age is extremely rich
in coccoliths, that is,
particles produced by coccolithophores (more commonly
addressed as nannofossils
by geologists). Siliceous
remains of diatoms, radiolarians, and silicoflagellates (all
planktonic and opaline) are characteristic of offshore
high-production areas, as mentioned. This general pattern must be
added as a clue to distance from the shore to the gradient in the
benthic-planktonic ratio in foraminifers that parallels distance
from the shore. In deep-sea sediments, planktonic remains
(specimens, not species) tend to outnumber benthic ones by ratios
of 10:1 to 100:1, the higher values being typical in areas of good
preservation.
4.7 Nonskeletal Carbonates
4.7.1 Carbonate Saturation and Precipitation
In the present ocean, carbonate
precipitation occurs either within organisms (shells, skeletons,
“internal precipitation”) or in association with their metabolic
activity (e.g., algal crusts, “bacterially mediated precipitation,”
“external precipitation”). This need not always have been the case:
at very high levels of saturation, there might have been inorganic
precipitation, which may have produced certain types of limestone
seen in the geologic record. Where would one look for possible
inorganic precipitation today? One would need to look in regions of
unusually high carbonate saturation. Seawater that spontaneously
precipitates a mineral is said to be supersaturated with this mineral phase.
In contrast, seawater that dissolves the mineral is undersaturated. Saturation obtains at
the point of balance between precipitation and dissolution in a
solution. For two-ion compounds (such as calcium carbonate), the
degree of saturation is expressed as the ratio between the products
of observed concentration of the two ions involved and the product
of concentrations at saturation. This criterion defines much
shallow tropical seawater as being supersaturated with calcium
carbonate. Yet spontaneous inorganic precipitation is not observed.
Presumably the expected reaction does not occur because of
interfering factors, that is, the presence of a blocking agent or
several such agents.
In the presence
of interference, if inorganic precipitation is to be observed,
especially high levels of supersaturation need to be extant (and
any blocking agent is to be removed), assuming that appropriate
nuclei for crystal formation are sufficiently abundant. Naturally,
when saturation is lowered substantially, for example, by the
addition of carbon dioxide (in modern times from human energy use),
it becomes more difficult to make carbonate sediment or shell. This
is, in a nutshell, the effect from the acidification of seawater. When
carbonate is dissolved, carbon dioxide is used up in the reaction.
Conversely, when precipitating carbonate, carbon dioxide is
released and any undissolved excess is at least partially expelled
from the sea to the atmosphere. Evidently, then, the reactions
involving carbonate are of prime importance when discussing changes
in the atmospheric concentration of carbon dioxide and associated
greenhouse effects on various time scales.
4.7.2 The Bahamas
One place that seemed ideally suited
for finding out whether inorganic precipitation is now occurring or
not is the archipelago known as “the Bahamas.”
The islands are surrounded by some of
the warmest and most alkaline waters in the ocean. There is hardly
any terrigenous input, so that pure carbonates can accumulate. The
Great Bahama Bank is very flat and very shallow over large areas
(Fig. 4.14).
Evaporation and warming by the sun combine to increase saturation.
Ubiquitous benthic algae remove carbon dioxide when growing. Thus,
conditions would seem very favorable indeed for the precipitation
of carbonate directly from seawater.
Fig.
4.14
Carbonate oolite deposits on the Bahama
Banks, seen from the air (Air photo courtesy D.L. Eicher, Boulder,
Colorado)
Two types of calcareous particles on
the seafloor of the Great Bank have been considered as possible
products of inorganic precipitation: aragonite needles (length of a few
micrometers) and oolites
(consisting of spherically layered particles (ooids) with a diameter near one third
of a millimeter) (Fig. 4.15). The origin of the aragonite needles has
puzzled geologists for some time; apparently they form within
certain algae. Oolites are a favorite object of geologic
discussion, famous for being mentioned by Mark Twain (in an ironic
remark on the quality of geologic reasoning). Oolites are abundant
in the geologic record, especially in subtropical shelf sediments.
In the Bahamas they occur especially on the outer rim of the Great
Bank in the shallowest water, suggesting a strong influence of
tidal currents and breaking waves in their origin. Apparently the
ooids only form in the presence of sufficient organic matter. This
is not good as support for the action of inorganic precipitation.
In fact, observations now suggest that precipitation is largely
through biocalcification by unicellular algae. It appears, then,
that even here in the Bahamas precipitation directly from seawater
is negligible in present-day conditions.
Fig.
4.15
SEM images of ooids, scale bar 5 μm.
Left, slightly etched
section. Three secondarily filled borings (marked “b”) intersect
the concentric laminae of primary oolitic coating. Right, close-up of oolite laminae
showing acicular aragonite needles (Images courtesy of D. Fϋtterer,
Kiel)
4.7.3 Dolomite
Calcite may contain magnesium in
different concentrations (magnesium calcite). In the present
shelf environment, as mentioned, magnesium-rich calcites are found
in greater abundance than pure calcite. Dolomite, a carbonate mineral with
equal amounts (by atom numbers) of magnesium and calcium, is a
different story. It does not precipitate in shells or other
biogenic products. Its presence in the geologic record poses many
unsolved questions. Thus one might expect that dolomite would
precipitate from seawater because it is much less soluble than
aragonite. There is, after all, plenty of magnesium in seawater.
However, such precipitation has not been observed. Instead,
dolomite apparently forms within sediments, either through
partial replacement of calcium with magnesium in preformed
carbonate or possibly also by precipitation from pore waters.
Either process would be referred to as “diagenesis.” Much of
“diagenesis” is about postdepositional geochemical problems. Thus,
geochemists have done much fundamental work on dolomitization
(Scripps examples: M. Kastner, P. Baker). In recent years,
petroleum and methane geologists and chemists have shown much
interest in the topic.
A classic area
for the study of incipient dolomite formation is the southern
margin of the Persian/Arabian Gulf. Here the intertidal flats lie
behind barrier islands within lagoons with warm and very saline
water. In the pore waters of sediments above low tide, magnesium is
greatly enriched with respect to calcium, while sulfate (which
apparently hinders dolomite formation in many surroundings) is
reduced in abundance owing to the precipitation of gypsum (calcium
sulfate) in adjoining evaporite pans (which are common in the
sabkha environment).
Microbial sulfate reduction in the uppermost few tens of meters of
continental margin sediments similarly seems to favor
“dolomitization.” One problem interfering with solving the
dolomitization question is that unknown conditions of the geologic
past and in a warm ocean may have played a large role in some of
the dolomite formation.
4.8 Hydrogenous Sediments
4.8.1 Marine Evaporites
Evaporites form when seawater
evaporates. Desert belts – high ratios of evaporation rates over
those of precipitation – are located near 25° of latitude, on both
hemispheres. Unsurprisingly, these also are the latitudes of the
highest salinity in the surface waters of the open ocean. For
precipitation of salt, however, concentrations have to exceed
saturation values, and this happens only upon restriction of
exchange of shelf-water bodies with the open ocean, commonly in a
semi-enclosed basin on the shelf (the one widely quoted exception
being the Mediterranean basin at the end of the Miocene, as
documented by deep-sea drilling during DSDP Leg 13.
How much salt can be produced by
evaporating a 1000-m-high 1-m2 column of seawater? Salt
is 3.5% (or 35 per mil) of the weight of the column; thus the
answer is 35 tons or 14 m at the common density of 2.5 tons per
cubic meter. Most of the readily recognized salt obtained would be
kitchen salt (“halite”). The salts precipitating first would not
include halite, though, but have abundant carbonates and
sulfates.
To precipitate halite the brine would
need to be concentrated about tenfold. Many evaporites only contain
carbonate and gypsum (or “anhydrite”), whereas others have thick
deposits of halite or (even more rarely) final layers of the
valuable (and very soluble) potassium salts. In marine sediments,
to get any one salt without the others, “fractionation” in linked
basins is necessary or periodic removal of certain salts by
dissolution, while preserving others.
4.8.2 Phosphorites
Phosphatic deposits presumably are
largely of biogenic origin. Phosphorites are of prime importance in
the cycling of phosphorus and hence in the productivity of the
ocean (Chap. 7). Also, they constitute an
important marine resource, used predominantly in the fertilizer
industry. Thus, phosphorites are encountered again in the chapter
on resources (Chap. 14).
4.8.3 Iron Compounds
Microbes are notably involved in the
origin of both iron sulfides and iron hydroxides, as has been known
for more than a century. Iron apparently is of prime importance in
the productivity of the ocean, with at least some of it mobilized
from the seafloor upon loss of oxygen. Sulfides and hydroxides are
important items in the geochemical cycling of temporarily free
oxygen, the abundance of which changes in Earth history, with
consequences for marine sedimentation (black mud, clay, and shale
and greenish deposits versus brown, reddish, and yellowish
deposits).
An iron-bearing
mineral that has attracted much attention by marine geologists is
glauconite. It is a
greenish silicate common in shallow marine areas and is commonly
found in association with phosphatic sediments in high-productivity
regions along certain continental margins, as, for example, off
Angola.
The origin of iron oolites, common in
Mesozoic marine sediments (e.g., “minette” ores of
Alsace-Lorraine), is not known. Apparently minette ores are not
forming in the present ocean in noticeable abundance.
4.9 Sedimentation Rates
The idea of
geologic time, so
fundamental in all of geology, is in fact quite young, compared
with the age of various branches of science. It was early discussed
by James Hutton (1726–1797), and its chief protagonists were
Charles Lyell (1797–1875) and Charles Darwin (1809–1882). Lyell,
like Hutton, invoked geologic time to create Earth’s morphology and
the geologic record. Darwin made use of geologic time in his
explanation of evolution. However, before the discovery of
radioactive decay at the end of the nineteenth century and the
application of this discovery to the geologic record, there was no
reliable way of telling just how much the geologic time scale
differs from the chronology derived from multimillennial human
time, that is, the account in Genesis, summarized by Bishop Sam
Wilberforce, FRS. Wilberforce’s clever but derogatory
pronouncements on Darwin’s ancestry were forcefully attacked by
T.H. Huxley, in 1860. The disputations reflected different belief
systems, not certain knowledge.
We now know that the guesses proposing
millions of years of Earth history were closer to the truth than
those postulating thousands of years. Modern determinations of
sedimentation rates span the gamut between 1 m per million years
and many km per million years, depending on the environment (see
Fig. 4.16). A
million years is a useful time interval to work with, for
geologists contemplating pre-ice age processes. To apply the rates
listed to ancient sediments, one has to consider compaction and
loss of porosity – about 40% in sands and about 70% in muds. On the
whole, high rates of sedimentation (say 100 m per million years)
occur at the edges of the continents. Exceptionally high rates are
found off many glaciers. The lowest rates occur on the deep
seafloor far away from continents.
Fig.
4.16
Typical rates of vertical crustal motion,
of denudation, and of sedimentation rates. Scale in mm per thousand
years (meters per million years). Postglacial sea-level rise shaded
(average of some 100 m in 10 thousand years; k kilo years). Logarithmic scale. The
values are approximate and for comparison (E. Seibold, 1975.
Naturwissenschaften 62: 62; modified)
Characteristic values of sedimentation
rates on continental slopes are 40–200 mm per thousand years, with
a typical value near 100 mm per millennium or 100 m per million
years. Theoretically, coral reefs can build up at rates near 1 cm
per year, that is, 10 m per millennium or 10 km per million years!
More commonly, one finds values somewhat less than half of that in
reef growth. In any case, such high rates make the sinking of the
seafloor irrelevant to the growth of stony reefs. Such findings
throw doubt therefore on the commonly quoted Darwinian origin of
atolls.
Reliable estimates of high
sedimentation rates are possible in the case of annual layers or varves. Counting varves in the Black
Sea have yielded rates near 400 mm per millennium. In a bay off the
Adriatic Island Mljet, 250 mm per millennium was found. In Santa
Barbara basin off Southern California, counts of varves yielded
values near 1 mm per year, that is, 1 m per millennium. Such high
rates of sedimentation are favorable for the detailed
reconstruction of environmental changes over the last few millennia
and centuries (Chap. 15).
In concluding
this chapter about the composition of marine sediments, some of the
more esoteric components may be mentioned. So-called cosmic
spherules were first described by John Murray of the Challenger Expedition. Black magnetic
spherical objects up to 0.2 mm in diameter, commonly rich in Fe and
Ni, can be found in pelagic sediments, typically several spherules
per gram of pelagic clay. Because it was thought that they
represent a steady influx of matter from space, their abundance was
used on occasion as an indicator for accumulation rates.
Glassy objects, normally up to 1 mm in
diameter, may be produced by the impact of meteorites on rocks.
They are called microtectites or simply tectites. Tectites were found in
deep-sea sediments in strewn fields around Australia (dated at 0.7
million years), off the Ivory Coast (1.1 million years), and in the
Caribbean (33–35 million years), for example. Where abundant, they
can be used as a time marker and serve for correlation of
sedimentary rocks by event
stratigraphy. Tectites commonly serve as evidence for impact
events.
Suggestions for Further Reading
Funnell, B.M., and W.R.
Riedel (eds.) 1971. The Micropalaeontology of Oceans. Cambridge
Univ. Press.
Hsü, K.J., and Jenkyns, H.C.
(eds.) 1974. Pelagic Sediments: On Land and Under the
Sea. Spec. Publ. Int.
Assoc. Sedimentol., 1:273 299.
Cook, H.E., Enos, P. (eds.)
1977. Deep-Water Carbonate Environments. Soc. Econ. Paleont.
Mineral. Spec. Publ. 25.
Bouma, S.G., W.R. Normark,
and N.E. Barnes (eds.) 1985. Submarine Fans and Related Turbidite
Systems. Springer, Heidelberg.
Anderson, J.B., and B. F.
Molnia, 1989. Glacial-marine Sedimentation. Short Course in
Geology, vol. 9, Am. Geophys. Union, Washington, D.C.
Hemleben, C. M. Spindler and
O.R. Anderson, 1989. Modern planktonic foraminifera. Springer, New
York.
Morse, J.W., and F.T.
Mackenzie, 1990. Geochemistry of Sedimentary Carbonates. Elsevier,
Amsterdam.
Einsele, G., W. Ricken, and
A. Seilacher (eds.) 1991. Cycles and Events in Stratigraphy.
Springer, Heidelberg.
Friedman, G.M., J.E. Sanders,
and D.C. Kopaska-Merkel, 1992. Principles of Sedimentary Deposits:
Stratigraphy and Sedimentology. McMillan, New York.
Heimann, M. (ed.) 1993. The
Global Carbon Cycle. Springer, Berlin &Heidelberg.
Schulz, H.D., and M. Zabel
(eds.) 2006. Marine Geochemistry (2nd edition). Springer,
Heidelberg.