You are on page 1of 23

A Brief Introduction to

TAPHONOMY
©
Gastaldo, Savrda, & Lewis. 1996. Deciphering Earth History: A
Laboratory Manual with Internet Exercises. Contemporary
Publishing Company of Raleigh, Inc. ISBN 0-89892-139-2

Not every organism that ever lived could become part of the fossil record. If you eat an
average of three meals a day, you test and prove this hypothesis daily. A large
percentage of all biological entities end up as food for other organisms higher on the
food chain. This fact alone may prevents these organisms from being preserved. Even
those organisms that avoid being eaten have a low probability of becoming fossilized
because most of them undergo decay and recycling of their chemical components. For
example, you can examine any forest-floor litter and find that beneath the top layer of
leaves, the organic matter has been degraded to an unrecognizable form (humus -- not
hummus, the garlic-laden spread served in health-food restaurants). This recycling
keeps the carbon, nitrogen, and sulfur cycles operating. In fact, many taphonomic biases
impact the odds of any organism being preserved.

The paleontological subdiscipline called Taphonomy, from the Greek taphos (death), is
concerned with the processes responsible for any organism becoming part of the fossil
record and how these processes influence information in the fossil record. Many
taphonomic processes must be considered when trying to understand fossilization.
These include events that affected the organism during life (changes in rainfall,
availability of food, and behavior for maximum growth, etc.), the transferral of that
organism (or a part of that organism) from the living world (biosphere) to the
sedimentary record (lithosphere; compare the death of a herd of vertebrates with the
autumnal leaf fall from a forest), and the physical and chemical interactions that affect
the organism from the time it is buried until the time it is collected in the field.

Any organism must successfully pass through three distinct, and separate, stages in
order to be seen in a museum display. These stages span the entire time from death of
the organism to collection. Necrology is the first stage, and involves the death or loss of
a part of the organism. The vast majority of animals must die before they can become
introduced to the next phase. It's true that if a starfish is cut in half, each half will
regenerate itself. The result will be two animals. Not many animals have this capability.
We suggest that you don't test this hypothesis with your beloved pet. On the other hand,
most plants do not have to die to contribute one or more of their parts to the potential
fossil record. When autumn leaves fall in temperate climates, the trees don't die. The
oldest living organism, bristlecone pines, are more than 5300 years old (as determined
by counting tree rings). Their present leaves are not the same ones that grew 5300 years
ago. When plants disperse their reproductive bodies (spores, pollen, or seeds), most do
not die thereafter. Of course there are exceptions, but these are a small percentage of all
extant (living) plants.

Once an organism has died or sheds a part, all the interactions involving its transferral
from the living world to the inorganic world (including burial) constitute the second
taphonomic stage. This is the Biostratinomy stage. Besides the conspicuous fossil
characteristics that you will be able to observe during this laboratory (those external and
internal features of the fossilized remain), less-obvious details often record what
happened to the organism (or part) before it became a fossil. By studying these details
paleontologists are able to understand, in a Sherlock Holmesian way, the mode of death
or disarticulation (breakup of an organism), any biological processes that may have
modified the remains before burial (such as scavenging), the response of the part to
transport (by animals, water and/or wind), and the amount of time the organism sat
around in the environment before it was finally entombed.

Ultimately the organic matter is buried. Burial plays an important role in potential
preservation of the organic matter. Very specific chemical and physical conditions must
exist in the burial environment to allow preservation in a form recognizable to us. It is
here that biological (e.g., enzymatic and bacterial) and chemical (e.g., enzymatic and
dissolution) processes must be slowed or eliminated. Once buried, the organic material
is subjected to the third taphonomic phase, or Diagenesis. Diagenesis involves all of the
processes responsible for lithification of the sediment and chemical interactions with
waters residing between clasts. The processes of fossilization appear to be site specific
with respect to depositional settings, resulting in a mosaic of preservational traits in the
terrestrial and marine realm. Few fossil assemblages are exactly identical, especially
with regard to the way in which they were formed, but general patterns do exist. An
understanding of taphonomic assemblage features within an environmental context
allows for a more accurate interpretation of the fossil record.

Most organic matter on Earth is used by some organism higher on the food chain and is,
therefore, ultimately recycled. This is the fate of almost all biomass on Earth. Most
organic matter is composed of easily degraded and digested compounds that are not
likely to be preserved even under the most favorable conditions. Those parts of an
organism that are already mineralized (such as your calcium-fortified skeleton) and,
hence have made the first step in the transition to "stone", have a higher probability of
preservation than any of the soft, fleshy tissues either around or within the skeleton. The
early inhabitants of Paris, France, the bones of whom are now stacked neatly in
catacombs beneath the city streets, attest to this fact.

Although the fossil record is incomplete, it still provides a useful survey of the history
of life because of the vast amounts of time represented within the rock record. Even if
the conditions for preserving organic matter existed only once every 10,000 years in
each contemporaneous depositional environment around the globe, a lithology that was
100 meters thick (330 feet) and encompassing 1 million years of time would contain
100 fossil assemblages. Such conditions are not unrealistic, particularly within the
ocean basins. If we then consider contemporaneous depositional settings around the
globe, the number of fossil assemblages that would be preserved during this 1 million
years of time increases dramatically. Of course, not all of these fossil sites are or would
be accessible for collection and study. Mountain-building processes associated with
plate tectonic activity (metamorphism of fossil-bearing sedimentary rock beyond
recognition) and the erosion of these folded (metasedimentary) and faulted
(sedimentary) rocks depletes the number of fossil localities available at the Earth's
surface through time. The quantity of fossiliferous rocks beneath ground level far
exceeds those available at the surface to be sampled and studied. Nevertheless, there are
far more fossils than paleontologists, which will continue to be the case far into the
future. Paleontologists are not wanting in their search for the history of life on Earth.

RADIOMETRIC TIME SCALE


The discovery of the natural radioactive decay of uranium in 1896 by Henry Becquerel,
the French physicist, opened new vistas in science. In 1905, the British physicist Lord
Rutherford--after defining the structure of the atom-- made the first clear suggestion for
using radioactivity as a tool for measuring geologic time directly; shortly thereafter, in
1907, Professor B. B. Boltwood, radiochemist of Yale Uniyersity, published a list of
geologic ages based on radioactivity. Although Boltwood's ages have since been
revised, they did show correctly that the duration of geologic time would be measured
in terms of hundreds-to-thousands of millions of years.

A technician of the U.S. Geological Survey


uses a mass spectrometer to determine the
proportions of neodymium isotopes
contained in a sample of igneous rock.

The next 40 years was a period of expanding research on the nature and behavior of
atoms, leading to the development of nuclear fission and fusion as energy sources. A
byproduct of this atomic research has been the development and continuing refinement
of the various methods and techniques used to measure the age of Earth materials.
Precise dating has been accomplished since 1950.

A chemical element consists of atoms with a specific number of protons in their nuclei
but different atomic weights owing to variations in the number of neutrons. Atoms of
the same element with differing atomic weights are called isotopes. Radioactive decay
is a spontaneous process in which an isotope (the parent) loses particles from its nucleus
to form an isotope of a new element (the daughter). The rate of decay is conveniently
expressed in terms of an isotope's half-life, or the time it takes for one-half of a
particular radioactive isotope in a sample to decay. Most radioactive isotopes have rapid
rates of decay (that is, short half-lives) and lose their radioactivity within a few days or
years. Some isotopes, however, decay slowly, and several of these are used as geologic
clocks. The parent isotopes and corresponding daughter products most commonly used
to determine the ages of ancient rocks are listed below:

Parent Isotope Stable Daughter Product Currently Accepted Half-Life Values


Uranium-238 Lead-206 4.5 billion years
Uranium-235 Lead-207 704 million years
Thorium-232 Lead-208 14.0 billion years
Rubidium-87 Strontium-87 48.8 billion years
Potassium-40 Argon-40 1.25 billion years
Samarium-147 Neodymium-143 106 billion years

The mathematical expression that relates radioactive decay to geologic time is called the
age equation and is:

Dating rocks by these radioactive timekeepers is simple in theory, but the laboratory
procedures are complex. The numbers of parent and daughter isotopes in each specimen
are determined by various kinds of analytical methods. The principal difficulty lies in
measuring precisely very small amounts of isotopes.

The potassium-argon method can be used on rocks as young as a few thousand years as
well as on the oldest rocks known. Potassium is found in most rock-forming minerals,
the half-life of its radioactive isotope potassium-40 is such that measurable quantities of
argon (daughter) have accumulated in potassium-bearing minerals of nearly all ages,
and the amounts of potassium and argon isotopes can be measured accurately, even in
very small quantities. Where feasible, two or more methods of analysis are used on the
same specimen of rock to confirm the results.

Another important atomic clock used for dating purposes is based on the radioactive
decay of the isotope carbon-14, which has a half-life of 5,730 years. Carbon-14 is
produced continuously in the Earth's upper atmosphere as a result of the bombardment
of nitrogen by neutrons from cosmic rays. This newly formed radiocarbon becomes
uniformly mixed with the nonradioactive carbon in the carbon dioxide of the air, and it
eventually finds its way into all living plants and animals. In effect, all carbon in living
organisms contains a constant proportion of radiocarbon to nonradioactive carbon. After
the death of the organism, the amount of radiocarbon gradually decreases as it reverts to
nitrogen-14 by radioactive decay. By measuring the amount of radioactivity remaining
in organic materials, the amount of carbon-14 in the materials can be calculated and the
time of death can be determined. For example, if carbon from a sample of wood is
found to contain only half as much carbon-14 as that from a living plant, the estimated
age of the old wood would be 5,730 years.

The radiocarbon clock has become an extremely useful and efficient tool in dating the
important episodes in the recent prehistory and history of man, but because of the
relatively short half-life of carbon-14, the clock can be used for dating events that have
taken place only within the past 50,000 years.
The following is a group of rocks and materials that have dated by various atomic clock
methods:

Approximate Age
Sample
in Years
Cloth wrappings from a mummified bull
Samples taken from a pyramid in Dashur, Egypt. This date
2,050
agrees with the age of the pyramid as estimated from historical
records
Charcoal
Sample, recovered from bed of ash near Crater Lake, Oregon,
is from a tree burned in the violent eruption of Mount Mazama
6,640
which created Crater Lake. This eruption blanketed several
States with ash, providing geologists with an excellent time
zone.
Charcoal
Sample collected from the "Marmes Man" site in southeastern
10,130
Washington. This rock shelter is believed to be among the
oldest known inhabited sites in North America
Spruce wood
Sample from the Two Creeks forest bed near Milwaukee,
11,640
Wisconsin, dates one of the last advances of the continental ice
sheet into the United States.
Bishop Tuff
Samples collected from volcanic ash and pumice that overlie
glacial debris in Owens Valley, California. This volcanic 700,000
episode provides an important reference datum in the glacial
history of North America.
Volcanic ash
Samples collected from strata in Olduvai Gorge, East Africa,
1,750,000
which sandwich the fossil remains of Zinjanthropus and Homo
habilis -- possible precursors of modern man.
Monzonite
Samples of copper-bearing rock from vast open-pit mine at 37,500,000
Bingham Canyon. Utah.
Quartz monzonite
Samples collected from Half Dome, Yosemite National Park, 80,000,000
California.
Conway Granite
Samples collected from Redstone Quarry in the White 180,000,000
Mountains of New Hampshire.
Rhyolite
Samples collected from Mount Rogers, the highest point in 820,000,000
Virginia.
Pikes Peak Granite
1,030,000,000
Samples collected on top of Pikes Peak, Colorado.
Gneiss
Samples from outcrops in the Karelian area of eastern Finland 2,700,000,000
are believed to represent the oldest rocks in the Baltic region.
The Old Granite
Samples from outcrops in the Transvaal, South Africa. These 3,200,000,000
rocks intrude even older rocks that have not been dated.
Morton Gneiss [see Editor's Note] 3,600,000,000
Samples from outcrops in southwestern Minnesota are
believed to represent some of the oldest rocks in North
America.

Carbon samples are converted to acetylene


gas by combustion in a vacuum line. The
acetylene gas is then analyzed in a mass
spectrometer to determine its carbon
isotopic composition.

Interweaving the relative time scale with the atomic time scale poses certain problems
because only certain types of rocks, chiefly the igneous variety, can be dated directly by
radiometric methods; but these rocks do not ordinarily contain fossils. Igneous rocks are
those such as granite and basalt which crystallize from molten material called "magma".

When igneous rocks crystallize, the newly formed minerals contain various amounts of
chemical elements, some of which have radioactive isotopes. These isotopes decay
within the rocks according to their half-life rates, and by selecting the appropriate
minerals (those that contain potassium, for instance) and measuring the relative amounts
of parent and daughter isotopes in them, the date at which the rock crystallized can be
determined. Most of the large igneous rock masses of the world have been dated in this
manner.

Most sedimentary rocks such as sandstone, limestone, and shale are related to the
radiometric time scale by bracketing them within time zones that are determined by
dating appropriately selected igneous rocks, as shown by a hypothetical example.

Literally thousands of dated materials are now available for use to bracket the various
episodes in the history of the Earth within specific time zones. Many points on the time
scale are being revised, however, as the behavior of isotopes in the Earth's crust is more
clearly understood. Thus the graphic illustration of the geologic time scale, showing
both relative time and radiometric time, represents only the present state of knowledge.
Certainly, revisions and modifications will be forthcoming as research continues to
improve our knowledge of Earth history.

Previous || Contents || Next

This page is URL: <http://pubs.usgs.gov/gip/geotime/radiometric.html>


Last updated June 13, 2001
Maintained by Kathie Watson

http://www.ucmp.berkeley.edu/IB181/VPL/Pres/Pres4.html

The fate of most organic material produced by living systems is to be decomposed


to carbon dioxide and water, and recycled into the biosphere. The circulation of
elements through biogeochemical cycles indicates that decomposition is, indeed,
efficient; however the presence of organic material in sedimentary rocks (e.g., coal,
petroleum, dispersed organic matter, and fossils) shows that some organic matter -
or its traces - escapes these cycles to be preserved in the rock record. The study of
paleobotany relies on this preserved material - fossils - as evidence of past life. In
the early history of modern paleontology, fossils were thought of mostly as static
parts of the rock record ("sticks in mud"). This fostered description and
classification as the main activities of scientific paleontologists. However, a shift in
emphasis to thinking of fossils as "once-living organisms" gave paleontology a more
biological flair and, more importantly, opened a new world of research questions.
One tangible outcome of this philosophical shift is the movement of many academic
paleontologists from Geology Departments to Biology Departments.

Plants become fossilized in a variety of ways. Each type of preservation carries


different information about the once-living organism. Thus, an appreciation of plant
fossils requires that one understand the processes of fossilization, and how each
type of preservation may influence our view of once-living organisms.

The study of how organisms or their parts become fossils is called taphonomy.
Taphonomy is literally everything that happens to an organism - or part of an
organism, as is often the case with plants - from the moment that it dies until it is
collected and curated for scientific study. Figure 3.1 illustrates some of the many
taphonomic pathways a plant or plant part can take from it's living community to
the museum drawer.

Figure 3.1. Some of the many possible fates of plants


or their parts as they enter the fossil record.

Levels of preservation
Plant fossil preservation can take place at a number of levels. Each level contains a
different type of information.

Cellular Level

Not all organic compounds are equally resistant to chemical degradation and decay.
Plant cell walls (composed primarily of the polysaccharide polymer cellulose) are far
more likely to escape decomposition than internal membranes and organelles,
which are rich in protein, lipids and sugars. Preservation of cytological details has
been reported in fossil plants, but occurrences are rare, and most reports of
fossilized nuclei and organelles should be read with caution. Secondary compounds,
such as those impregnating or covering cell walls, can also be resistant to
decomposition; examples include lignin, waxes, cutin (which comprises plant
cuticle), and sporopollenin, which forms the external shell of spores, pollen, and the
resting cysts of some marine algae.

Tissue Level

Decay-resistant materials are distributed differentially throughout the plant.


Consequently, some tissues are more amenable to preservation in the fossil record
than others. With respect to vascular tissue, xylem is often preserved, while phloem
commonly is not. This is because the cell walls of xylem are impregnated with
decay-resistant lignin, while phloem cell walls are cellulosic. Cuticle, composed of
the resistant material cutin and various waxes, is more likely to be preserved than
actual epidermal cells; however, the shape and distribution of epidermal cells,
including guard cells, is faithfully preserved in cuticle (VG 1:2). Spores and pollen
(VG 1:8), because of their resistant spore coats, are the most abundant and
ubiquitous structural remains of vascular plants preserved in the rock record.
Because they are easily preserved and found in great numbers, pollen and spores
(palynomorphs) provide important quantitative data for vegetation reconstruction
and a variety of paleoecological questions.

Organ Level

Plants break apart, both in life (organ senescence or dispersal) and after death;
dispersed parts may be transported before settling into the sediment to be buried
and become fossils. Assemblages of plant fossils that are preserved close to where
their parent-plants originally grew are called autochthonous; assemblages that
have been transported are referred to as allochthonous. Whether an assemblage is
autochthonous or transported has obvious implications for what sorts of ecological
interpretations we can make from it. Therefore, as you study each type of plant
fossil preservation, think about whether fossils produced in these ways are
autochthonous or not.

The differential hydrodynamic properties of different plant organs often results in


segregation of different parts in the fossil record. Some deposits contain only
palynomorphs (e.g., pollen, spores, and a few other do-dads), while others may
contain only large chunks of wood, or scattered leaves. Reproductive organs,
especially flowers, are relatively rare in the fossil record because they are delicate
and full of energy-rich, easily broken-down compounds. Other reproductive
structures are sometimes preserved, however, and constitute a valuable source of
taxonomic and evolutionary information. Fruits and seeds often contain resistant
tissues (e.g., sclerenchyma) that yield clues to the reproductive morphology and
biology of ancient plants.
The differential dispersal or plant organs before they enter the fossil record creates
an interesting problem for the paleobotanist: We have to reassemble our plants
before we can think of them as unified organisms. Botanists studying living plants
take for granted that they know which leaves, wood, and reproductive structures
belong to a species. For the paleobotanist, many of our plants come as isolated
parts that have to be reunited. Obviously, the strongest evidence that organs
belong together is actual organic connection. For example, we find leaves and
reproductive structures both attached to a single branch. Another example is
isolating spores from sporangia that are preserved on their parent leaf. Somewhat
weaker inference is possible when two organs found in isolation have a unique,
derived feature. For example, particular glands on the epidermis of stem, petiole,
foliage and seeds allowed F.W. Oliver and D.H. Scott (1904) to unite these separate
organs into a single form: the Devonian seed plant Lyginopteris. The weakest
inference for uniting organs into a single plant species concept is when a suite of
organs always occur together. However, a variety of other processes could generate
this pattern and so caution is advised.

Because we don't always know which leaves belong to which seeds when they are
first discovered, we use the convention of form taxa. When organs are found
isolated (not in organic connection), each type of leaf and seed is given it's own
binomial name (genus and species name according to the International Code of
Botanical Nomenclature), without making any assumption about what belongs to
what. To use the example discussed by Oliver and Scott (1904), leaves were
described as Lyginopteris (genus only for brevity), seeds as Lagenostoma, and
stems as Lyginodendron. The similarity of the first syllable gives a hint that the
describing paleobotanists (others besides Oliver and Scott) suspected some
relationship, but were unable to make a strong inference link. The last syllable of
each name gives a hint to the organ type: "dendron" = stem, "pteris" is often used
for frond-like foliage, "stoma" = seed. However, after Oliver and Scott's recognition
of the unique glands on Lagenostoma lomaxi and species in the other organ form
genera, they were able to make the whole-plant link with greater confidence. The
whole plant then takes the name of the organ first described, in this
caseLyginopteris. When you are writing, take care to make clear whether you are
talking about form taxa (organs) or whole-plants.

Form taxa are not unique to paleobotany. For example, the planktonic larvae of
many marine invertebrates are commonly described as separate species when they
are first discovered in the ocean. Only later when they can be reared in the
laboratory can the link to their adult form be recognized. Similarly, the different life
stages of many fungi are given different names because they have different
physical forms and hosts. Only through detailed inoculation studies can mycologists
work out which forms are members of the same life cycle. Since some fungi may
have more than five discrete life cycle stages, this can be a long process. Similar
problems exist for some marine algae and multiple-host parasitic organisms of
many kinds. Even among well-studied vertebrates, some tropical birds have been
described as separate species until they are observed to mate and rear young
together.

Organism Level

Not all plants in a given community are equally likely to find their way into the fossil
record. Processes of fossilization often favor large or woody plants with resistant
tissues over small herbs. Likewise, wind dispersed pollen is much more common in
the fossil record than pollen dispersed by animals. Also, plants growing in or near
an area of preservation (e.g., riverbank or swamp) are more commonly preserved
than their counterparts growing far from water or anoxic sedimentary
environments.
Environmental Level

Plant preservation depends on removing the organic material from the zone of
aerobic decomposition. This is most easily accomplished by burying the plant.
Consequently, swamps, deltas, lakes, lowland flood plains, and volcanic areas are
good spots for fossilization (VG 1:9)(VG 1:12)(VG 1:13). Arid regions and
mountainsides are not likely candidates for plant fossil preservation (with the
exception of exquisitely preserved plant material from Pleistocene packrat middens
in the southwestern United States.

All of these taphonomic factors influence the information that can be recovered
from the plant fossil record. While taphonomic filtering does not preclude biological
interpretation of fossil material, taphonomy can introduce substantial biases into
the record and influence our interpretation of the fossils, and thus our
reconstruction of the ancient plants. It is, therefore, important always to keep in
mind the mode of preservation of fossils when making any interpretations from
fossils.

Conditions Required for Plant Fossil


Preservation
Three conditions are required for the preservation of plant fossils:

1. Removing the material from oxygen-rich environment of


aerobic decay;
2. "Fixing" the organic material to retard anaerobic decay;
3. Introducing the fossil to the sedimentary rock record
(a.k.a., burial).

Consequently, plant fossils are generally preserved in environments


very low in oxygen (e.g., anaerobic sediment) because most
decomposers (e.g., fungi, most decomposing bacteria and
invertebrates) require oxygen for metabolism. Such sediments are
commonly gray, green or black rather than red, a sedimentary signal
of oxygen-rich conditions. The "fixing" requirements means that plant
material must fall into an environment rich in humic acids or clay
minerals, which can retard decay by blocking the chemical sites onto
which decomposers fasten their degrading enzymes. Plant material
can also be "fixed" by removing degradable organic compounds
during the process of charring by wildfire. This is a common and
spectacular mode of preservation for flowers. Plant material can then
be incorporated into the rock record in areas where sediment is being
deposited, which usually, but not always, requires the presence of
water. Consequently, streams, flood plains, lakes, swamps, and the
ocean are good candidates for fossil-forming systems. Plant fossils
are commonly preserved in fine-grained sediment such as sand, silt,
or clay, or in association with organic deposits such as peat (coal).

As you look at the various modes of preservation in lab, note the


characteristics of the rock matrix in which plant material is preserved.
Note color, grain size (i.e., sand, silt, clay), mineral composition
(quartz, clay, mica, organic-rich, organic-poor), and any other
unusual features. If you aren't familiar with the basic features of
sedimentary rocks, they may start out all looking the sameÉdon't
worry. Take some extra time to make systematic observations of each
specimen: What color is it? Can you see the sediment grains with
your naked eye? With a hand lens? How would you classify them
(round or angular)? Is the rock shiny or dull? Does it have uniform or
mixed composition? Do you notice sedimentary layers or other
features (ripple marks, animal tracks)? Just be patient; you'll train
your eye to recognize rock types in no time.

Types of Plant Fossils


Six broad categories of plant fossils are commonly recognized.
Although these categories seem well-defined, a given fossil may fall
into several categories or may elude them all. Consequently, these
categories should be thought of as broad modes of preservation
rather than shoe boxes into which all fossils must go. When thinking
about types of fossils and modes of preservation, it is more important
to consider what types of biologically interesting information is or is
not present than to fret over strict classifications. With that caveat,
the basic types of plant fossils include:

o Compressions -- 2-dimensional, with organic material (VG


1:1)(VG 1:2) .
o Impressions -- imprints, 2-dimensional, devoid of organic
matter (VG 1:1)(VG 1:3) .
o Casts and Molds -- 3-dimensional, may have a surface layer of
organic material (VG 1:4)(VG 1:5).
o Permineralization -- 3-dimensional, tissue infiltrated by minerals
allowing internal preservation (VG 1:6)(VG 1:7) .
o Compactions -- 3 dimensional, reduced volume, flattened,
wholly organic (VG 1:8) .
o Molecular Fossils -- non-structural, preserves organic
compounds.

Each type of plant fossil carries different types of anatomical and


biological information. Consequently, to piece together the most
complete picture of an ancient plant, paleobotanists hope for the
same type of plant or plant part will be preserved in several different
styles.
Compressions

Compressions are plant parts that have suffered physical deformation


such that the three-dimensional plant part is compressed to more-or-
less two-dimensions. Compressions retain organic matter, usually
more or less coalified. Compressions of leaves (VG 1:1) , for
example, differ from impressions in that some organic substance,
often cuticle, is preserved. Peat, lignite, and coal are essentially
compressions of thick accumulations of plant debris relatively free of
encasing mineral sediment.

Compressions are excellent records of external form, especially for


planar structures like leaves. They often preserve cuticle that can be
recovered by dissolving the mineral matter in hydrofluoric acid (HF)
or disaggregating in mild peroxide (VG 1:2). The cuticle retains the
imprint of epidermal cells, but other than this, cellular information
can seldom be recovered from compressions. Consequently,
compressions generally preserve plants at the organ, organism,
and/or environment level. In addition, because compressions
preserve organic material, carbon isotopic studies can be performed
on compressions. From these studies, paleobotanists can sometimes
recognize the biochemical signature of C3, C4, and CAM
photosynthetic physiology in extinct species.

Study the assortment of compressions available. Note color, texture,


and type of deformation (usually flattening) experienced by each
specimen. Notice distortion in original morphology is introduced by
flattening.

Examine samples of coal, noting their weight, texture and surface


features (dull vs. shiny). Identify discrete plant parts in the different
ranks (essentially grades of metamorphism) of coal. Coal geologists
recognize a continuum of degree of metamorphism in coal: peat -
lignite - bituminous coal (A-C) - anthracite. Peat is an accumulation of
virtually unaltered plant material, while anthracite is nearly pure
carbon with little trace of the original plant material. All materials can
be burned for fuel, but the energy content per weight increases with
degree of metamorphism and the proportion of impurities generally
decreases. Study thin sections of coal at the microscope. Can you
identify specific plant parts in the thin sections? Study samples of
fusain (fossil charcoal).

Impressions

Impressions are two-dimensional imprints of plants or their parts


found, most commonly, in fine-grained sediment such as silt or clay.
Impressions are essentially compressions sans organic material. If
the sediment is very fine-grained, impressions may faithfully replicate
remarkable details of original external form, regardless of subsequent
consolidation of the sediment. Study the specimens of impressions,
noting the several categories of plant parts represented. Because of
their shape, texture, and abundance, leaves are among the most
common organ preserved in impression (VG 1:1). Impressions may
also occur if, when layers of rock are split apart, the organic material
adheres to only one side of the rock. In this case, the side with
organic material is the compression, known as the "part", while the
corresponding impression known as the "counterpart".

One particularly interesting type of impression forms in "dirty" sand


(VG 1:3) . In this type of sediment, relatively coarse sand grains are
mixed with silt and clay. This type of sediment is common in river and
flood plain environments so is important for terrestrial plant
preservation. When a leaf falls into this type of sediment and begins
to decay, the first organic bonds to break leave charged molecular
tails hanging off the leaf surface. This charged tail attracts clay
particles with opposite charge that linger within the sediment. The
clay migrates to the leaf surface, coating the organic structure. This
has two remarkable consequences: First, further decay is retarded
because clay is occupying sites of organic reaction. Second, the fine
clay allows remarkable detail to be preserved. Because most of the
sediment is relatively coarse (sand), the organic material is lost later,
but an exquisitely detailed impression is retained in the clay film. This
mode of preservation is important in the Dakota Sandstone flora of
Cretaceous age. It is also important in the preservation of remarkable
animal fossils such as the Jurassic bird Archaeopteryx and the strange
Cambrian invertebrates of the Burgess Shale.

Impressions, like compressions, record information about external


shape and morphology of plant organs. However, because they lack
organic material, cuticle and organic carbon cannot be recovered
from them. In cases of impressions in very fine-grained sediment,
some cellular detail can be recovered by making a latex of silicone
rubber cast of the impression.

Casts and Molds

When sediment is deposited into cavities left by the decay of plant


parts, a cast results (VG 1:4). A mold is essentially a cavity left in
the sediment by the decayed plant tissue. Molds are generally
unfilled, or may be partially filled with sediment. Casts and molds
commonly lack organic matter, but a resistant structure like periderm
may be preserved as a compression on the outside of the cast or the
inside of a mold. Casts and molds may be found together with the
cast filling the mold.
Molds are formed when soft sediment surrounding the structure
lithifies or hardens before the structure decays. When the mold fills in
with sediment that subsequently hardens, a cast is formed. Casts of
an internal hollow structure like a pith cavity are also common. Pith
casts can be confusing because you are looking at the inside of the
fossil-what in life would have been empty space. Like compressions
and impressions, casts and molds record external (or sometimes
internal) organ features well, but provide no cellular or tissue
information. Unlike compressions/impressions, molds and casts often
are truer records of the original three-dimensional shape of the
structure. Casts of ancient trees are among the most impressive plant
fossils (VG 1:5).

Permineralizations

Permineralization occurs when the plant tissues are infiltrated with


mineral-rich fluid. Minerals (commonly silica, carbonate, phosphate or
pyrite or rarely other minerals) precipitate in cell lumens and
intercellular spaces, thus preserving internal structures of plant parts
in three dimensions. This type of preservation is known as "structural
preservation". Because organic material (commonly cell walls but in
some cases finer detail) is preserved, permineralizations can yield
detailed information about the internal structure of the once-living
plant (VG 1:6). When mineral matter actually replaces the cell-wall
and other internal structures, the preservation may be called
petrifaction. In petrified specimens, cellular details are lost with the
organic material of the cell wall. Please note that we are using these
words in a precise, scientific sense. "Petrified" also has a colloquial
meaning that might encompass what we distinguish as
"permineralized". Therefore, for the purpose of this course,
permineralized wood preserves the cellular detail of wood anatomy
and the lignin of cell walls that has been "fixed" by a mineral in-filling.
This is much like bioplastic in-filling cellular structures when one
makes a histological thin section. Petrified wood, on the other hand,
lacks such cellular preservation.

Silica permineralization (silification (VG 1:7)) commonly occurs in


areas where silica-rich volcaniclastic sediments are weathering, for
example the famous upright trees in Yellowstone National Park or
nearby Calistoga, California. Silification is also an important
preservational mode for Precambrian microbial remains deposited in
near-shore marine environments.

Permineralization with calcium carbonate (calcite or dolomite) is


particularly common in Carboniferous coal seams (VG 1:9), where
whole regions of peat were permineralized. Called coal balls (because
of their sometimes round or ellipsoidal shape (VG 1:10)) or widow
makers (because of their tendency to drop out of mine roofs onto the
heads of unsuspecting miners), these fossils commonly preserve a
hodge-podge of plants and plant organs.

Permineralizations in pyrite (an iron-sulfur mineral) are particularly


important in Devonian rocks where coal balls and well-preserved
compactions are rare or unknown. These pyritized fossils often occur
in the presence of sea water (a source of sulfur), and are
characteristic of plant tissues washed into marine basins. Pyrite
permineralizations offer a challenge to the museum curator because
iron in pyrite exists in a reduced state and tends to oxidize when
exposed to air. Upon oxidization, most of the structures are lost. This
is called "pyrite disease" in fossils and is characterized by a mold-like
appearance on the cut surface of the coal ball. To prevent destruction,
the surface can be coated with a sealant. Coal balls can also be
stored in an low-oxygen medium like glycerin or antifreeze.

Permineralization with phosphate is uncommon for land plants, but


can be important in some types of marine settings.

Compactions

In peat, brown coals (lignite), middens and soft sediments, plant


remains may retain their external form with only slight volume
reduction due to compaction. Such tissues are not mineralized, retain
resistant organic material, and may show unidirectional compression
(flattening). Internal structure, especially of thick-walled, hard fruits
is sometimes well preserved. These fossils may be sectioned by
microtome or embedded and treated much like living tissues.
Compactions are most common in the youngest plant fossils. Examine
specimens of Tertiary fruits and seeds, and thin sections made from
them. Pollen and spores are also preserved as compactions (VG
1:8). The material making up their outer shells (sporopollenin) is
extremely resistant to decay and can remain for hundreds of millions
of year practically unaltered in the rock record. However, the pollen
and spore shells, once spherical, are flattened by the compressive
forces of lithification.

Molecular Fossils

As more becomes known about the chemistry of modern plants,


paleobotanists have begun to examine the fossil record for
corresponding chemical data. For example, characteristic breakdown
products of chlorophylls and lignins have been found in well-
preserved fossil leaves. Lipids and their derivatives have also been
recovered from sediments. Some carbohydrate break-down products
may also survive in sediment. A special class of these, oleananes, are
formed by flowering plants, some ferns and lichens. An increase in
abundance of these molecules in sediments of mid to Late Cretaceous
age is used to document the increasing abundance of flowering plants
(Moldowan et al, 1994). In another stunning example, genetic
material was recovered from Tertiary leaves, and the age of material
from which DNA and RNA is recovered seems to be greater with every
issue of Nature. As testament to the anoxic requirement for
preservation of most molecular fossils, the RNA recovered from fossil
leaves degrades within a few seconds when exposed to air, so special
preparation techniques were developed to harvest and transport the
material to the lab for processing.

Molecular fossils are recovered and studied using chromatographic


techniques, mass spectrometry, and spectrophotometry. The
preservation of these chemical products is highly variable, and
depends on oxygen levels during deposition, temperatures
experienced by the rocks since preservation, and many other physical
and chemical factors. In a similar vein, geochemists have investigated
the chemistry of petroleum and its precursors in an attempt to
understand its formation.

Fossil DNA and RNA have also been making headlines in the scientific
press. In some exceptional cases, genetic material or proteins have
been sufficiently well-preserved to permit their use in the
reconstruction of evolutionary relationships, in much the same way as
one might sequence living organisms. However, much of this work is
controversial due to the difficulty of preserving and isolating these
fragile molecules. Also, contamination by other materials is a
common and difficult to recognize problem.

A more mainstream application of organic chemistry to the study of


ancient plants is that of stable carbon isotopes. During
photosynthesis, plants reduce carbon from carbon dioxide to form
organic molecules. This ratio of carbon-12 to carbon-13 in the
resulting compounds gives information on the proportion of these
isotopes in the atmosphere (interesting for geological questions
relating to global carbon cycling) and about the physiology of the
plant itself. In lecture we will discuss several applications of these
techniques.

Read More About It!

• Gastaldo, Savrda, & Lewis. 1996: Deciphering Earth


History: Taphonomy -- This WWW site provides more info
on preservation and taphonomy which might be helpful.
• Taphonomy Virginia Tech -- This WWW site might also be
of use.
Anderson, T.F., M.E. Brownlees, and T.L. Phillips. 1981. A stable
isotope study of the origin of permineralized peat zones in the Herrin
coal. Journal of Geology 88:713-722.

Cohen, A.D. and W. Spackman. 1980. Phytogenetic organic sediments


and sedimentary environments in the Everglades -- Mangrove
complex. Part III. The alteration of plant material in peat and the
origin of coal macerals. Palaeontographica 172B:125-149.

Ferguson, D.K. 1985. The origin of leaf assemblages--new light on an


old problem. Review of Paleobotany and Palynology 48:117-188.

Gastaldo, R. 1986. Selected aspects of plant taphonomic processes in


coastal deltaic regions. University of Tennessee Studies in Geology
15:27-44.

Karowe, A. and T. Jefferson. 1987. Burial of trees by eruptions of Mt.


St. Helens, Washington: implications for the interpretation of fossil
forests. Geology Magazine 124:191-204.

Knoll, A.H. 1985. Exceptional preservation of photosynthetic


organisms in silicified carbonates and silicified peats. Philosophical
Transaction of the Royal Society of London 311B:111-122.

Moldowan, J.M., J. Dahl, B.J. Huizinga, F.J. Fago, L.J. Hickey, T.M.
Peakman, and D.W. Taylor. 1994. The molecular fossil record of
loeanane and its relation to angiosperms. Science 265:768-771.

Oliver, F.W. and D.H. Scott. 1904. On the structure of the Palaeozoic
seed Lagenostoma lomaxi, wit a statment of the evidence upon which
it is referred to Lyginodendron. Philosophical Transactions of the
Royal Society of London 197B: 193-248.

Schopf, J.M. 1975. Modes of fossil preservation. Review of


Paleobotany and Palynology 20:27-53.

Scott, A.L. and G. Rex. 1985. The formation and significance of


Carboniferous coal balls. Philosophical Transaction of the Royal
Society of London 311B:123-137.

Spicer, R.A. and A. Greer. 1986. Plant taphonomy in fluvial and


lacustrine systems. University of Tennessee Studies in Geology
15:10-26.

TAPHONOMY & PRESERVATION
INTRODUCTION
As fossils are the preserved remains of ancient organisms or their traces, 
understanding the process of preservation, and more importantly, being able to 
recognize and identify fossil remains after their discovery is an integral part of 
paleobiology. Protective cover (sediments) and stabilizing chemical environments 
are of prime importance in the preservation of once living organisms. Due to the 
process of aerobic decay and physical/chemical destruction, most animals leave no 
evidence of their existence.

In order to make a correct interpretation of taphonomic processes and mode of 
preservation, it is often necessary to have a prior knowledge of the structural 
features or morphology of original skeleton in addition to knowing its original 
mineralogical composition. This limitation should diminish as you become 
familiar with the various fossil groups throughout the semester.

TAPHONOMY
Taphonomy is the study of what happens to an organism after its death and until 
its discovery as a fossil. This includes decomposition, post­mortem transport, 
burial, compaction, and other chemical, biologic, or physical activity which affects 
the remains of the organism. Being able to recognize taphonomic processes that 
have taken place can often lead to a better understanding of paleoenvironments 
and even life­history of the once­living organism.

In addition, understanding which taphonomic processes a fossil occurrence has 
undergone, and to what degree, may have implication on interpreting the 
significance of the fossil deposit and clearer understanding of the biases in the 
sample.

An outline of the pathways affecting the preservation of once living organisms can 
be found in Figure 1 below. As discussed below, this encompasses both the 
processes of biostratinomy and diagenesis. 

Figure 1 - The field of Taphonomy as it relates to steps in 
transformation from living organisms to fossils.
Modified from McRoberts (1998)

Processes that occur between the death of an organism and its subsequent burial in 
the sediment are termed biostratinomy. Generally, this includes the 
decomposition and scavenging of the animal's soft parts, and at least some amount 
of post­mortem transport. Such things as the amount of shell breakage and the 
concentration of shells in layers often indicate the level of water energy and post­
mortem transport. For example, the shell­hash or coquina  has experienced a 
significant amount of shell breakage and probably post­mortem transport 
suggesting deposition in high energy environments; whereas, the articulated plant 
remains  are intact suggesting little or no post­mortem transport and deposition 
in a very low energy and oxygen­free environment. In Table 1 below are various 
taphonomic indicators and their environmental implications.

The physical and/or chemical effects after burial are called diagenesis. This is the 
realm in which dissolution, replacement, or recrystallization of original shell 
material occurs, as can the formation of molds and casts. A more detailed 
description of diagenesis with regards to fossil preservation in the next section.

Table 1

Summary of Taphonomic Indicators and Their Paleoenvironmental 
Implications

TAPHONOMIC 
IMPLICATIONS
FEATURE

The wearing­down of skeletons owing to differential 
movement with respect to sediments is an indicator of 
Abrasion environmental energy. Significant abrasion is most commonly 
found on skeletal material collected from beaches, or areas of 
strong currents or wave action.
Multi­element skeletons are soon disarticulated after death. 
Articulated skeletons, then, indicate rapid burial or otherwise 
Articulation
removing the skeleton from the effects of energy of the 
original environment.

Bioerosion encompasses the many different corrosive 
processes by organisms. The most pervasive causes of 
degradation are boring and grazing. Bioerosion erases 
information from the fossil record, but it also leaves 
Bioerosion
identifiable traces made by organisms on remaining hard 
skeletons or surfaces. Therefore, trace fossils produced by 
bioerosion add information on the diversity of ancient 
assemblages.

Skeletal remains commonly are in equilibrium with 
surrounding waters, but changes in chemical conditions can 
cause skeletons to dissolve. Dissolution represents fluctuation 
Dissolution
in temperature, pH or pCO2 in calcium carbonate skeletons. 
Siliceous skeletons also can dissolve because normal sea water 
is usually undersaturated with respect to silica.

Broken edges of skeletons become rounded owing to 
dissolution and/or abrasion of exposed surfaces. Processes 
Rounding that control edge rounding probably include a combination of 
dissolution, abrasion, and bioerosion. Rounding gives an 
estimate of time since breakage.

The growth of hard skeleton substrates by other organisms is a 
common occurrence. Besides indicating exposure of the 
skeleton above the sediment­water interface, encrustation can 
Encrustation
specify a particular environment. Different patterns of 
encrustation, as well as different biota, occur in different 
environments.

Breakage of skeletons is usually an indication of high energy 
resulting from wave action or current energy. Fragmentation 
Fragmentation
also can be caused by other organisms through either 
predation or scavenging.

Orientation After death, skeletal remains are moved by the transporting 
medium and oriented relative to their hydrodynamic 
properties. Fossil skeletons in life position indicate rapid 
burial, attachment to a firm substrate, or death of in­place 
infauna. Hard parts tend to orient long­axis parallel to 
unidirectional flow in current­dominated areas and 
perpendicular to wave crests on wave­dominated bottoms.

After death and if not rapidly buried, a skeleton behaves as a 
sedimentary particle and is moved and sorted with respect to 
Size the carrying capacity of the flow of currents, waves, or tides. 
Size can, therefore, be an effective indicator of flow capacity in 
a hydraulic or wind­driven system.

From McRoberts (1998)

FORMS OF PRESERVATION

UNALTERED

This form of preservation is rare in most of the geologic column, but becomes more 
frequent in younger sedimentary rocks. Types of unaltered preservation where 
even the soft body parts are preserved include: (i) mummification, (ii) encasement 
in tar, (iii) encasement in amber, (iv) encasement in sediment, and (v) freezing. 
More frequently, however, only the hard skeletal material is preserved after 
removal of soft body parts.

Examples of unaltered preservation include the skeleton of a horseshoe crab, 
whose shell is composed of interlocking plates and jointed appendages which 
quickly disarticulate after death; cockle bivalved molluscs,  whose outer­most 
shell layer has been removed by abrasion, yet the original shell material of the 
inner layers remains; an ammonoid  from the Cretaceous period in which you 
should note the pearly luster which is original aragonite shell material; and an 
insect encased in amber  .

MOLDS & CASTS

This general class of preservation entails making "replicas" of the skeletal hard 
parts of organisms. In general, a mold is an impression in the sediment of a 
skeleton or shell. Once encased in lithified sediment, the dissolution of skeletal 
material leaves behind the impression or mold of original skeletal form. Thus, a 
mold is a "mirror image" of the original skeleton. An internal mold (sometimes 
called a steinkern) is the impression of the inside surface of skeletal hard parts. An 
external mold is the impression of the outside surface of skeleton or bone. An 
example of both types of molds can be seen in this image of a trilobite  . 

A cast is formed by the filling­in of a mold. It is thus a true replica (not a "mirror 
image") of the original skeleton or shell. By this definition, the cast one gets for a 
broken limb is not really a cast at all but an external mold.

A graphical representation of the formation of casts and molds is provided in 
Figure 2 below.

Figure 2 ­ Different diagenetic processes leading to different 
preservational styles in skeletal materials.

* Note that molds are produced directly as imprints of the shell and casts are produced from 
molds.

Modified from McRoberts (1998)

REPLACEMENT & RECRYSTALLIZATION

This common form of preservation involves chemical and/or physical alteration or 
replacement of original skeletal material. To properly identify replacement and 
recrystallization, one must know what the original constituents of the organism's 
skeleton were. These are provided in Figure 1.3. Replacement occurs often by the 
filling in (by various minerals) of the void space after dissolution of original 
skeletal material. Sometimes, the replacement occurs on a molecule by molecule 
basis. Common replacement minerals that you should be able to recognize include 
Silica (SiO2) as shown in the coral,  and Pyrite (FeS2) shown in the ammonoid. 

Recrystallization involves the physical re­arrangement of crystalline structure of 
skeletal material. This is a common phenomenon in shells which were originally 
aragonite and/or calcite (both forms of calcium carbonate­ CaCO3). Examples, 
both of which are now calcite, include a gastropod which was originally aragonite 
and a brachiopod which was originally calcite. 

CARBONIZATION

As organic remains decompose in sediments, volatile constituents such as oxygen, 
hydrogen, and nitrogen are slowly lost to the surrounding sediments frequently 
leaving behind a carbon film. This process is carbonization (or sometimes called 
distillation), and occurs most frequently in oxygen deficient, organic­rich 
environments such as basinal black shales, and coal swamps. The carbon films 
often show exquisite details of plants and soft­body parts of animals not readily 
preserved, and can often be recognized by a dark gray or black film with a metallic 
sheen such as these fern­like fossil plants. 

PERMINERALIZATION

Permineralization involves the filling­in of pore and/or void spaces in shell or 
bone by secondary mineral matter in solution. With permineralization, the tiny 
pore spaces in the fossil are filled and the original skeletal material is still retained. 
However, it is often common for other types of preservation (e.g. replacement) to 
occur during and/or after permineralization. Because of its porous nature, bone 
and wood  is especially prone to permineralization.

You might also like