You are on page 1of 30

geosciences

Article
Non-Mineralized Fossil Wood
George E. Mustoe ID

Geology Department, Western Washington University, Bellingham WA 98225, USA; mustoeg@wwu.edu;


Tel.: +1-360-650-3582

Received: 25 May 2018; Accepted: 11 June 2018; Published: 18 June 2018 

Abstract: Under conditions where buried wood is protected from microbial degradation and exposure
to oxygen or harsh chemical environments, the tissues may remain unmineralized. If the original
organic matter is present in relatively unaltered form, wood is considered to be mummified. Exposure
to high temperatures, whether from wild fires or pyroclastic flows, may cause wood to be converted
to charcoal. Coalification occurs when plant matter undergoes gradual metamorphosis, producing
bituminous alteration products. Examples of all three types of non-mineralized wood are common
in the geologic record. This report describes some of the most notable occurrences, reviews past
research and introduces data from several localities in North America.

Keywords: fossil wood; carbonized; charcoal; coalified; mummified; paleobotany

1. Introduction
From a deep time perspective, forests are a transient phenomenon, their existence depending on
environmental factors that include elevation change, atmospheric physics, ocean currents patterns, and
tectonic plate motion. These factors determine whether a region is an arid desert or a lush rainforest.
Foresters describe forests that contain trees with ages of only a few hundred years as “old growth”.
The oldest known age record for an individual tree is a Bristlecone Pine (Pinus longaeva) growing
in White Mountains of California USA; a tree core taken in 2012 was observed to have 5067 annual
rings [1]. The ages of clonal tree clusters, where individual trunks share a common root network,
may be much greater, e.g. the 80,000 year age estimated for ~47,000 component trees of a Quaking
Aspen (Populus tremuloides) grove in central Utah, USA [2]. The ages of these living trees are brief in
comparison to fossil evidence of trees that comprised ancient forests. The infiltration and replacement
of inorganic minerals produces petrified wood that has sufficient durability to be preserved for many
millions of years. Recent reports [3,4] describe these mineralization processes in detail, and list earlier
literature. Less well-known are fossil forests where ancient trees have been preserved as original tissue,
or as organic material that has undergone structural alteration.
This paper provides a detailed review of fossil wood occurrences where tissues have not been
mineralized. These fossils are sometimes described as “carbonized wood”, but unmineralized ancient
wood can be divided into three categories. Mummified wood consists of original tissues that may
have undergone anatomical distortion or desiccation, but which are otherwise free of alteration.
Charcoalified wood originates when wood is combusted in an anaerobic environment, causing much
of the organic materials to be reduced to pure carbon. Coalified wood consists of tissues that have
been altered by heat and pressure during deep burial, converting the original organic constituents to a
mixture of pure carbon and various hydrocarbons.

2. Analytical Methods
Scanning electron microscope (SEM) images were taken by the author using a Tescan Vega SEM
at Western Washington University, Bellingham, WA USA.

Geosciences 2018, 8, 223; doi:10.3390/geosciences8060223 www.mdpi.com/journal/geosciences


Geosciences
Geosciences 2018, 2018,
8, 2238, x FOR PEER REVIEW 2 of 322 of 30

Scanning electron microscope (SEM) images were taken by the author using a Tescan Vega SEM
at Western Wood
3. Mummified Washington University, Bellingham, WA USA.

A common question asked of fossil wood researchers is “how long does it take for wood to become
3. Mummified Wood
petrified?” There is no uniform answer, because fossilization processes are highly variable. In siliceous
A common question asked of fossil wood researchers is “how long does it take for wood to
hot springs, wood may become rapidly impregnated with silica [3,5–7], but in other environments,
become petrified?” There is no uniform answer, because fossilization processes are highly variable.
buried wood may be preserved for tens of millions of years without experiencing mineralization.
In siliceous hot springs, wood may become rapidly impregnated with silica [3,5–7], but in other
The most important conditions required for mummification are an absence of microbial and chemical
environments, buried wood may be preserved for tens of millions of years without experiencing
degradation, and burial
mineralization. conditions
The most where
important mineral-bearing
conditions required for groundwater
mummification is unable
are antoabsence
penetrate
of the
microbial and chemical degradation, and burial conditions where mineral-bearing groundwater is of
tissues. Mummified wood consists of original tissues that have undergone minimal degradation
unable
cellular to penetrateThis
constituents. the preservation
tissues. Mummified woodof
is a result consists
severalof factors:
original inhibition
tissues that of have undergone
wood-destroying
minimal
microbes, degradation
decreased oxygen of availability,
cellular constituents. This preservation
and the absence is a result
of harsh chemical of several
or physical factors: (e.g.,
conditions
high inhibition
alkalinity of or wood-destroying
acidity, or elevated microbes, decreasedEnvironments
temperatures). oxygen availability, and the
that favor absence of harsh
mummification include
chemical or physical conditions (e.g., high alkalinity or acidity, or
deeply submerged wood, burial in impermeable sediments, aridity, or low temperatures [8–12]. Ligninelevated temperatures).
Environments that favor mummification include deeply submerged wood, burial in impermeable
and cellulose, the primary constituents of wood, may undergo alteration on a microscopic level, even
sediments, aridity, or low temperatures [8–12]. Lignin and cellulose, the primary constituents of
though the overall appearance of the ancient wood may appear unchanged. Wood cell walls consist of
wood, may undergo alteration on a microscopic level, even though the overall appearance of the
an outer primary wall that encloses a multilayered secondary wall. The primary wall, which may be
ancient wood may appear unchanged. Wood cell walls consist of an outer primary wall that encloses
relatively thick, consists
a multilayered mostly
secondary wall.ofThe
cellulose
primaryandwall,hemicellulose,
which may be in contrast
relatively to the
thick, thinner
consists secondary
mostly of
cellulose and hemicellulose, in contrast to the thinner secondary wall that is largely composed of a
wall that is largely composed of a mixture of cellulose and lignin. This secondary wall commonly
mixture
consists of cellulose
of three discreteand lignin.
layers This secondary
(lamellae) that differwallincommonly consists
the orientation of of three
their discrete aggregates
cellulosic layers
(lamellae)
and the degreethat differ in the orientation
of lignification. Microbialof their cellulosic
degradation aggregates
of wood may and thein
result degree of lignification.
destruction of the wood
sugarsMicrobial
cellulose degradation of wood may
and hemicellulose, resultininrelative
resulting destruction of the of
enrichment woodligninsugars cellulose
[13,14]. and
Deterioration
hemicellulose, resulting in relative enrichment of lignin [13,14]. Deterioration
of the lignin-rich secondary wall may not greatly disrupt the overall cellular architecture as long of the lignin-rich
as thesecondary
cellulosic wall may not greatly disrupt the overall cellular architecture as long as the cellulosic outer
outer walls remain relatively intact (Figure 1). When it is preserved, cellulose has
walls remain relatively intact (Figure 1). When it is preserved, cellulose has scientific value because
scientific value because carbon and oxygen isotope ratios in this substance can be used to estimate
carbon and oxygen isotope ratios in this substance can be used to estimate paleoprecipitation and
paleoprecipitation and paleotemperature [15–18].
paleotemperature [15–18].

Figure 1. Anatomical preservation in mummified food from western Washington, USA. (A,B)
Figure 1. Anatomical preservation in mummified food from western Washington, USA.
Transverse views, late Pleistocene wood, Whidbey Formation, Swantown, Whidbey Island, Island
(A,B) Transverse views, late Pleistocene wood, Whidbey Formation, Swantown, Whidbey Island,
Island County, Washington State, USA. (C) Transverse view, Wilkes Formation compressed wood.
(D) Late Pleistocene wood showing extensive decay. Radial view. Net-like structures are relict ray cell
walls, Whidbey Formation, Cama Beach, Camano Island, Island County, Washington, USA.
Geosciences 2018, 8, x FOR PEER REVIEW 3 of 32

County, Washington State, USA. (C) Transverse view, Wilkes Formation compressed wood. (D) Late
Pleistocene wood showing extensive decay. Radial view. Net-like structures are relict ray cell walls,
Whidbey Formation, Cama Beach, Camano Island, Island County, Washington, USA.
Geosciences 2018, 8, 223 3 of 30

Reported occurrences of mummified wood range in age from Paleocene to Holocene. This
subfossil wood occurrences
Reported is often locally
of abundant,
mummified particularly
wood range in Quaternary
in age fromdeposits,
Paleocenewhere glacial, fluvial,
to Holocene. This
and volcaniclastic
subfossil laharlocally
wood is often deposits provide
abundant, deep layers
particularly of fine sediment
in Quaternary deposits,that
wherepreserve
glacial,organic
fluvial,
materials. The following
and volcaniclastic section
lahar deposits describes
provide deepsome
layersof the sediment
of fine best-known
thatlocalities, but it materials.
preserve organic is not a
comprehensive list. describes some of the best-known localities, but it is not a comprehensive list.
The following section

3.1.
3.1. Canadian
Canadian Arctic
Arctic Localities
Localities

Spectacular
Spectacular examples
examples of
of mummified
mummified wood wood occur
occur in
in the
the Canadian
Canadian Arctic,
Arctic, ranging
ranging in
in age
age from
from
Paleocene
Paleoceneto toPliocene
Pliocene(Figure
(Figure2).2).Although
Although thethe
modern
modernHigh Arctic
High is a polar
Arctic desert
is a polar typified
desert by long
typified by
cold
long winters and short
cold winters and cool
shortsummers,
cool summers, early Tertiary conditions
early Tertiary were very
conditions weredifferent, with a with
very different, warma
ice-free environment
warm ice-free where where
environment plants plants
benefited from high
benefited fromhumidity and aand
high humidity growing season
a growing that had
season that
three months of continuous light. Three months of continuous darkness during Arctic Circle winters
had three months of continuous light. Three months of continuous darkness during Arctic Circle
favored conifersconifers
winters favored rather rather
than angiosperms.
than angiosperms. TheseThese
fossilfossil
forests arearetypically
forests typicallydominated
dominated by by
Metasequoia,
Metasequoia, an
an unusual
unusual deciduous
deciduous conifer
conifer particularly
particularly suited
suited for
for long
long dark
dark winters.
winters. Many
Many other
other taxa
taxa
are
are represented,
represented, particularly
particularly in
in the
the Miocene
Miocene and
and Pliocene
Pliocene occurrences [19,20].
occurrences [19,20].

Figure
Figure 2.
2. Fossil forests
Fossil in in
forests thethe
Canadian Arctic.
Canadian (A) Axel
Arctic. Heiberg
(A) Axel IslandIsland
Heiberg (Eocene); (B) Banks
(Eocene); (B) Island
Banks
(Miocene/Pliocene); (C) Cornwallis
Island (Miocene/Pliocene); Island
(C) Cornwallis (Miocene);
Island (E) (E)
(Miocene); Ellesmere
EllesmereIsland
Island (Paleocene/Eocene,
(Paleocene/Eocene,
Pliocene);
Pliocene); (M)
(M) Meighen
Meighen Island
Island (Miocene).
(Miocene).
Geosciences 2018, 8, 223 4 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 4 of 32

3.2.
3.2. Ellesmere
Ellesmere Island—Late
Island—Late Paleocene/early
Paleocene/earlyEocene,
Eocene,Pliocene
Pliocene
Fossil
Fossil wood
wood was
was discovered
discovered at at northern
northern Ellesmere
Ellesmere Island
Island inin 1883
1883 at
at aa site
site later
later described
described as as
“Brainard’s
“Brainard’s Petrified Forest”, named after a Sargent D.L. Brainard, a survivor of the ill-fated Greely
Petrified Forest”, named after a Sargent D.L. Brainard, a survivor of the ill-fated Greely
Expedition
Expedition of of 1881–1883
1881–1883 [21].
[21]. Interest
Interest was
was revived
revived inin the
the 1980s
1980swhen
whenGeological
Geological Survey
Survey of ofCanada
Canada
geologist
geologist Neil McMillan observed fossil wood at Ellesmere Island. The main fossil forest areas
Neil McMillan observed fossil wood at Ellesmere Island. The main fossil forest areas atat
Strathcona
Strathcona and
and Stenkul
Stenkul Fjords
Fjordscontain
contain siderite-mineralized
siderite-mineralized wood wood preserved
preserved in in Late
Late Paleocene-Early
Paleocene-Early
Eocene
Eocene coal-bearing
coal-bearing sediments
sediments of of the
the Iceberg
Iceberg Bay
BayFormation
Formation[22–24].
[22–24]. Pliocene
Pliocene mummified
mummified wood wood
occurs
occurs at the Beaver Pond fossil site near the head of Strathcona Fjord, where braidedriver
at the Beaver Pond fossil site near the head of Strathcona Fjord, where braided riverdeposits
deposits
of
of the
the Beaufort
Beaufort Formation
Formation preserve
preserve woody
woody debris
debris [25].
[25]. Many
Many of of the
the twigs
twigs have
have teeth
teeth marks
marks left
left by
by
ancestral beaver (Figure 3). Pliocene wood also occurs at nearby Meighen Island,
ancestral beaver (Figure 3). Pliocene wood also occurs at nearby Meighen Island, including both including both
individual
individuallogs
logsand
andfine
finewoody
woodydebris
debris[26].
[26].

Figure 3. Pliocene wood from Canadian Arctic sites. (A) Natalia Rybcynski with log, Fyles Leaf Bed,
Figure 3. Pliocene wood from Canadian Arctic sites. (A) Natalia Rybcynski with log, Fyles Leaf Bed,
~3 km south of Beaver Pond locality. Ellesmere Island. (B) Beaver-gnawed twigs preserved in
~3 km south of Beaver Pond locality. Ellesmere Island. (B) Beaver-gnawed twigs preserved in Pliocene
Pliocene sediments at Beaver Pond site. (C) Tree trunk in Beaufort Fm. Strata at Meighen Island. (D)
sediments at Beaver Pond site. (C) Tree trunk in Beaufort Fm. Strata at Meighen Island. (D) Fine woody
Fine woody debris. Meighen Island. Photos courtesy of Natalia Rybcynski. (E) William Hagopian
debris. Meighen Island. Photos courtesy of Natalia Rybcynski. (E) William Hagopian with freshly
with freshly excavated trunk fragment, Stenkul Fjord, Ellesmere Island 2005. Photo by Ann Jefferson.
excavated trunk fragment, Stenkul Fjord, Ellesmere Island 2005. Photo by Ann Jefferson.

3.3. Axel Heiberg Island—Middle Eocene


3.3. Axel Heiberg Island—Middle Eocene
The fossil forest on this island was discovered in 1985 by helicopter pilot Paul Tudge, who had
The fossil forest on this island was discovered in 1985 by helicopter pilot Paul Tudge, who had
been transporting Canadian Geological Survey scientists to Arctic locations for years. Tudge notified
been transporting Canadian Geological Survey scientists to Arctic locations for years. Tudge notified
University of Saskatchewan paleobotanist James Basinger, who organized the first expeditions to
University of Saskatchewan paleobotanist James Basinger, who organized the first expeditions to visit
visit the site.
the site.
Located in the Geodetic Hills, the fossil forest is preserved in sediments of the middle Eocene
Located in the Geodetic Hills, the fossil forest is preserved in sediments of the middle Eocene
Buchanan Lake Formation. The fossils are preserved in a basin where fluvial sedimentation included
Buchanan Lake Formation. The fossils are preserved in a basin where fluvial sedimentation included
episodic floods that killed trees, burying stumps in situ together with fallen logs and abundant forest
episodic floods that killed trees, burying stumps in situ together with fallen logs and abundant
floor leaf litter. The total depth of sediment exceeds 150 m, and preserves at least 33 separate
forest floor leaf litter. The total depth of sediment exceeds 150 m, and preserves at least 33 separate
fossil-bearing layers [27]. As noted later in this report, although mummified wood is a dominant
fossil-bearing layers [27]. As noted later in this report, although mummified wood is a dominant
component or the Axel Heiberg Island fossil forest, some specimens have been coalified.
component or the Axel Heiberg Island fossil forest, some specimens have been coalified.
Metasequoia (Dawn Redwood) wood is by far the most abundant fossil (Figure 4). Other conifers
Metasequoia (Dawn Redwood) wood is by far the most abundant fossil (Figure 4). Other conifers
include fir, cypress, ginkgo, larch, redwood, spruce, pine and hemlock; Angiosperm trees include
include fir, cypress, ginkgo, larch, redwood, spruce, pine and hemlock; Angiosperm trees include
maple, alder, birch, hickory, chestnut, beech, ash, holly, walnut, sweetgum, sycamore, oak, willow,
maple, alder, birch, hickory, chestnut, beech, ash, holly, walnut, sweetgum, sycamore, oak, willow,
and elm. Herbaceous plants are represented by honeysuckle and sumac; other taxa include several
and elm. Herbaceous plants are represented by honeysuckle and sumac; other taxa include several
species of ferns and mosses [27]. This botanical diversity is evidence of a taxonomically diverse
species of ferns and mosses [27]. This botanical diversity is evidence of a taxonomically diverse
ecosystem, where plant communities flourished in a continually-changing floodplain environment.
Despite the rigors of conducting research in a remote location with a harsh modern polar climate, the
Axel Heiberg fossil forest has been the subject of enthusiastic scientific inquiry. Published research
Geosciences 2018, 8, 223 5 of 30

ecosystem, where plant communities flourished in a continually-changing floodplain environment.


Despite the rigors of conducting research in a remote location with a harsh modern polar climate,
Geosciences 2018, 8, x FOR PEER REVIEW 5 of 32
the Axel Heiberg fossil forest has been the subject of enthusiastic scientific inquiry. Published research
includes
includes general
general descriptions
descriptions [19,22,28–32], as well
[19,22,28–32], as well as
as more
more specialized
specialized reports. These include
reports. These include analyses
analyses
of paleoclimate and paleoecology [15,18,33–40], and plant anatomy and taxonomy
of paleoclimate and paleoecology [15,18,33–40], and plant anatomy and taxonomy [14,41–43]. [14,41–43].

Figure 4. Axel Heiberg Island, Canadian Arctic. (A) Geodetic Hills site, 1999. (B) Metasequoia stumps
Figure 4. Axel Heiberg Island, Canadian Arctic. (A) Geodetic Hills site, 1999. (B) Metasequoia stumps
exposed by weathering. Photos courtesy of David Greenwood. (C) Metasequoia stump, 2000, photo by
exposed by weathering. Photos courtesy of David Greenwood. (C) Metasequoia stump, 2000, photo
William Hagopian. (D) Close view of Metasequoia wood, showing annual rings and desiccation
by William Hagopian. (D) Close view of Metasequoia wood, showing annual rings and desiccation
fractures. Photo from Kaelin et al. [44].
fractures. Photo from Kaelin et al. [44].

3.4. Banks Island—Middle Miocene-Pliocene


3.4. Banks Island—Middle Miocene-Pliocene
At Banks Island, the Miocene Ballast Brook Formation consists of sediments deposited on an
At Banks Island, the Miocene Ballast Brook Formation consists of sediments deposited on an
ancient floodplain, where an upper peat layer preserves cones, foliage, twigs, and logs (Figure 5)
ancient floodplain, where an upper peat layer preserves cones, foliage, twigs, and logs (Figure 5) [45].
[45]. In situ stumps are exposed at the top of the peat player; logs buried in the peat lie adjacent to
In situ stumps are exposed at the top of the peat player; logs buried in the peat lie adjacent to their
their corresponding stumps [46]. Separated by an unconformity, the overlying Beaufort Formation is
corresponding stumps [46]. Separated by an unconformity, the overlying Beaufort Formation is a
a Pliocene braided river deposit where logs and woody debris are preserved as transported material
Pliocene braided river deposit where logs and woody debris are preserved as transported material
(Williams 2006. Williams et al. 2008). Paleobotanical studies date to the 1960s, when University of
(Williams 2006. Williams et al. 2008). Paleobotanical studies date to the 1960s, when University of
Alberta professor Len Hills began studying Banks Island plant fossils [47–50].
Alberta professor Len Hills began studying Banks Island plant fossils [47–50].
Geosciences 2018, 8, 223 6 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 6 of 32

Figure 5. Miocene Beaufort Formation at Ballast Brook, Banks Island, Canadian Arctic. (A) General
Figure 5. Miocene Beaufort Formation at Ballast Brook, Banks Island, Canadian Arctic. (A) General
site view.
site view. (B)
(B)Tree
Treetrunk with
trunk intact
with root
intact crown.
root (C) Excavated
crown. trunktrunk
(C) Excavated showing branching.
showing 2005 photos
branching. 2005
courtesy of Christopher J. Williams. (D,E) Partially exposed logs, 2010, photos courtesy of William
photos courtesy of Christopher J. Williams. (D,E) Partially exposed logs, 2010, photos courtesy of
Hagopian.
William Hagopian.

3.5. Cornwallis Island—Miocene


3.5. Cornwallis Island—Miocene
A single sample of mummified collected from a roadcut at Resolute has been described as
A single sample of mummified collected from a roadcut at Resolute has been described as having
having elevated levels of Ca, Fe, and S relative to modern wood, but scanning and transmission
elevated levels of Ca, Fe, and S relative to modern wood, but scanning and transmission electron
electron microscopy show no evidence of mineralization [14].
microscopy show no evidence of mineralization [14].
3.6. Northwestern
3.6. Northwestern Canada
Canada
Late Paleocene/early
Late Paleocene/early Eocene
Eocene woods
woods have
have been
been preserved
preserved inin volcaniclastic
volcaniclastic material
material filling the
filling the
upper zones
upper zones ofofkimberlite
kimberlitepipes
pipesinin
thethe
LacLac
de de
GrasGras region
region of Slave
of Slave Province,
Province, Canada,
Canada, theofsite
the site manyof
many open pit diamond mines (Figure 6). The wood is inferred to represent a tree
open pit diamond mines (Figure 6). The wood is inferred to represent a tree from the boreal forest from the boreal
forest surrounding
surrounding the eruption
the eruption zone, wherezone, where
a trunk a trunk
collapsed intocollapsed into becoming
the diatreme, the diatreme, becoming
preserved in the
preserved in the sterile environment of the volcaniclastic kimberlite. Other mummified
sterile environment of the volcaniclastic kimberlite. Other mummified woods have been recovered woods have
been recovered from another diamond mine in the same region [51].
13 Cellulose
18
from another diamond mine in the same region [51]. Cellulose δ C, δ O and δ H values indicate δ2 13C, δ18O and δ2H

values
the indicateclimate
Paleogene the Paleogene climate of subarctic
of the Canadian the Canadian
was subarctic
12–17 ◦ C was
warmer12–17and
°C four
warmer andwetter
times four times
than
wetter than present
present [52,53]. [52,53].
Geosciences 2018, 8, 223 7 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 7 of 32

Geosciences 2018, 8, x FOR PEER REVIEW 7 of 32

Figure 6. Paleocene/Eocene mummified wood from kimberklite pipes, Subarctic Canada. (A,B)
Figure 6. Paleocene/Eocene mummified wood from kimberklite pipes, Subarctic Canada.
Metasequoia heartwood showing tyloses (arrow), Panda Pipe Ekati Diamond Mine. Specimen
(A,B) Metasequoia
provided byheartwood showing
Alex Wolfe. (C) tyloses
Piceaoxylon (arrow),
from Diavik PandaMine.
Diamond PipePhoto
EkatibyDiamond Mine.[52].
Benjamin Hook Specimen
provided by Alex Wolfe. (C) Piceaoxylon from Diavik Diamond Mine. Photo by Benjamin Hook
Figure 6. Paleocene/Eocene mummified wood from kimberklite pipes, Subarctic Canada. (A,B) [52].
3.7. United States ofheartwood
Metasequoia America showing tyloses (arrow), Panda Pipe Ekati Diamond Mine. Specimen
provided
3.7. United States by Alex Wolfe. (C) Piceaoxylon from Diavik Diamond Mine. Photo by Benjamin Hook [52].
of America
Mummified wood occurs in many locations, particularly in Pleistocene and Holocene
sediments. Examples where intact logs are preserved include the Farmdale Geosol in Illinois [54]
3.7. Unitedwood
Mummified States of Americain many locations, particularly in Pleistocene and Holocene sediments.
occurs
and Mt. Pleasant Bluff, Louisiana [55]. This report focuses on three locations that offer notable
Examples Mummified
where
examples intact
of woodlogs wood
are occurs
mummification inthat
preserved many
rangelocations,
include particularly
the from
in age Farmdale in Pleistocene
Geosol
late Miocene toin and
Illinois
late [54]Holocene
Pleistocene and Mt.7).Pleasant
(Figure
sediments. Examples where intact logs are preserved include the Farmdale Geosol
Bluff, Louisiana [55]. This report focuses on three locations that offer notable examples of wood in Illinois [54]
and Mt. Pleasant Bluff, Louisiana [55]. This report focuses on three locations that offer notable
mummification that range in age from late Miocene to late Pleistocene (Figure 7).
examples of wood mummification that range in age from late Miocene to late Pleistocene (Figure 7).

Figure 7. Mummified wood localities in USA. (1) Wilkes Formation, Washington, late Miocene.
(2) Northern Puget Sound, Washington, late Pleistocene. (3) Two Creeks Forest, Wisconsin, late
Pleistocene. (4) Farmdale Geosol, Illinois, late Pleistocene. (5) Mt. Pleasant Bluff, Louisiana,
late Pleistocene.
3.8. Wilkes Formation, Washington, USA—Late Miocene
The Wilkes formation consists of a stacked series of volcanic mudflow (lahar) sediments that
episodically inundated a lowland forest, alternating with lacustrine strata that were deposited
Geosciences 2018, 8, 223 8 of 30
during quiescent intervals. Abundant mummified wood is evidence of paleoenvironment: upright
stems where mudflows buried living vegetation (Figure 8A), and layers of woody peat that
represents
3.8. organic debris
Wilkes Formation, transported
Washington, by theMiocene
USA—Late mudflow (Figure 8B). Radiometric dating of a tephra
interbed gave a Late Miocene age of 6.13 ± 0.08 Ma [56]. Taxonomy of the wood has not been
The Wilkes
attempted, formation
but fossil pollenconsists
providesofevidence
a stackedofseries of volcanic
a swamp mudflow (lahar)
forest dominated sediments
by Taxodium (Swampthat
episodically inundated a lowland forest, alternating with lacustrine
Cypress) and Nyssa (Tupelo), with diverse angiosperms as minor constituents [57]. strata that were deposited during
quiescent intervals.quality
Preservation Abundant mummified
of the wood is evidence
wood is variable, typicallyofconsisting
paleoenvironment:
of wood thatupright
hasstems
been
where
compressed from burial pressure, and deformed by desiccation. Exposure to weathering organic
mudflows buried living vegetation (Figure 8A), and layers of woody peat that represents causes
debris
many transported
mummified by the mudflow
specimens to be(Figure 8B). (Figure
shattered Radiometric
9A,B),dating of a tephra
but intact interbed
specimens cangave
be afound
Late
Miocene
(Figure 9C).age of 6.13 ±mummified
Brown 0.08 Ma [56]. Taxonomy
wood of the wood
is abundant (Figurehas9D),
not been attempted,
as well as woodbut fossil
that haspollen
been
provides evidence of a swamp forest dominated by Taxodium (Swamp Cypress)
partially coalified (Figure 9E,F). Even in the freshest-looking specimens, tissues commonly and Nyssa (Tupelo),
show
with diverse
structural angiosperms
damage (Figure as minor constituents [57].
1C).

Figure 8. Miocene mummified wood in Wilkes Fm. Strata at Salmon Creek, Lewis County,
Figure 8. Miocene mummified wood in Wilkes Fm. Strata at Salmon Creek, Lewis County, Washington,
Washington, USA. (A) Upright stems buried by volcanic mudflow. (B) Transverse view of in situ
USA. (A) Upright stems buried by volcanic mudflow. (B) Transverse view of in situ stump, showing
stump, showing compression of original circular shape and deformation of annual rings. (C) Three
compression of original circular shape and deformation of annual rings. (C) Three thin wood mat layers
thin wood mat layers separated by thin mudflow sediments. (D) Close view of the lowest wood mat,
separated by thin mudflow sediments. (D) Close view of the lowest wood mat, showing flattening of
showing flattening of individual stems.
individual stems.

Preservation quality of the wood is variable, typically consisting of wood that has been compressed
from burial pressure, and deformed by desiccation. Exposure to weathering causes many mummified
specimens to be shattered (Figure 9A,B), but intact specimens can be found (Figure 9C). Brown
mummified wood is abundant (Figure 9D), as well as wood that has been partially coalified
(Figure 9E,F). Even in the freshest-looking specimens, tissues commonly show structural damage
(Figure 1C).
Geosciences 2018, 8, 223 9 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 9 of 32

Figure 9. Late Miocene wood from Wilkes Formation at Salmon Creek, Washington USA. (A,B)
Figure 9. LateHighly
Miocene wood from Wilkes Formation at Salmon Creek, Washington USA. (A,B) Highly
fractured wood in clay bed hat has been exposed to surface weathering. (C,D) Mummified
fractured wood woodin clay bed
typically showshat has been
compaction and exposed to surface
distortion from weathering.
burial pressures. (E,F) Some(C,D) Mummified wood
small specimens
typically shows showcompaction
incipient coalification.
and distortion from burial pressures. (E,F) Some small specimens show
incipient 3.9.
coalification.
Two Creek Forest, Wisconsin USA—Late Pleistocene
The Two Creeks forest bed is perhaps the most famous Quaternary deposit in the Great Lakes
3.9. Two Creekregion
Forest, WisconsinUSA
in north-central USA—Late Pleistocene
[58]. The widespread distribution of subfossil wood was first mapped in
1970 [59], though the Two Creeks fossil forest had been described a century earlier [60], with later
The Twostudies
Creeks forestinbed
published is perhaps
the 1930s the most
[61,62]. Radiometric famous
ages Quaternary
for the Two Creeks forestdeposit
sedimentsin the Great Lakes
average
13,500 14C years
region in north-central USA BP (Before
[58]. Present), equivalent to distribution
The widespread a calendar age ofof11,850 years BP
subfossil [57,63],was
wood a time first mapped
when spruce forest covered much of the Lake Michigan region following the recession of the great
in 1970 [59], though the Two Creeks fossil forest had been described a century earlier [60], with later
continental ice sheet. Tree-ring analyses show a duration age of the forest to have been 252 years [64]
studies published in the
to 310 years [65].1930s [61,62].
The wood Radiometric
is extremely agestypically
well preserved, for theshowing
Two Creeks
shrinkageforest
cracks,sediments
but few average
14 other structural defects (Figure 10). Detailed descriptions of the
13,500 C years BP (Before Present), equivalent to a calendar age of 11,850 years BP [57,63], a timegeologic setting and
when sprucepaleoenvironment
forest coveredcanmuch be found in references [59,66–69].
of the Lake Michigan region following the recession of the great
continental ice sheet. Tree-ring analyses show a duration age of the forest to have been 252 years [64]
to 310 years [65]. The wood is extremely well preserved, typically showing shrinkage cracks, but few
other structural defects (Figure 10). Detailed descriptions of the geologic setting and paleoenvironment
can be foundGeosciences
in references [59,66–69].
2018, 8, x FOR PEER REVIEW 10 of 32

Figure 10. Late Pleistocene Spruce log from Two Creeks Forest, east-central Wisconsin, USA.
Figure 10. Late Pleistocene Spruce log from Two Creeks Forest, east-central Wisconsin, USA.
3.10. Puget Sound, Washington—Late Pleistocene
Mummified wood is abundant in Late Pleistocene interglacial deposits in the northern Puget
Sound lowlands, northwest Washington, USA, where tree trunks, limbs, and small wood fragments
occur in the Whidbey Formation. The region experienced multiple advances and retreats of the
cordilleran ice sheet, but evidence of earlier events has largely been obscured by erosion and
deposition related to the last glacial episode, the Fraser Glaciation that occurred between
~25,000–10,500 years ago. An exception occurs at Whidbey Island, where Fraser glacial deposits are
Geosciences 2018, 8, 223 10 of 30

3.10. Puget Sound, Washington—Late Pleistocene


Mummified wood is abundant in Late Pleistocene interglacial deposits in the northern Puget
Sound lowlands, northwest Washington, USA, where tree trunks, limbs, and small wood fragments
occur in the Whidbey Formation. The region experienced multiple advances and retreats of the
cordilleran ice sheet, but evidence of earlier events has largely been obscured by erosion and deposition
related to the last glacial episode, the Fraser Glaciation that occurred between ~25,000–10,500 years
ago. An exception occurs at Whidbey Island, where Fraser glacial deposits are underlain by prominent
exposures of the much older Whidbey Formation. These sediments have ages that exceed the maximum
age that can be determined by 14 C (>40,000 years); thermoluminescence dates suggest an age of
150,000–100,000 years BP [70]. The Whidbey Formation has an exposed thickness of ~60 m at the type
section; other localities expose up to 90 m of strata. In general, the strata originated as fluvial and
lacustrine sediments deposited by meandering streams flowing along a coastal floodplain.
Wood samples used in this study were collected from coastal exposures at Swantown, Lagoon
Point, and Double Bluffs on Whidbey Island, and Cama Beach on nearby Camano Island (Figure 11).
Geosciences 2018, 8, x FOR PEER REVIEW 11 of 32

Figure 11. Late Pleistocene Whidbey Formation mummified wood localities, Whidbey and Camano
Figure 11. LateWashington
Islands, Pleistocene Whidbey
USA. Formation
(1) Swantown, mummified
(2) Lagoon wood localities,
Point, (3) Double Whidbey
Bluffs, (4) Cama Beach. and Camano
Islands, Washington USA. (1) Swantown, (2) Lagoon Point, (3) Double Bluffs, (4) Cama Beach.
Wood occurs in two sedimentary environments. Individual logs are preserved in thick sandy
beds, where they have been transported by fluvial processes. (Figure 12) Wood also occurs in beds of
Wood occurs in two sedimentary environments. Individual logs are preserved in thick sandy
woody peat, ranging in form from abundant flattened stems and small twigs to large branches and
beds, where they
rare tree havethat
trunks been transported
accumulated by fluvial
in local processes.
bogs (Figure 13). Peat(Figure 12) at
beds occur Wood alsolevels
multiple occurs in beds of
in the
woody Whidbey
peat, ranging in form
Formation, from abundant
interspersed with sandyflattened stems(Figure
fluvial deposits and small
14). twigs to large branches and
rare tree trunks that accumulated in local bogs (Figure 13). Peat beds occur at multiple levels in the
Whidbey Formation, interspersed with sandy fluvial deposits (Figure 14).
Geosciences 2018, 8, 223 11 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 12 of 32

Figure 12. Logs in Whidbey Formation sand beds, Swantown beach, Whidbey Island, Washington
Figure 12.
USA.Logs in Whidbey
(A) View Formation
of Whidbey Formation sand
strata. beds, Swantown
(B–D) Logs exposedbeach, Whidbey Island, Washington
by weathering.
USA. (A) View of Whidbey Formation strata. (B–D) Logs exposed by weathering.
Geosciences 2018, 8, x FOR PEER REVIEW 13 of 32
The Whidbey Formation wood is well-known to local geoscientists, but this material has
received little scientific study. Poor anatomical preservation hinders taxonomic identification of the
wood, but the forest that produced it is known from pollen obtained from Whidbey Formation peat
beds at the Double Bluffs type locality (Figure 14) [71,72]. Pinus contorta (Lodgepole Pine), Tsuga
heterophylla (Western Hemlock), and Pseuodtsuga menziesii (Douglas Fir) are abundant elements in
the palynoflora. Pollen analysis from the main wood-bearing peat bed in the Whidbey Formation
type section contains the above-mentioned conifers as the most abundant taxa, but small amounts of
Juniperus (Juniper), Abies (Balsam Fir), and Betula (Birch) are present. A thin peat bed ~20 m higher in
the stratigraphic section is dominated by Alnus (Alder) [72]. The pollen flora indicates that during
the Whidbey Formation interglacial period the region was inhabited by boreal conifer forest that
closely resembles the modern flora.
Preservation quality is variable, largely affected by compression from the weight of overlying
sediments. Small twigs and stems in woody peat layers are commonly highly flattened, but larger
limbs and trunks are more likely to retain three-dimensional shapes. Regardless of shape, it is rare
for specimens to preserve uncompressed cells (Figures 1A and 15A-D).

Figure 13. Logs in Whidbey Formation peat, Double Bluff, Whidbey Island, Washington USA. (A)
Logs instrata
Figure 13.Pleistocene Whidbey
at DoubleFormation
Bluffs. Largepeat, Double occur
wood fragments Bluff,in Whidbey Island,
2 m-thick peat Washington
beds (arrow). (B) USA.
Blocks of
(A) Pleistocene peat that
strata have fallen
at Double to lie atLarge
Bluffs. the base of the fragments
wood bluff. (C,D) Wood
occurin in
fallen peat blocks.
2 m-thick peat beds (arrow).
(B) Blocks of peat that have fallen to lie at the base of the bluff. (C,D) Wood in fallen peat blocks.
Geosciences 2018, 8, 223 12 of 30

The Whidbey Formation wood is well-known to local geoscientists, but this material has received
little scientific study. Poor anatomical preservation hinders taxonomic identification of the wood,
but the forest that produced it is known from pollen obtained from Whidbey Formation peat beds at
the Double Bluffs type locality (Figure 14) [71,72]. Pinus contorta (Lodgepole Pine), Tsuga heterophylla
(Western Hemlock), and Pseuodtsuga menziesii (Douglas Fir) are abundant elements in the palynoflora.
Pollen analysis from the main wood-bearing peat bed in the Whidbey Formation type section contains
the above-mentioned conifers as the most abundant taxa, but small amounts of Juniperus (Juniper),
Abies (Balsam Fir), and Betula (Birch) are present. A thin peat bed ~20 m higher in the stratigraphic
section is dominated by Alnus (Alder) [72]. The pollen flora indicates that during the Whidbey
Formation interglacial period the region was inhabited by boreal conifer forest that closely resembles
Geosciences
the modern 2018, 8, x FOR PEER REVIEW
flora. 14 of 32

Figure14.
Figure 14.Stratigraphic
Stratigraphicsection
sectionofoflate
latePleistocene
PleistoceneWhidbey
WhidbeyFormation
Formationofofinterglacial
interglacialsediment
sedimentatat
Double Bluff, Whidbey Island, Washington USA. Adapted from
Double Bluff, Whidbey Island, Washington USA. Adapted from [73]. [73].

Preservation quality is variable, largely affected by compression from the weight of overlying
sediments. Small twigs and stems in woody peat layers are commonly highly flattened, but larger
limbs and trunks are more likely to retain three-dimensional shapes. Regardless of shape, it is rare for
specimens to preserve uncompressed cells (Figures 1A and 15A–D).
Geosciences 2018, 8, 223 13 of 30

Whidbey Formation wood is not visibly mineralized, the only exception being a few specimens
that have thin8,coatings
Geosciences 2018, vivianite (Fe2+ Fe2+ 2 (PO4 )2 ·8H2 O), (Figure 16).
of REVIEW
x FOR PEER 15 of 32

Figure 15. Scanning


Figure 15. Scanning electron
electron microscopy
microscopy (SEM)
(SEM) photos
photos ofof mummified
mummified conifer
conifer wood
wood fromfrom the
the late
late
Pleistocene Whidbey Formation, Whidbey Island, Washington USA. (A) Wood
Pleistocene Whidbey Formation, Whidbey Island, Washington USA. (A) Wood from peat bed at Double from peat bed at
Double Bluff. Radial
Bluff. Radial view ofview of tracheids,
tracheids, showing
showing circular
circular pits (arrow).
pits (arrow). (B) Wood
(B) Wood fromfrom
log log in sand
in sand bedbed
at
at Swantown beach. Cell walls are composed of relict organic matter, but anatomical
Swantown beach. Cell walls are composed of relict organic matter, but anatomical details are mostly details are
mostly obliterated.
obliterated. (C) Pits (C) Pitscells
in ray in ofraywood
cells from
of wood
Double from
BluffDouble Bluff peat,
peat, showing showing
excellent excellent
preservation.
preservation.
(D) Radial view of wood from Swantown log, showing two pits on a cell that has altered surfacethat
(D) Radial view of wood from Swantown log, showing two pits on a cell has
texture.
altered surface texture. (E) Transverse view of wood from a limb buried in peat bed
(E) Transverse view of wood from a limb buried in peat bed at Double Bluff. Cells walls preserve at Double Bluff.
Cells walls
original preserve architecture,
multilayered original multilayered
but cells arearchitecture, but cells
highly compressed. (F)are highly view
Transverse compressed.
of the same(F)
Transverse viewmagnification,
wood at higher of the same wood
showingat higher magnification,
the sigmoidal shape showing
of a singlethe sigmoidal tracheid.
compressed shape of a single
compressed tracheid.

Whidbey Formation wood is not visibly mineralized, the only exception being a few specimens
that have thin coatings of vivianite (Fe2+Fe2+2(PO4)2·8H2O), (Figure 16).
Geosciences 2018, 8, 223 14 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 16 of 32

Figure
Figure 16.
16. Late
Late Pleistocene
Pleistocene wood
wood from
from northwest
northwest Washington,
Washington, USA,
USA, containing
containing coatings
coatings of
of blue
blue
vivianite
vivianite (arrows).
(arrows).

3.11.
3.11. European
European Locations
Locations
Mummified
Mummified woodwood occurs
occurs atatmany
manylocations
locationsininthroughout
throughoutthetheworld,
world,including
including numerous
numerous sites
sites in
in
Europe. Examples that have been reported in scientific literature include localities in Belgium [74],
Europe. Examples that have been reported in scientific literature include localities in Belgium [74],
England
England [75],
[75], Austria
Austria [76],
[76], and
and Siberia
Siberia [77,78].
[77,78]. Other
Othermummified
mummifiedwoodwoodsites
sitesare
arelocated
locatedininGermany,
Germany,
Poland,
Poland, Czech Republic, and The Netherlands. The following section describes some of the
Czech Republic, and The Netherlands. The following section describes some of the most
most
spectacular occurrences at localities where efforts have been made to display evidence
spectacular occurrences at localities where efforts have been made to display evidence of fossil offorests
fossil
forests for viewing.
for public public viewing.

3.12.
3.12. Ipolytarnóc
Ipolytarnóc Fossil
Fossil Forest,
Forest, Hungary—Late
Hungary—Late Miocene
Miocene
In
In July
July 2007,
2007, miners
miners working
working at at an
an open-pit
open-pit coal
coal mine
mine at
at Bükkábrány,
Bükkábrány, Hungary
Hungary discovered
discovered 16 16
upright treetrunks
upright tree trunksininthethe
sandsand
layerlayer
just just
above above the lignite
the lignite coal(Figure
coal seam seam 17A).
(Figure 17A).ofAgeologists
A team team of
geologists and paleontologists
and paleontologists was quickly wasorganized
quickly organized to study
to study this this spectacular
spectacular occurrence, occurrence,
the only the only
location
location
worldwide worldwide where
where large large
trees trees are preserved
are preserved intendingintending in theirforest
in their original original forest
setting setting
[79,80]. [79,80].
The fossil
The fossil
forest forest isto
is believed believed to have originated
have originated when rising whenlakerising lakedrowned
waters waters drowned
the forestthe forest the
during during
late
the late Miocene,
Miocene, with thewith
trunks theprotected
trunks protected
from decay fromby decay by their
their deep deep
burial burial in water-saturated
in water-saturated sediment.
sediment. Thebeen
The trees have treesidentified
have been identified as Glyptostroboxylon
as Glyptostroboxylon rudolphii
rudolphii (Figure 17B), a taxon(Figure 17B), described
previously a taxon
previously
from a site described
in Germany from a site In
[81,82]. in 2010,
Germany a new [81,82]. Incenter
visitor 2010, awas
newopened
visitor atcenter was openedFossils
the Ipolytarnóc at the
Ipolytarnóc
Nature Reserve,Fossils Nature
where fiveReserve, wherefrom
large stumps five Bükkábrány
large stumpsare from Bükkábrány
displayed (Figureare17C,D).
displayed (Figure
17C,D).
The fossil trees have commonly been described as being mummified, with the wood soft enough
to cutThe
withfossil trees
a razor have
blade commonly
after specimensbeen weredescribed
soaked in as beingwater
boiling mummified, with theexamination
[80]. Microscopic wood soft
enough to cutanalyses
and chemical with a confirm
razor blade afterwood
that the specimens were soaked
of the buried in boilingpristine;
trees is relatively water [80].
levelsMicroscopic
of cellulose
examination
and phenolicand chemicalin
compounds analyses confirm
wood within thethat the woodlignite
underlying of theare
buried trees is relatively
significantly pristine;
lower, evidence of
levels of cellulose
incipient and
gelification phenolic compounds in wood within the underlying lignite are significantly
[83].
lower, evidence of incipient gelification [83].
Geosciences 2018, 8, 223 15 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 17 of 32

Figure 17. Fossil forest at Bükkábrány, northern Hungary. (A) Miocene stumps excavated in 2007
Figure 17. Fossil forest at Bükkábrány, northern Hungary. (A) Miocene stumps excavated in 2007 open
open pit lignite mine. (B) Glyptostroboxylon rudolphii wood. Photo by Zsigmund Atilla. (C) Large
pit lignite mine. (B) Glyptostroboxylon rudolphii wood. Photo by Zsigmund Atilla. (C) Large stumps that
stumps that have been relocated for outdoors display at Ipolytarnóc Fossils Nature Center, opened in
have been relocated for outdoors display at Ipolytarnóc Fossils Nature Center, opened in 2010. Photo
2010. Photo courtesy of Tamás Elter. (D). Interior of Nature Center. Photo (D) courtesy of Sándor
courtesy of Tamás Elter. (D). Interior of Nature Center. Photo (D) courtesy of Sándor Kovács.
Kovács.

3.13. Dunarobba Fossil Forest, Italy—Pliocene


3.13. Dunarobba Fossil Forest, Italy—Pliocene
Discovered during clay mining in 1980, Dunarobba Fossil Forest is preserved in lacustrine
Discovered during clay mining in 1980, Dunarobba Fossil Forest is preserved in lacustrine
sediments of the Tiberino Basin, a Plio-Plesitocene intermontane basin along the western margin of
sediments of the Tiberino Basin, a Plio-Plesitocene intermontane basin along the western margin of
the Apenine Mountains in central Italy [84]. More than 50 upright trunks up to 2.5 m in diameter and
the Apenine Mountains in central Italy [84]. More than 50 upright trunks up to 2.5 m in diameter and
9 m in height are preserved in clay and fine sand sediment, representing a Pliocene conifer forest that
9 m in height are preserved in clay and fine sand sediment, representing a Pliocene conifer forest
flourished along a marshy lake shore ~3–2 Ma (Figure 18). Originally named as a member of the form
that flourished along a marshy lake shore ~3–2 Ma (Figure 18). Originally named as a member of the
genus Taxodioxylon [83], the dominant species is now considered to be Glyptostrobus europaeus based on
form genus Taxodioxylon [83], the dominant species is now considered to be Glyptostrobus europaeus
the “whole plant” association of foliage, cones, and wood [85]. The wood is commonly described as
based on the “whole plant” association of foliage, cones, and wood [85]. The wood is commonly
mummified, resulting from the protection against degradation provided by the impermeable clay-rich
described as mummified, resulting from the protection against degradation provided by the
matrix, which restricted influx of oxygenated groundwater. However, the composition of Dunarobba
impermeable clay-rich matrix, which restricted influx of oxygenated groundwater. However, the
wood has complexities. Chemical analyses reveal that leaching has removed most polyoses and
composition of Dunarobba wood has complexities. Chemical analyses reveal that leaching has
cellulose, leaving lignin as the major structural component [86,87]. Some specimens are partially
removed most polyoses and cellulose, leaving lignin as the major structural component [86,87].
mineralized with calcite, chlorapatite, chloromagnesite, and clays [88]. Permineralized wood collected
Some specimens are partially mineralized with calcite, chlorapatite, chloromagnesite, and clays [88].
from clays above the main fossil forest layer are mineralized by a combination of siderite, goethite, and
Permineralized wood collected from clays above the main fossil forest layer are mineralized by a
ferroan calcite [89,90]. Mineral composition may vary within a single specimen, evidence that mineral
combination of siderite, goethite, and ferroan calcite [89,90]. Mineral composition may vary within a
precipitation was strongly influenced by localized changes in Eh and pH.
single specimen, evidence that mineral precipitation was strongly influenced by localized changes in
Eh and pH.
Conservation efforts in the 1990s led to the establishment of Dunrobba Botanic Palaeontology
Centre, which includes both a small exhibition hall and a path that directs visitors to outdoor
displays where upright stumps are protected by individual A-frame shelters. Preservation efforts
continue in an ongoing attempt to minimize degradation caused by microbial decay and
atmospheric exposure.
Geosciences 2018, 8, 223 16 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 18 of 32

Figure 18. Dunarobba Fossil Forest, Umbria region of central Italy. (A) Fossil forest, circa 1995, prior
Figure 18. Dunarobba Fossil Forest, Umbria region of central Italy. (A) Fossil forest, circa 1995, prior to
to conservation. Wikipedia photo. (B) One of many outdoor stumps protected by shelters at
conservation. Wikipedia photo. (B) One of many outdoor stumps protected by shelters at Dunarobba
Dunarobba Fossil Forest. Copyrighted photo from www.alamy.com, used with permission. (C) View
Fossil Forest. Copyrighted photo from www.alamy.com, used with permission. (C) View of stump
of stump within its shelter. (D) Close-up view of mummified wood showing shrinkage cracks.
within its shelter. (D) Close-up view of mummified wood showing shrinkage cracks. Photos (C,D) by
Photos Bianchi,
Franco (C,D) bycourtesy
Franco Bianchi,
of Martacourtesy of Marta Pinzaglia.
Pinzaglia.

3.14. Fossano Fossil Forest, Northern Italy—Pliocene


Conservation efforts in the 1990s led to the establishment of Dunrobba Botanic Palaeontology
A major
Centre, whichPliocene
includes fossil
both aforest
smalloccurs in northwestern
exhibition Italy,
hall and a path thatwhere
directsupright
visitorsstumps are displays
to outdoor exposed
where upright stumps are protected by individual A-frame shelters. Preservation efforts continue as
along the banks of the Stura di Dimonte River near the small town of Fossano. The site is known in
theongoing
an Fossanoattempt
Fossil Forest. The site
to minimize is in proximity
degradation to microbial
caused by the alps; decay
the river
andhas incised Quaternary
atmospheric exposure.
sediments to expose Pliocene continental sediments (Figure 19). Two successive forests are
represented
3.14. Fossano in separate
Fossil Forest,stratigraphic layers [91–93]. As in other Pliocene swamp forests in Italy,
Northern Italy—Pliocene
Glyptostrobus europaeusfossil
A major Pliocene is theforest
dominant
occurstree type. The scientific
in northwestern significance
Italy, where uprightof stumps
this newly-exposed
are exposed
along the banks of the Stura di Dimonte River near the small town of Fossano. The siteyears.
fossil forest is sure to result in additional descriptions and interpretations in future is known as the
Fossano Fossil Forest. The site is in proximity to the alps; the river has incised Quaternary sediments to
expose Pliocene continental sediments (Figure 19). Two successive forests are represented in separate
stratigraphic layers [91–93]. As in other Pliocene swamp forests in Italy, Glyptostrobus europaeus is the
dominant tree type. The scientific significance of this newly-exposed fossil forest is sure to result in
additional descriptions and interpretations in future years.
Geosciences 2018, 8, 223 17 of 30
Geosciences 2018, 8, x FOR PEER REVIEW 19 of 32

Figure 19. Fossano Fossil Forest, Italy. (A,B) Upright stumps of Glyptostrobus europaeus exposed in
Figure 19. Fossano Fossil Forest, Italy. (A,B) Upright stumps of Glyptostrobus europaeus exposed in
Pliocene sediment that forms the bed of the Stura di Dimonte River, NW Italy. (C) Wood exposed to
Pliocene sediment that forms the bed of the Stura di Dimonte River, NW Italy. (C) Wood exposed to
surface conditions. Photos courtesy of Edoardo Martinetto.
surface conditions. Photos courtesy of Edoardo Martinetto.
3.15. Stura di Lanzo Fossil Forest, Northwest Italy—Pliocene
3.15. SturaLike
di Lanzo Fossil Forest, Northwest Italy—Pliocene
the Fossano Fossil Forest, the Stura de Lanzo locality is a site where mummified Pliocene
tree trunks
Like have been
the Fossano exposed
Fossil bythe
Forest, river erosion
Stura [94–96].locality
de Lanzo The muddy sediments
is a site that enclose large
where mummified Pliocene
Glyptostrobus stumps also preserve seeds and cones, allowing “whole plant” identification
tree trunks have been exposed by river erosion [94–96]. The muddy sediments that enclose [85]. Thelarge
Stura di Lanzo Fossil Forest has been described in detail in a 2005 booklet [97] published in Italian by
Glyptostrobus stumps also preserve seeds and cones, allowing “whole plant” identification [85].
Parco la Mandria (Mandria Provincial Park). Conservation efforts at the site have included
The Stura di Lanzo Fossil Forest has been described in detail in a 2005 booklet [97] published in
participation of teachers and young students [98,99].
Italian by Parco la Mandria (Mandria Provincial Park). Conservation efforts at the site have included
3.16. Other
participation ofLocations
teachers and young students [98,99].
InLocations
3.16. Other East Asia, Oligocene mummified wood occurs in the Nanning Basin of southern China. The
deposit includes fossilized mollusks, fish, reptiles, and mammals as well as plants. Wood occurs as
stumps
In East and trunks
Asia, in a lagerstätte
Oligocene that also
mummified preserves
wood foliage
occurs andNanning
in the seeds [100]. Pliocene
Basin mummified
of southern China.
wood has
The deposit been described
includes from
fossilized central Japan
mollusks, fish,[101].
reptiles, and mammals as well as plants. Wood occurs
as stumps and trunks in a lagerstätte that also preserves foliage and seeds [100]. Pliocene mummified
4. Charcoalified Wood
wood has been described from central Japan [101].
publications [102–106]. Several papers provide general overviews of fossil charcoal research
[104-106].
Because charred wood typically has excellent anatomical preservation, intact charcoal
specimens are well suited for taxonomic identifications. For relatively young deposits, charcoal is in
an ideal material for 14C radiometric dating. Like other plant fossils, charcoal can be used to
Geosciences 2018, 8, 223 18 of 30
reconstruct paleoenvironmental conditions. On a global scale, ignition temperature of plant material
is related to atmospheric oxygen levels, so the presence of charcoal can be used to trace atmospheric
evolution [107].
4. Charcoalified Wood
Charring of wood by volcanic processes has been described for the Soufrière Hills Volcano,
Monserrat
Fire has been[103], the Miocene
present volcanicthe
throughout field at Cormandel
Earth’s history,Peninsula, New Zealand
so the abundance [108] andwood
of charred Recentin the
ignimbrite
geologic recordinisthenotTaupo volcano,Production
surprising. New Zealand [109]. In the
of charcoal haslatter
two example,
primary charring
causes: temperatures
natural wildfires
estimated
(Figure 20), and from optical of
ignition reflectance
wood by ranged from flows
volcanic 267 °C or
to 414
hot °C, generally
volcanic ashdecreasing at distances
(ignimbrite) (Figure 21).
away from the volcanic vent. At Montseratt, charring temperatures are estimated to mainly be in the
Charcoal in the fossil record ranges from intact logs to tiny fragments dispersed in clastic sediments
range of 200°C–340 °C, with a possible maximum of ~450 °C [104]. In volcanic environments, it is
and coal. The latter form is commonly called fusain. Regardless of size, ancient charcoal may yield
common for tree trunks to be completely charred, though some specimens are less charred in
important
interiorinformation
regions. about Earth history.

Figure 20. Charcoalification of wood in modern wild fires. (A) California, USA, 2015. Photo: National
Figure 20. Charcoalification of wood in modern wild fires. (A) California, USA, 2015. Photo: National
Atmospheric & Space Administration. (B) Charred trees after 2007 fire, Minnesota, USA. Photo
Atmospheric
courtesy &
of Space Administration. (B) Charred trees after 2007 fire, Minnesota, USA. Photo courtesy
Eli Sagor.
of Eli Sagor.
Geosciences 2018, 8, x FOR PEER REVIEW 21 of 32

Figure 21. Charred wood resulting from volcanic processes. (A) Tree ignited by basalt flow from
Figure 21. Charred wood resulting from volcanic processes. (A) Tree ignited by basalt flow from
Kilauea Volcano, Hawaii, USA. Photo by Brian W. Schaller (Wikipedia). (B) Charred trees after 1959
Kilauea Volcano, Hawaii, USA. Photo by Brian W. Schaller (Wikipedia). (B) Charred trees after 1959
Kilauea eruption. USGS Hawaii Volcano Observatory photo. (C) Tephra cloud engulfing palm grove,
Kilauea eruption. USGSeruption,
1991 Mt. Pinatubo Hawaii Philippine
Volcano Observatory photo.
Islands. Photo (C) Tephra
by Alberto cloud
Garcia. engulfing
(D) Recent palmingrove,
charcoal
1991 ignimbrite, Taupo
Mt. Pinatubo volcanic field,
eruption, New Zealand.
Philippine Islands.Photo by Brent
Photo Alloway.
by Alberto Garcia. (D) Recent charcoal in
ignimbrite, Taupo volcanic field, New Zealand. Photo by Brent Alloway.
4.1. Evidence of Ancient Wildfires
Fossil evidence of wildfires primarily comes from charcoal. The oldest known charcoal comes
from the Silurian [110]. However, wildfires did not become prominent until the appearance of
forests in the Middle Devonian [111,112]. By the Carboniferous, high-biomass forests were
episodically swept by fires, ultimately producing coal that contains as much as 20% charcoal by
volume [99]. The occurrence of fire is related to atmospheric oxygen levels, which determine the
ignition temperature of organic materials. Atmospheric oxygen levels during Phanerozoic time was
estimated from the organic carbon and sulfur content of clastic sediments [113–115], but the
Geosciences 2018, 8, 223 19 of 30

Much of our knowledge of ancient charcoal comes from the work of University of London
professor Andrew C. Scott. His research is described in a recent book [96], and in many journal
publications [102–106]. Several papers provide general overviews of fossil charcoal research [104-106].
Because charred wood typically has excellent anatomical preservation, intact charcoal specimens
are well suited for taxonomic identifications. For relatively young deposits, charcoal is in an ideal
material for 14 C radiometric dating. Like other plant fossils, charcoal can be used to reconstruct
paleoenvironmental conditions. On a global scale, ignition temperature of plant material is related to
atmospheric oxygen levels, so the presence of charcoal can be used to trace atmospheric evolution [107].
Charring of wood by volcanic processes has been described for the Soufrière Hills Volcano,
Monserrat [103], the Miocene volcanic field at Cormandel Peninsula, New Zealand [108] and Recent
ignimbrite in the Taupo volcano, New Zealand [109]. In the latter example, charring temperatures
estimated from optical reflectance ranged from 267 ◦ C to 414 ◦ C, generally decreasing at distances
away from the volcanic vent. At Montseratt, charring temperatures are estimated to mainly be in
the range of 200◦ C–340 ◦ C, with a possible maximum of ~450 ◦ C [104]. In volcanic environments,
it is common for tree trunks to be completely charred, though some specimens are less charred in
interior regions.

4.1. Evidence of Ancient Wildfires


Fossil evidence of wildfires primarily comes from charcoal. The oldest known charcoal comes
from the Silurian [110]. However, wildfires did not become prominent until the appearance of forests in
the Middle Devonian [111,112]. By the Carboniferous, high-biomass forests were episodically swept by
fires, ultimately producing coal that contains as much as 20% charcoal by volume [99]. The occurrence
of fire is related to atmospheric oxygen levels, which determine the ignition temperature of organic
materials. Atmospheric oxygen levels during Phanerozoic time was estimated from the organic carbon
and sulfur content of clastic sediments [113–115], but the usefulness of fossil charcoal for making more
precise evaluations was quickly recognized. A minimum oxygen concentration of 13% is required for
wildfires; oxygen levels above 35% would reduce ignition temperatures to a point where sustained
forest growth would be prohibited [114]. Atmospheric oxygen concentrations are estimated to have
been above 26% throughout the Carboniferous and Permian periods, declining abruptly near the time
of the Permian-Triassic mass extinction, fluctuating during the Triassic and Jurassic, and declining
beginning in the mid Cretaceous to reach the present-day value of ~21% [116–118]. Dispersed charcoal
is commonly preserved in clastic sediments, but also occurs in lignite coal [119].

4.2. Anatomical Evidence


Dispersed charcoal fragments are useful for tracing wildfire history, but macroscopic specimens
are valuable for taxonomic purposes. Light microscopy can be used, but SEM images are richer in
detail. Although most research has been done on charred wood, vascular plants (e.g., ferns) may
also be represented as fossil charcoal [120]. In higher plants, charred tissues become black and brittle,
with a silky sheen. Because charcoal results from heating, compression forces are limited to minor
shrinkage effects, so wood retains its 3-dimensional structure The reflectance of the charcoal can be
used to estimate charring temperature [105, 107] The most significant structural change is the fusing
of the cell walls into a single homogeneous layer, obliterating the original multilayered architecture.
Experimental studies show that cell wall homogenization may occur when modern conifer wood is
heated for one hour at 350 ◦ C [105,116]. In other aspects, anatomical preservation may be excellent
(Figure 22).
used to estimate charring temperature [105, 107] The most significant structural change is the fusing
of the cell walls into a single homogeneous layer, obliterating the original multilayered architecture.
Experimental studies show that cell wall homogenization may occur when modern conifer wood is
heated for
Geosciences one
2018, hour at 350 °C [105,116]. In other aspects, anatomical preservation may be excellent
8, 223 20 of 30
(Figure 22).

Figure
Figure 22.22. Modern
Modern charred
charred wood
wood from
from Whatcom
Whatcom County,
County, Washington
Washington USA. USA. (A)
(A) Thuja
Thuja plicata
plicata (Red
(Red
Cedar), wildfirewood;
Cedar), wildfire wood;(B–D)
(B–D)AcerAcer macrophyllum
macrophyllum (Bigleaf
(Bigleaf Maple),
Maple), campcamp fire wood.
fire wood. (B) Tangential
(B) Tangential view,
view, showing cross-section of a ray with a conspicuous resin canal. Adjacent tracheid
showing cross-section of a ray with a conspicuous resin canal. Adjacent tracheid cell walls show cell brittle
walls
show brittle
fracture. fracture. (C),
(C), Transverse viewTransverse view showing
showing earlywood (largeearlywood (large cells)
cells) and latewood (smalland latewood
cells). (small
(E,F) Charred
cells). (E,F) Charred conifer wood from an excavated landfill. (E) Tangential view,
conifer wood from an excavated landfill. (E) Tangential view, showing cross-section view of ray cells. showing
cross-section
(F) Radial view view of ray well-preserved
showing cells. (F) Radialbordered
view showing
pits. well-preserved bordered pits.

A few occurrences have been reported where partially-charred wood has become silicified
A few occurrences have been reported where partially-charred wood has become
[108,121]. A previously unreported occurrence is represented by a single specimen from the Miocene
silicified [108,121]. A previously unreported occurrence is represented by a single specimen from
Columbia Plateau basalts of central Washington, USA (Figure 23A). Although the opal mines at
the Miocene Columbia Plateau basalts of central Washington, USA (Figure 23A). Although the opal
Virgin Valley are famous for silicified wood, a few samples show charring, presumably from
mines at Virgin Valley are famous for silicified wood, a few samples show charring, presumably from
pyroclastic flows that affected forests bordering the ancient caldera. These examples provide
pyroclastic flows that affected forests bordering the ancient caldera. These examples provide evidence
evidence that charred wood is not susceptible to silicification, probably because the pyrolized tissues
that charred wood is not susceptible to silicification, probably because the pyrolized tissues are not
capable of acting as organic templating matrices for the precipitation of silica. However, uncharred
interior regions and empty cell lumen may be sites for silica precipitation (Figure 23B,C).
Geosciences 2018, 8, x FOR PEER REVIEW 23 of 32

are not
Geosciences 2018, capable of acting as organic templating matrices for the precipitation of silica. However,
8, 223 21 of 30
uncharred interior regions and empty cell lumen may be sites for silica precipitation (Figure 23B,C).

Figure 23. Silicification of charred wood. (A) Miocene wood from Columbia River Basalt Group,
Figure 23. Silicification of charred wood. (A) Miocene wood from Columbia River Basalt Group, central
central Washington, USA. Exterior charred layer surrounds silicified interior wood. (B,C) Miocene
Washington,
wood USA.
fromExterior charred
Royal Peacock Opallayer
Mine,surrounds silicified
Virgin Valley, Nevadainterior wood. homogenization
USA, showing (B,C) Mioceneofwood from
charredOpal
Royal Peacock cell walls.
Mine, Cell lumen Valley,
Virgin contain lepispheres
Nevada USA,of opal-CT.
showing homogenization of charred cell walls.
Cell lumen contain lepispheres of opal-CT.
5. Coalified Wood
Mummified wood retains a high proportion of the original cellulosic constituents, in contrast to
5. Coalified Wood
coalified wood, where loss of cellulose and hemicellulose results in relative enrichment of lignin [73].
During coalification,
Mummified wood retains woodaand other
high plat remainsof
proportion arethe
transformed
originaltocellulosic
peat as a result of humification
constituents, in contrast to
and gelification,
coalified wood, where loss which are biochemical
of cellulose processes. Humification
and hemicellulose resultsinvolves
in relativethe enrichment
decompositionof of lignin [73].
original organic constituents to produce dark brown organic polymers. Because these polymers are
During coalification, wood and other plat remains are transformed to peat as a result of humification
resistant to microbial attack, peat is a relatively stable material. The amorphous nature of these
and gelification,
organic which areproducts
alteration biochemical
causes processes.
plant tissuesHumification involves
to lose their original the decomposition
architecture, producing a of original
gelatinous texture.
organic constituents The conversion
to produce dark brown of peat to coalpolymers.
organic is a metamorphic
Because process,
these where
polymers organic
are resistant
compounds are altered as a result of heat and pressure, leading to an increase
to microbial attack, peat is a relatively stable material. The amorphous nature of these organic in carbon content and
decreases in oxygen and hydrogen. These reactions involve dehydration, bituminization, and
alterationgraphitization,
products causes plant tissues to lose their original architecture, producing a gelatinous
and they can be divided into four stages: peat is successively changed into lignite,
texture. The conversion of peat
subbituminous coal, bituminous, to coal is a andmetamorphic
anthracite process, whereis organic
[122]. Coal composed compounds
of three are altered
as a resultmicroscopically-recognizable
of heat and pressure, leading groups to of an
organic constituents
increase in carbonthat are knownand
content as macerals.
decreases Derived
in oxygen and
hydrogen.fromThesecellular tissues,involve
reactions vitrinite is the most abundant
dehydration, and most homogeneous
bituminization, maceral, theand
and graphitization, mainthey can be
contributor to the shiny black appearance of coal. U.S. Coals typically contain as much as 80 wt %
divided into four stages: peat is successively changed into lignite, subbituminous coal, bituminous, and
vitrinite [123]. Liptinite (also called exinite) originates from spores, pollen, resins, lipids, waxes,
anthracitealgae,
[122].andCoal is composed
bacterial of threeismicroscopically-recognizable
proteins. Inertinite derived from oxidation. Coalified wood groups of organic
is most commonly constituents
found inas
that are known themacerals.
lowest-rank coals, lignite
Derived from andcellular
subbituminous,
tissues, where the macroscopic
vitrinite is the most organic remains and most
abundant
homogeneousmay be preserved.
maceral, thePreservation of wood in
main contributor to lignite (brown
the shiny coal)appearance
black is very common (e.g., [124–126]).
of coal. U.S. Coals typically
contain as much as 80 wt % vitrinite [123]. Liptinite (also called exinite) originates from spores, pollen,
resins, lipids, waxes, algae, and bacterial proteins. Inertinite is derived from oxidation. Coalified wood
is most commonly found in the lowest-rank coals, lignite and subbituminous, where the macroscopic
organic remains may be preserved. Preservation of wood in lignite (brown coal) is very common
(e.g., [124–126]). However, cellular tissues in a mineralized Permian tree fern have been observed to
consist of anthracite [127]. Conversion of original tissue to vitrinite involves the diagenetic alteration
of lignocelluloses. In general, both hemi-cellulose and cellulose are greatly reduced or destroyed, and
lignin-based components are selectively enriched. [127,128]. Key drivers of vitrinite formation are
removal of cellulosic materials followed by physical compression. The result is that cellular anatomy
is likely to be destroyed, producing logs and limbs that retain their original form, but often with
significant flattening. Coalified wood typically has a vitreous or semi-vitreous luster, brittle with
Geosciences 2018, 8, x FOR PEER REVIEW 24 of 32

However, cellular tissues in a mineralized Permian tree fern have been observed to consist of
anthracite [127]. Conversion of original tissue to vitrinite involves the diagenetic alteration of
lignocelluloses. In general, both hemi-cellulose and cellulose are greatly reduced or destroyed, and
lignin-based components are selectively enriched. [127,128]. Key drivers of vitrinite formation are
Geosciences 2018, 8, 223 22 of 30
removal of cellulosic materials followed by physical compression. The result is that cellular anatomy
is likely to be destroyed, producing logs and limbs that retain their original form, but often with
significant flattening. Coalified wood typically has a vitreous or semi-vitreous luster, brittle with
conchoidal fracture (Figure 24). Microscopically, coalified wood cells are typically comminuted and
conchoidal fracture (Figure 24). Microscopically, coalified wood cells are typically comminuted and
homogenized (Figure
homogenized 25). 25).
(Figure

Figure 24. Coalified wood. (A) Coalified wood fragment from Eocene Colestin Formation, Siskiyou
Pass,Coalified
Figure 24. Oregon USAwood. (A)(B–E)
[129]. Coalified
Eocenewood fragment
Chuckanut from Eocene
Formation, WhatcomColestin
County, Formation,
WA USA. (B)Siskiyou
coalified
Pass, Oregon log, [129].
USA Bellingham
(B–E)Coal Mine.Chuckanut
Eocene (C–E) Driftwood logs in fluvial
Formation, sandstone
Whatcom at Lookout
County, Mountain.
WA USA. (B) coalified
Outer barkCoal
log, Bellingham layer Mine.
has been(C–E)
coalified, inner wood
Driftwood logshasin
decayed
fluvialand been replaced
sandstone by sandstone.
at Lookout Mountain. Outer
bark layer has been coalified, inner wood has decayed and been replaced by sandstone.
Geosciences 2018, 8, x FOR PEER REVIEW 25 of 32

Figure 25. SEM images of coalified wood. (A,B) Late Pleistocene Whidbey Formation, Swantown
Figure 25. SEM
beach, images of
Washington, USA.coalified wood. views
(A) longitudinal (A,B)showing
Late Pleistocene Whidbey
brittle fracture Formation,
of tracheids, Swantown
(B) transverse
view showing homogenized cell walls, (C,D) Oligocene Colestin Formation, Siskiyou Pass, Oregon
beach, Washington, USA. (A) longitudinal views showing brittle fracture of tracheids, (B) transverse
USA. Arrows show relict pits.
view showing homogenized cell walls, (C,D) Oligocene Colestin Formation, Siskiyou Pass, Oregon
USA. Arrows show relict pits.
5.1. Coexistence of Mummifoed and Coalified Wood
Fossil forests in the Whidbey Formation and Wilkes Formation in Washington, USA and Axel
Heiberg and Ellesmere Islands in the Canadian Arctic primarily contain both mummified wood, but
coalified wood is also present. The occurrence of both mummified and coalified wood in a single
formation, sometimes within a single stratum, is evidence that conditions of preservation are subject
to localized variation. Because the alteration of original organic matter to bituminous products
involves a reduction in volume, coalified specimens are typically compressed. At the Eocene fossil
Geosciences 2018, 8, 223 23 of 30

5.1. Coexistence of Mummifoed and Coalified Wood


Fossil forests in the Whidbey Formation and Wilkes Formation in Washington, USA and Axel
Heiberg and Ellesmere Islands in the Canadian Arctic primarily contain both mummified wood, but
coalified wood is also present. The occurrence of both mummified and coalified wood in a single
formation, sometimes within a single stratum, is evidence that conditions of preservation are subject to
localized variation. Because the alteration of original organic matter to bituminous products involves a
reduction in volume, coalified specimens are typically compressed. At the Eocene fossil forest at Axel
Heiberg Island, Canadian Arctic, coalified wood is estimated to have been compressed at a ratio of
~6:1 relative to the original wood [44].

5.2. Accessory Minerals in Mummified and Coalified Wood


The original preface of this paper is a description of fossil woods that are composed of original
organic matter or its alteration products. However, nature seldom obeys simple rules, and it is not
unusual for non-silicified wood to contain small amounts of inorganic minerals that were introduced
during diagenesis. This phenomenon usually involves the precipitation of microcrystalline minerals
within empty cell lumen. Figure 26 shows Eocene wood from Yuba River, California. Cell walls have
been fused to produce a homogenous carbon layer, suggestive of charring. Individual cell lumen
contain a variety of different diagenetic minerals, including apatite, pyrite, and gypsum. Siderite
(FeCO3 ) has previously been reported as a major component of Neogene coalified wood in Alaska,
USA [130]. Cretaceous wood from New Mexico, USA, has been reported to contain quartz, apatite,
and calcite [131]; small amounts of pyrite, siderite, and clay minerals occur in coalified Miocene tree
stumps
Geosciencesat2018,
the 8,Bilna
x FORMine, Czech Republic [132].
PEER REVIEW 26 of 32

Figure 26. Cont.


Geosciences 2018, 8, 223 24 of 30

Figure 26. SEM


Figure 26. SEM images
images ofof accessory
accessory minerals
minerals inin Eocene
Eocene charred
charred wood
wood from from Yuba
Yuba River,
River, California,
California,
USA.
USA. (A) Transverse view
(A) Transverse view shows
shows homogenized
homogenized cell cell walls,
walls, with
with diagenetic
diagenetic apatite
apatite crystals
crystals dispersed
dispersed
over
over inner
inner surfaces
surfaces of
of some
some cells.
cells. (B)
(B) Longitudinal
Longitudinal view
view of
of two
two tracheids,
tracheids, showing
showing abundant
abundant apatite.
apatite.
Apatite does not occur in intracellular spaces. (C) Apatite crystals are attached to surfaces of cell wall
Apatite does not occur in intracellular spaces. (C) Apatite crystals are attached to surfaces of cell wall
separating
separating two
two adjacent
adjacent tracheids.
tracheids. (D)
(D) Tabular
Tabular gypsum
gypsum crystals
crystals occur
occur on on outer
outer surfaces
surfaces of
of the
the wood,
wood,
apparently representing aa very
apparently representing very late
late stage
stage precipitation
precipitation event.
event. (E,F)
(E,F) Pyrite
Pyrite crystals
crystals are present in
are present in some
some
cell lumen. The
cell lumen. The presence
presence of of gypsum
gypsum (CaSO
(CaSO44··2H
2H22O)
O) and
and pyrite
pyrite (FeS
(FeS22)) are
are evidence
evidence ofof the
the availability
availability of
of
dissolved sulfur in
dissolved sulfur in pore
pore fluids
fluids during
during diagenesis.
diagenesis.

Funding: This research received no external funding.

Acknowledgments:
Funding: This research The author no
received thanks thefunding.
external following people for generously providing photographs: Brent
Alloway, Franco Bianchi,
Acknowledgments: The Tamás
authorElter, Alberto
thanks the Garcia,
followingDavid Greenwood,
people Hope Jahren,
for generously William
providing Hagopian,
photographs:
Benjamin Hook,Franco
Brent Alloway, SándorBianchi,
Kovács,Tamás
Edoardo Martinetto,
Elter, Mark Orsen,
Alberto Garcia, Marta Pinzagli,
David Greenwood, Hope Natalia Rybcynski,
Jahren, Eli Sagor,
William Hagopian,
Benjamin
Brian Hook,Chris
Schaller, Sándor Kovács,Fossil
Williams. Edoardo
wood Martinetto,
specimens Mark
wereOrsen, Marta Pinzagli,
contributed by DavidNatalia
Lawler,Rybcynski, Eli Sagor,
Jim Meissner, and
Brian Schaller, Chris Williams. Fossil wood specimens were contributed by David Lawler, Jim Meissner, and
Alex Wolfe. Charles Wampler and Dan Carnevale provided technical support for the SEM used in this research.
Alex Wolfe. Charles Wampler and Dan Carnevale provided technical support for the SEM used in this research.
Conflicts of Interest:
Conflicts of interest: The
The author
author declares
declares no
no conflicts
conflicts of
of interest.
interest.

References
References
1. Rocky Mountain Tree Ring Research. Available online: www.rmtrr.org/oldlist.htm (accessed on
18 May 2018).
2. Grant, M.C. The trembling giant. Available online: www.discovermagazine.com/1993/oct/
thetremblingian285 (accessed on 18 May 2018).
3. Mustoe, G.E. Wood petrifaction: A new view of permineralization and replacement. Geosciences 2017, 7, 1–17.
[CrossRef]
4. Mustoe, G.E. Mineralogy of non-silicified wood. Geosciences 2018, 8, 1–32. [CrossRef]
5. Akahane, H.; Furuno, T.; Miyajima, H.; Yoshikawa, T.; Yamamoto, S. Rapid wood silicification in hot spring
water: An explanation of silicification of wood during the Earth’s history. Sediment. Geol. 2004, 169, 219–228.
[CrossRef]
6. Hellawell, J.; Ballhaus, C.; Gee, C.T.; Mustoe, G.E.; Nagel, T.J.; Wirth, R.; Rethemeyer, J.; Tomaschek, F.;
Geisler, T.; Greef, K.; Mansfeldt, T. Incipient silicification of recent conifer wood at a Yellowstone hot spring.
Geochim. Cosmochim. Acta 2014, 149, 79–87. [CrossRef]
7. Channing, A.; Edwards, D. Experimental taphonmy: Silicification of plants in Yellowstone hot-spring
environments. Trans. R. Soc. Edinb. 2003, 94, 503–521. [CrossRef]
8. Ericksen, K.E.L.; Blanchette, R.A.; Ander, P. Microbial and Enzymatic Degradation of Wood and Wood Components;
Springer Verlag: Berlin, Germany, 1990; 407p.
9. Passialis, C.N. Physico-chemical characteristic of waterlogged archaeological wood. Holzforschung 1997, 51,
111–113. [CrossRef]
10. Fengel, D. Ageing and fossilization of wood and its components. Wood Sci. Technol. 1991, 25, 153–177.
[CrossRef]
Geosciences 2018, 8, 223 25 of 30

11. Kim, Y.S.; Sing, A. Micromorphological characteristics of wood biodegradation in wet environments: A
review. IAWA J. 2000, 21, 135–155. [CrossRef]
12. Klusek, M.; Pawelczyk, S. Stable carbon isotope analysis of subfossil wood from Australian Alps.
Geochronometrica 2014, 41, 400–408.
13. Barghoorn, E. Degradation of plant tissues in organic sediments. J. Sediment. Petrogrol. 1952, 2, 34–41.
14. Blanchette, R.A.; Cease, K.R.; Abad, A.R.; Burnes, T.A. Ultrastructural characterization of wood from Tertiary
fossil forests in the Canadian Arctic. Can. J. Bot. 1991, 69, 560–568. [CrossRef]
15. Jahren, A.H.; Sternberg, L.S.L. Humidity evidence for the middle Eocene Arctic rain forest. Geology 2003, 31,
463–466. [CrossRef]
16. Richter, S.L.; Johnson, A.H.; Dranoff, M.M.; LePage, B.A.; Williams, C.J. Oxygen isotope ratios in fossil wood
cellulose: Isotopic composition of Eocene-to-Holocene-aged cellulose. Geochim. Cosochim. Acta 2008, 72,
2744–2753. [CrossRef]
17. Jahren, A.H.; Sternberg, L.S.L. Annual patterns within tree rings of the Arctic middle Eocene. (ca. 45 Ma):
Isotopic evidence of precipitation, relative humidity, and deciduousness. Geology 2008, 36, 99–102. [CrossRef]
18. Eberle, J.J.; Greenwood, D.R. Life at the top of the greenhouse Eocene world—A review of the Eocene floara
and vertebrate fauna from Canada’s High Arctic. Geol. Soc. Am. Bull. 2012, 124, 3–23. [CrossRef]
19. Taggart, R.E.; Cross, A.T. Global greenhouse to icehouse and back again: The origin and future of the Boreal
Forest biome. Glob. Planet. Chang. 2009, 65, 115–121. [CrossRef]
20. Greely, A.W. Report of Sergeant D.L. Brainard on a petrified forests discovered May 20, 1883, near Cape Baird
81◦ 30’ N, 64◦ W. In Three Years of Arctic Service, Volume 2, Appendix 14; Charles Scribner’s Sons: New York
City, NY, USA, 1886; pp. 419–420.
21. Francis, J.E. A 50-million year old fossil forest from Strathcona Fjord, Ellesmere Island, Arctic Canada. Arctic
1988, 41, 314–318. [CrossRef]
22. Williams, C.J.; Johnson, A.H.; LePage, B.; Vann, D.R.; Sweda, T. Reconstruction of Tertiary forests. II.
Structure, biomass, and productivity of the Eocene floodplain forests in the Canadian Arctic. Paleobiology
2003, 29, 271–292. [CrossRef]
23. Dolezych, M.; Estrada, S. A fossil wood of Taxodium vandenburghi in Palaeogene sediment of Ellesmere Island
(Nunavut, Canada). Z. Dt. Ges. Geowiss. 2012, 163, 283–292.
24. Mitchell, W.T.; Rybcyznski, N.; Schröeder-Adama, C.; Hamilton, P.N.; Smith, R.; Douglas, M. Stratigraphic
and paleoenvironemtnal reconstruction of amid-Pliocene fossil site in the High Arctic (Ellesmere Island,
Nunavut): Evidence of an ancient peatland with beaver activity. Arctic 2016, 68, 185–204. [CrossRef]
25. Davies, N.S.; Gosse, J.C.; Rybczynski, N. Cross-bedded woody debris from a Pliocene forester rivers stem in
the High Arctic Beaufort Formation, Meighen Island, Canada. J. Sediment. Res. 2014, 84, 19–25. [CrossRef]
26. Jahren, A.H. The Arctic forest of the middle Eocene. Annu. Rev. Earth Planet. Sci. 2007, 35, 509–540.
[CrossRef]
27. Basinger, J.F.; McIver, E.E.; LePage, B.A. The fossil forests of Axel Heiberg Island. Musk-Ox 1988, 36, 50–55.
28. Basinger, J.F. The fossil forests of the Buchanan Lake Formation (early Tertiary), Axel Heiberg Island,
Canadian Arctic Archipelago: Preliminary floristics and paleoclimate. In Tertiary Fossil Forests of the Geodetic
Hills, Axel Heiberg Island, Arctic Archipelago; Christie, R.L., McMillan, N.J., Eds.; Geological Survey of Canada
Bulletin 403: Ottawa, ON, Canada, 1991; pp. 39–65.
29. Francis, J.E. Polar fossil forests. Geol. Today 1990, 6, 92–95. [CrossRef]
30. Francis, J.E. The dynamics of polar fossil forests: Tertiary fossil forests of Axel Heiberg Island, Canadian
Arctic Archipelago. In Tertiary Fossil Forests of the Geodetic Hills, Axel Heiberg Island, Arctic Archipelago;
Christie, R.L., McMillan, N.J., Eds.; Geological Survey of Canada: Ottawa, ON, Canada, 1991; pp. 29–38.
31. Christie, R.L.; McMillan, N.J. Tertiary Fossil Forests of the Geodetic Hills, Axel Heiberg Island, Arctic Archipelago;
Geological Survey of Canada Bull. 403: Ottawa, ON, Canada, 1991; pp. 1–227.
32. Greenwood, D.R.; Basinger, J.F. Stratigraphy and floristics 0f peat-coal layers in Eocene swamp forests from
Axel Heiberg Island, Canadian Arctic Archipelago. Can. J. Earth Sci. 1993, 30, 1914–1923. [CrossRef]
33. Greenwood, D.R.; Basinger, J.F. The paleoecology of high-latitude Eocene swamp forests from Axel Heiberg
Island, Canadian High Arctic. Rev. Palaeobot. Palynol. 1994, 81, 83–97. [CrossRef]
Christie, R.L., McMillan, N.J., Eds.; Geological Survey of Canada: Ottawa, ON, Canada, 1991; pp. 29–38.
31. Christie, R.L.; McMillan, N.J. Tertiary Fossil Forests of the Geodetic Hills, Axel Heiberg Island, Arctic
Archipelago; Geological Survey of Canada Bull. 403: Ottawa, ON, Canada, 1991; pp. 1–227.
32. Greenwood, D.R.; Basinger, J.F. Stratigraphy and floristics 0f peat-coal layers in Eocene swamp forests
from Axel Heiberg Island, Canadian Arctic Archipelago. Can. J. Earth Sci. 1993, 30, 1914–1923.
Geosciences 2018, 8, 223 26 of 30
33. Greenwood, D.R.; Basinger, J.F. The paleoecology of high-latitude Eocene swamp forests from Axel
Heiberg Island, Canadian High Arctic. Rev. Palaeobot. Palynol. 1994, 81, 83–97.
34. Irving,
34. E.; Wynne,
Irving, E.; Wynne, P.J. The paleoaltitude
P.J. The paleoaltitude of the Eocene
of the Eocene fossil forests
fossil of the
forests Geodetic
of the Hills,
Geodetic AxelAxel
Hills, Heiberg
Heiberg
Island, Arctic
Island, ArcticArchiplago.
Archiplago.InInTertiary Fossil Forests
Tertiary Fossil Forestsof of thethe Geodetic
Geodetic Hills,Hills, Axel Heiberg
Axel Heiberg Island,Archipelago;
Island, Arctic Arctic
Archipelago;
Christie,Christie, R.L., McMillan,
R.L., McMillan, N.J., Eds.;N.J., Eds.; Geological
Geological Survey ofSurvey CanadaofBull. Canada403: Bull.
Ottawa,403:ON,
Ottawa,
Canada,ON,1991;
Canada, 1991;
pp. 209–211. pp. 209–211
35. Basinger,
35. Basinger,J.F.; J.F.;
Greenwood,
Greenwood, D.G.; Sweda,
D.G.; Sweda,T. Early Tertiary
T. Early vegetation
Tertiary vegetation of Arctic Canada
of Arctic and and
Canada its relevance to to
its relevance
paleoclimatic
paleoclimatic interpretation.
interpretation. In Cenozoic
In Cenozoic Plants
PlantsandandClimates
Climatesofofthe theHigh
HighArctic
Arctic v.
v. 27; Boulter, M.C.,
27; Boulter, M.C.,Fisher,
Fisher,H.C.,
H.C.,Eds.;
Eds.;NATO
NATO ASI ASI Series
Series I. Springer-Verlag:
I; Springer-Verlag: Berlin,
Berlin, Germany,
Germany, 1994;
1994; pp.pp. 175–198.
175–198.
36. Kumagai, H.; Sweda, T.; Hayashi, K.; Kojima, S.;
36. Kumagai, H.; Sweda, T.; Hayashi, K.; Kojima, S.; Basinger, J.F.; Shibuya, Basinger, J.F.; Shibuya, M.; M.;
Fukaoa, Y. Growth
Fukaoa, Y. Growthringring
analysis
analysisof of early
earlyTertiary conifer woods
Tertiary conifer woodsfrom from the Canadian
the Canadian High and
High Arctic Arctic and its paleoclimatic
its paleoclimatic interpretation.
interpretation.
Palaeogeogr.Palaeogeogr.
Palaeoclimatol. Palaeoclimatol. Palaeoecol.
Palaeoecol. 1995, 1995, 116,[CrossRef]
116, 247–262. 247–262.
37. Jahren,
37. A.H.;A.H.;
Jahren, Byrne, M.C.;M.C.;
Byrne, Graham,Graham,H.V.; H.V.;
Sternberg,
Sternberg,L.S.L.;L.S.L.;
Summons, Summons, R.E. The
R.E.environmental
The environmental water ofwater
the of
middle EoceneEocene
the middle Arctic:Arctic:
Evidence from from
Evidence δD, δδD, 180 and
δ18 0 δ1
and 3C within
δ13 C within specific compounds.
specific compounds. Palaeogeogr.
Palaeogeogr.
Palaeoclimatol.
Palaeoclimatol. Palaeoecol. 2009,2009,
Palaeoecol. 271, 271,
96–103.
96–103. [CrossRef]
38. Jahren,
38. A.H.;
Jahren, LePage,
A.H.; LePage,B.A.; Werts,
B.A.; Werts,S.P. Methanogenesis
S.P. Methanogenesis ininEocene
EoceneArcticArcticsoils
soilsinferred
inferred from
from (EҰ13C
(E 13C of of tree
tree fossil
fossil carbonates.
carbonates. Palaeogeogr.
Palaeogeogr. Palaeoclimatol.
Palaeoclimatol. Palaeoecol.
Palaeoecol. 2004,2004, 214,
214, 347–358.
347–358. [CrossRef]
39. Williams,
39. Williams, C.J.;C.J.;
Mendell,
Mendell,E.K.;E.K.;
Murphy,
Murphy, J.; Court, W.M.;
J.; Court, Johnson,
W.M.; Johnson,A.H.; Richter,
A.H.; S.L.S.L.
Richter, Paleoenvironmental
Paleoenvironmental
reconstruction
reconstruction of of a aMiocene
Miocene forest fromthe
forest from thewestern
western Canadian
Canadian Arctic.Arctic. Palaeogeogr.
Palaeogeogr. Palaeoclimatol.
Palaeoclimatol. Palaeoecol.
Palaeoecol.
2008, 261,2008, 261, 160–176.
160–176. [CrossRef]
40. LePage,
40. LePage,B.A.;B.A.;
Basinger,
Basinger,J.F. J.F.
Early Tertiary
Early TertiaryLarix from
Larix from thethe
Buchanan
Buchanan LakeLakeFormation,
Formation, Canadian
Canadian Arctic, and
Arctic, and a
a consideration
consideration of ofthethe
phytogeography
phytogeography of of
thethegenus.
genus. Geol.
Geol.Surv.
Surv.Can.Can.Bull. 1991,
Bull. 403,
1991, 67–82.
403, 67–82.
41. LePage,
41. LePage,B.A.; Basinger,
B.A.; Basinger, J.F.J.F.
AA newnewspecies
speciesofofLarixLarix (Pinaceae)
(Pinaceae) from from the theearly
earlyTertiary
Tertiaryof of Axel
Axel Heiberg
Heiberg Island,
Island, Arctic
Arctic Canada.
Canada. Rev.Rev. Palaeobot.
Palaeobot. Palynol.
Palynol. 1991,1991,
70, 70, 89–111.
89–111. [CrossRef]
42. Harrington,
42. Harrington, G.J.;G.J.;
Eberle, J.; LePage,
Eberle, J.; LePage,B.A.;B.A.;
Dawson,Dawson, M.; Hutchison,
M.; Hutchison, J.H. J.H.
Arctic plantplant
Arctic diversity in the
diversity in Early
the Early
Eocene
Eocenegreenhouse.
greenhouse. Proc.Proc.
R. Soc. B Biol.
R. Soc. Sci. 2010,
B Biol. 279, 1515–1521.
Sci. 2010, 279, 1515–1521. [CrossRef] [PubMed]
43. Kaelin,
43. P.E.;P.E.;
Kaelin, Hugget,
Hugget, W.H.; Anderson,
W.H.; Anderson, K.B.K.B.Comparison
Comparison of of
vitrified
vitrified andandunvitrified
unvitrifiedEocene
Eocenewoody
woodytissue
tissue by
by TMAH
TMAHthermochemolysis—Implications
thermochemolysis—Implications for for the early stages
the early stages of offormation
formationof ofvirtrinite.
virtrinite.Geochem.
Geochem.Trans.
Trans.2006,
2006,7,7,1–12.
1–12.[CrossRef] [PubMed]
44. Fyles,
44. J.G.; J.G.;
Fyles, Hills,Hills,
L.V.;L.V.;
Matthews,
Matthews, J.V., J.V.,
Jr.; Barendregt,
Jr.; Barendregt, R.; Baker,
R.; Baker,J.; Irving, E.; Jette,
J.; Irving, H. Ballast
E.; Jette, Brook
H. Ballast and and
Brook
Beaufort Formations (Late Tertiary) on northern Banks Island, Arctic Canada. Quat. Int. 1994, 22/23, 141–171.
Beaufort Formations (Late Tertiary) on northern Banks Island, Arctic Canada. Quat. Int. 1994, 22/23, 141–171.
45. Mendell, E. Using fossil trees to estimate paleoclimate of Banks Island, Arctic Canada and the effects of
[CrossRef]
modern
45. climate
Mendell, changefossil
E. Using on the Arctic.
trees In Proceedings
to estimate paleoclimate of the of19th
Banks Annual
Island,Keck Symposium
Arctic Canada and in Geology,
the effects of
Amherst,
modern MA, USA, change
climate 20–23 April
on the2006;Arctic.pp. In14–19. Available
Proceedings of online:
the 19thhttp:/keck.wooster.edu/publications
Annual Keck Symposium in Geology,
(accessed
Amherst, on 18 MA,April 2018).
USA, 20–23 April 2006; pp. 14–19. Available online: http:/keck.wooster.edu/publications
46. Hills, L.V. Beaufort Formation,
(accessed on 18 April 2018). Northwestern Banks Island, District of Franklin; Paper 69-1A; Geological
Survey
46. Hills,of L.V.
Canada Bull.Formation,
Beaufort 403: Ottawa, ON, Canada,
Northwestern Banks 1969; pp. District
Island, 204–207. of Franklin; Paper 69-1A; Geological Survey
47. Hills, L.V. The
of Canada Bull.stratigraphy,
403: Ottawa,sedimentology,
ON, Canada, 1969; andpp.paleobotany
204–207. of the Beaufort Formation, Arctic
Archipelago, Canada. In Proceedings of the Workshop on Circum-Arctic
47. Hills, L.V. The stratigraphy, sedimentology, and paleobotany of the Late Formation,
Beaufort Tertiary/Early Pleistocene
Arctic Archipelago,
Stratigraphy and Environments, Denver, CO, USA, 15–17 October 1987; pp. 1–3.
Canada. In Proceedings of the Workshop on Circum-Arctic Late Tertiary/Early Pleistocene Stratigraphy and
48. Hills, L.V.; Ogilvie,Denver,
Environments, R.T. PiceaCO,banksii n. sp.,October
USA, 15–17 Beaufort Formation
1987; pp. 1–3. (Tertiary), northwestern Banks Island,
Arctic
48. Canada.
Hills, Can. J. Bot.
L.V.; Ogilvie, R.T.1970,
Picea48, 457–464.
banksii n. sp., Beaufort Formation (Tertiary), northwestern Banks Island, Arctic
Canada. Can. J. Bot. 1970, 48, 457–464. [CrossRef]
49. Hills, L.V.; Klovan, J.E.; Sweet, A.R. Juglans eocinera n. sp., Beaufort Formation (Tertiary), southwestern Banks
Island, Arctic Canada. Can. J. Bot. 1974, 52, 65–90. [CrossRef]
50. Hook, B.A. Paleoclimatology of the Paleocene/Eocene Using Kimberlite-Hosted Mummified Wood from
the Canadian Subarctic. Ph.D. Thesis, Department of Earth Sciences, University of Toronto, Toronto, ON,
Canada, 2014. 165p.
51. Wolfe, A.P.; Csank, A.Z.; Reyes, A.V.; McKellar, R.C.; Tappert, R. Pristine Early Eocene wood buried deeply
in kimberlite from northern Canada. PLoS ONE 2012, 17, e45537. [CrossRef] [PubMed]
52. Hook, B.A.; Halfar, J.; Gedalof, Z.; Bollman, J.; Schulze, D.J. Stable isotope paleoclimatology from the earliest
Eocene using kimberlite-hosted mummified wood from the Canadian Subarctic. Biogeosciences 2015, 12,
5899–5914. [CrossRef]
Geosciences 2018, 8, 223 27 of 30

53. Hansel, A.K.; Johnson, W.H. Wedron and Mason Groups: Lithostratigraphic Reclassification of Deposits of the
Wisconsin Episode, Lake Michigan Lobe Area; Illinois State Geological Survey Bulletin 104: Champaign, IL, USA,
1996; 116p.
54. Mossa, J.; Autin, W.J. Quaternary Geomorphology and Stratigraphy of the Florida Parishes, Southeastern Louisiana:
Friends of the Pleistocene, Field Trip Guide Book; South-Central Cell: Baton Rouge, LA, USA, 1986; pp. 76–94.
55. Yancey, T.E.; Mustoe, G.E.; Leopold, E.B.; Heizler, M.T. Mudflow disturbance in latest Miocene forests in
Lewis County, Washington. PALAIOS 2013, 28, 343–358. [CrossRef]
56. Mustoe, G.E.; Leopold, E.B. Paleobotanical evidence for the post-Miocene uplift of the Cascade Range. Can. J.
Earth Sci. 2014, 51, 809–824. [CrossRef]
57. Mickelson, D.M.; Hooyer, T.S.; Soch, B.J.; Winguth, C. Late Glacial ice advances and vegetation changes
in east-central Wisonsin. In Late Glacial History of East-Cenral Wisconson; Survey Open-File Report 2007-1;
Hooyer, T.S., Ed.; Wisconsin Geological and Natural History: Madison, WI, USA, 2007; pp. 72–87.
58. Black, R.F. Glacial Geology of the Two Creeks Forest Bed, Valderan Type Locality, and Northern Kettle Moraine State
Forest (Information Circular 13); Geological and Natural History Survey, University of Wisconsin: Madison,
WI, USA, 1970; pp. 21–30.
59. Goldthwait, J.W. The abandoned shorelines of eastern Wisconsin. Wis. Geol. Nat. Hist. Surv. Bull. 1907, 17,
1–134.
60. Wilson, L.R. The Two Creeks forest bed, Manitowoc County, Wisconsin. Trans. Wis. Acad. Sci. Arts Lett. 1932,
27, 31–46.
61. Wilson, L.R. Fossil studies of the Two Creeks forest bed, Manitowoc County, Wisconsin. Bull. Torrey Bot. Club
1936, 63, 317–325. [CrossRef]
62. Broecker, W.S.; Farrand, W.R. Radiocarbon age of the Two Creeks Forest bed, Wisconsin. Geol. Soc. Am. Bull.
1963, 74, 647–649. [CrossRef]
63. Kaiser, K.F. Two Creeks interstade dated through dendrochronology and AMS. Quat. Res. 1994, 32, 288–298.
[CrossRef]
64. Panyushkina, I.P.; Leavitt, S.W. Ancient boreal forests under the environmental instability of the glacial to
post-glacial transition in the Great Lakes region (14,000–11,000 years BP). Can. J. For. Res. 2013, 3, 1032–1039.
[CrossRef]
65. Panyshukina, I.P.; Leavitt, S.W.; Mode, W.N. A 1400-Year Bølling-Allerød Tree-Ring Record from the U.S.
Great Lakes Region. Tree-Ring Res. 2017, 73, 102–112. [CrossRef]
66. Black, R.F.; Geology of the Ice Age Natural Science Reserve of Wisconsin. National Park Service Scientific
Monograph No. 2, Chapter 2: Two Creeks Forest Bed 1974. Available online: https://www.nps.gov/
parkhistory/online_books/science/2/chap2.htm (accessed on 21 April 2018).
67. Panyushkina, I.P.; Leavitt, S.W.; Schneider, A.F.; Thompson, T.A.; Lange, T. Environment and paleoecology
of a 12 ka mid-North American Younger Dryas forest chronicled in tree rings. Quat. Res. 2008, 70, 433–441.
[CrossRef]
68. Panyushkina, I.P.; Leavitt, S.W. Tapping ancient tree-ring archives in the U.S. Great Lakes region. Eos Trans.
Am. Geophys. Union 2010, 91, 489–490. [CrossRef]
69. Easterbrook, D.J. Stratigraphy and chronology of early to late Pleistocene glacial and interglacial sediments in
the Puget Lowland, Washingon. In Geologic Field Trips in the Pacific Northwest, Proceedings of the 1994 Geological
Society of America Annual Meeting, Seattle, WA, USA, 24–27 October 1994; Swanson, D.A., Haugerud, R.A.,
Eds.; Dept. of Geological Sciences, Univ. of Washington: Seattle, WA, USA, 1994; Volume 1, pp. IJ1–IJ38.
70. Hansen, H.P.; Mackin, J.H. A pre-Wisconsin forest succession in the Puget Lowland, Washington. Am. J. Sci.
1949, 247, 833–855. [CrossRef]
71. Easterbrook, D.J.; Crandell, D.R.; Leopold, E.B. Pre-Olympia stratigraphy and chronology in the central
Puget Lowland, Washington. Geol. Soc. Am. Bull. 1967, 78, 13–20. [CrossRef]
72. Stoffel, K.L. Stratigraphy of pre-Vashon Quaternary sediments applied to the evolution of a proposed major
tectonic structure in Island County, Washington. In Washington Division of Geology and Earth Resources, Open
File Report OFR 80-0; Washington Division of Geology and Earth Resources: Olympia, WA, USA, 1980; 161p.
73. Lechien, V.; Rodriguez, C.; Onenga, M.; Hiligsmann, S.; Rulmont, A.; Thonart, P. Physiochemical and
biochemical characterization of non-biodegradable cellulose in Miocene gymnosperm wood from the
Entre-Sambre-et-Meuse, southern Belgium. Org. Chem. 2006, 37, 1465–1476.
Geosciences 2018, 8, 223 28 of 30

74. Simpson, I.M.; West, R.G. On the stratigraphy and palaeobotany of a late-Pleistocene organic deposit at
Chelford, Cheshire. New Phytol. 1958, 57, 239–250. [CrossRef]
75. Klusik, M.A.; Pawelczyk, S.R. Stable carbon isotope analysis of subfossil wood from the Austrian Alps.
Geochronometria 2014, 41, 400–408.
76. Dorofeyev, P.I. Oligocene flora of the Dunayevsky Crag on the Tym River in western Siberia. Dokl. Akademii
Nauk SSSR 1960, 132, 659–661.
77. Gnibidenko, Z.N.; Semakov, N.N. Paleomagnetism of boundary Oligocene-Miocene deposits in the
Kompasski Bor tract on the Tym River (western Siberia). Izvestia, Phys. Solid Earth 2009, 45, 70–79. [CrossRef]
78. Kázmér, M. The Miocene Bükkábrány Fossil Forest in Hungary—Field observations and project outline. In
125th Anniversary of the Department of Palaeontology at Budapest University, a Jubilee Volume; Galácz, A, Ed..
Hantkeniana 2008, 6, 220–244.
79. Erdei, B.; Dolezych, M.; Hably, L. The buried Miocene forest at Bükkábrány, Hungary. Rev. Palaeobot. Palynol.
2009, 15, 69–79. [CrossRef]
80. Gryce, V.; Sakala, J. Identification of fossil trunks from Bükkábrány newly installed in the visitor centre of
the Ipolytarnóc Fossils Nature Reserve (Novhrad-Nóográd Geopark) in northern Hungary. Acta Univ. Agric.
Silvic. Mendel. Brun. 2010, 58, 117–122. [CrossRef]
81. Dolezych, M.; Van der Burgh, J. Xylotomische Untersuchungen an inkholten Hölzern aus dem
Braunkohlentagebau Berzdorf (Oberlausitz, Deutschland). Feddes Repert. 2009, 115, 397–437. [CrossRef]
82. Hámor-Vidó, M.; Hofmann, T.; Albert, L. In Situ preservation and paleonvironmental assessment of Taxodiacea
fossil trees in the Bükkalja Lignite Formation, Bükkábrány open cast mine, Hungary. Int. J. Coal Geol. 2010,
81, 203–210. [CrossRef]
83. Baldanza, A.; Sabatino, G.; Triscari, M.; De Angeles, M.C. The Dunarobba Fossil Forest (Umbria, Italy):
Mineralogical transformations evidences as possible decay effects. Per. Mineral. 2009, 78, 51–60.
84. Vassio, E.; Martinetto, E.; Dolezych, M.; Van der Burg, J. Wood anatomy of the Glyptostrobus europaeus
“whole-plant” from a Pliocene fossil forest in Italy. Rev. Palaeobot. Palynol. 2008, 151, 81–89. [CrossRef]
85. Staccioli, G.; Menchi, G.; Matteoli, U.; Seraglia, R.; Traldi, P. Chemotaxonomic onservations on some pliocenic
woods from Arno Basin and fossil forest of Dunarobba (Italy). Flora Mediterr. 1996, 6, 113–117.
86. Staccioli, G.; Bartolini, G. New biomarkers of the extinct species Taxodioxylon gypsaceum. Wood. Sci. Technol.
1997, 31, 311–315. [CrossRef]
87. Staccioli, G.; Fratini, F.; Meli, A.; Lazzeri, S. Mineralization processes in some samples from the fossil forest
of Dunarobba (Umbria, Central Italy). Wood. Sci. Technol. 2001, 35, 353–362. [CrossRef]
88. Boyce, C.K.; Hazen, R.M.; Knoll, A.H. Nondestructive, in-situ, cellular scale mapping of elemental
abundances including organic carbon in permineralized fossils. Proc. Nat. Acad. Sci. USA 2001, 98,
5970–5974. [CrossRef] [PubMed]
89. Scott, A.C.; Collinson, M.E. Non-destructive multiple approaches to interpret the preservation of plant
fossils: Implications for calcium-rich permineralizations. J. Geol. Sci. Lond. 2003, 160, 857–862. [CrossRef]
90. Martinetto, E.; Bertini, A.; Basilica, G.; Baldanza, A.; Bizzarri, R.; Cherin, M.; Gentili, S.; Pontini, M.R. The
plant record of the Dunarobba and Pietrafitta sites in the Plio-Pleistocene palaeoenvironmental context of
central Italy. Alp. Mediter. Quat. 2014, 27, 29–72.
91. Macaluso, L.; Martinetto, E.; Vigna, B.; Bertini, A.; Cilia, A.; Teodordis, V.; Kvaček, S. Palaeofloral and
stratigraphic context of a new fossil forest from the Pliocene of NW Italy. Rev. Palaeobot. Palynol. 2018, 248,
15–33. [CrossRef]
92. Ambrosetti, P.; Basilici, G.; Ciangherotti, A.D.; Codipietro, G.; Corona, E.; Esu, D.; Girotta, O.; Lo Manco, A.;
Meneghina, M.; Paganelli, A.; et al. La Foresta Fossile di Dunarobba (Terni, Unbria, Italia Centrale): Contest
lithostratigraphico, sedimentologico, palinologico, dendrochronologico e paleomalacologico. Ital. J. Quat.
Sci. 1995, 8, 465–508.
93. Martinetto, E.; Sardia, G.; Varrone, D. Magnetostratigraphy of the Stura di Lanzo fossil forest succession
(Piedmont, Italy). Riv. Ital. Paleontol. Stratigr. 2007, 113, 109–125.
94. Ferrero, E. Foresta fossile: Fruizione didattica. In La Foresta Fossile del Torrente Stura di Lanzo. I Quaderni de La
Mandria, Ente Parco Regionale La Mandria, Torino; Martinetto, E., Farina, T., Eds.; Ente di Gestione del Parco
Regionale La Mandria e dei Parchi e delle Riserve Naturali delle Valli di Lanzo: Venaria, Italy, 2005; Volume
1, pp. 44–45.
Geosciences 2018, 8, 223 29 of 30

95. Godone, F.; Baaaldo, M.; Maraga, F. Una foresta nel’alveo del F. Stura di Lanzo (TO). Rilevamentu topografici,
cartografia e GIS. Atti 8◦ Conferenza Nationale ASITA, Roma, 14–17 dic. 2004, Artesanpa sas. Varese 2004, 2,
1219–1224.
96. Chiariglione, A.; Farina, T.; Ferrero, O.; Forna, M.G.; Gattiglio, M.; Lucchesi, S.; Martinetto, E. La Foresta
Fossile dello Stura de Lanzo, Quaderni de la Mandria, Parco La Mandria 2005, 48p. Available online:
https://www.researchgate.net/profile/Edoardo_Martinetto (accessed on 19 May 2018).
97. Ferrero, E.; Magagna, A. Natural hazards and geological heritage in Earth science education projects. In
Geoethics: The Role and Responsibility of Geoscientists; Peppoloni, S., Di Capua, G., Eds.; Geological Society of
London Special Publications: London, UK, 2015; p. 419.
98. Ferrero, E.; Gimigliano, D.; Pogliano, A. Educazione ambientale in un’area protetta del territorio piemontese.
Il progetto ‘La foresta ritrovata’. In Proceedings of the Atti 3rd World Environmental Education Congress,
Turin, Italy, 2–6 October 2005; pp. 170–177.
99. Quan, C.; Fu, Q.Y.; Shu, G.L.; Li, Y.S.; Lu, L.; Liu, X.Y.; Jin, J.H. First Oligocene mummified plant Lagerstätte
at low latitudes of East Asia. Sci. China Earth Sci. 2016, 59, 445–560. [CrossRef]
100. Yamakawa, C.; Momohara, A.; Saito, T.; Nunotani, T. Composition and paleoenvironment of wetland forests
dominated by Glyptostrobus and Metasequoia in the latest Pliocene (2.6Ma) in central Japan. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 2017, 467, 191–210. [CrossRef]
101. Scott, A.C. The pre-Quaternary history of fire. Palaeogeogr. Palaoclimatol. Palaeoecol. 2000, 154, 281–329.
[CrossRef]
102. Scott, A.C.; Cripps, J.A.; Collinson, M.E.; Nicjols, G.J. The taphonomy of charcoal in a Recent heathland fire
and some implications for the interpretation of fossil charcoal deposits. Palaeogeogr. Palaeoclimatol. Palaeoecol.
2000, 164, 1–31. [CrossRef]
103. Scott, A.C. The preservation of woods in volcanic pyroclastic flows and surges. In Proceedings of the
Geological Society of America 2001 Annual Meeting, Boston, MA, USA, 5–8 November; p. 68.
104. Scott, A.C. Charcoal recognition, taphonomy, and uses in paleoenvironmenetal analysis. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 2010, 291, 11–39. [CrossRef]
105. Scott, A.C.; Damblon, F. Charcoal: Taphonomy and significance in geology, botany, and archaeology.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 2010, 291, 1–10. [CrossRef]
106. Sander, O.M.; Gee, C.T. Fossil charcoal: Techniques and applications. Rev. Palaeobot. Palynol. 1990, 63,
269–279. [CrossRef]
107. Glasspool, I.J.; Scott, A.C. Phanerozoic concentrations of atmospheric oxygen reconstructed from sedimentary
charcoal. Nat. Geosci. 2010, 3, 627–630. [CrossRef]
108. Moore, P.R.; Wallace, R. Petrified wood from the Miocene volcanic sequence of Coromandel Peninsula,
northern New Zealand. J. R. Soc. N. Z. 2000, 30, 115–130. [CrossRef]
109. Hudspeth, V.A.; Scott, A.C.; Wilson, C.J.N.; Collinson, M.E. Charring of woods by volcanic processes: An
example from the Taupo ignimbrite, New Zealand. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2010, 291, 40–51.
[CrossRef]
110. Glasspool, I.J.; Edwards, D.; Axe, L. Charcoal in the Silurian as evidence for the earliest wildfiew. Geology
2004, 32, 381–383. [CrossRef]
111. Glasspool, I.J.; Edwards, D.; Axe, L. Charcoal in the Early Devonian: A wildfire-derived
Konservate-Lagerstätte. Rev. Paleobot. Palynol. 2006, 142, 131–136. [CrossRef]
112. Prestianni, C.; Decombeix, A.L.; Thorez, J.; Fokan, D.; Gerrienne, P. Famennian charcoal of Belgium.
Palaeoecology 2010, 291, 60–71. [CrossRef]
113. Berner, R.Z.; Canfield, D.E. A new model for atmospheric oxygen over Phanerozoic time. Am. J. Sci. 1989,
289, 333–361. [CrossRef] [PubMed]
114. Berner, R.A.; Beerling, D.J.; Dudley, R.; Robinson, J.M.; Wildman, R.A., Jr. Phanerozoic atmospheric oxygen.
Annu. Rev. Earth Planet. Sci. 2003, 31, 105–134. [CrossRef]
115. Berner, R.A. The carbon and sulfur cycles and atmospheric oxygen from Middle Permian to late Triassic.
Geochim. Cosmochim. Acta 2005, 69, 3211–3217. [CrossRef]
116. Jones, T.P.; Chaloner, W.H. Fossil charcoal, its recognition and paleoatmospheric significance. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 1991, 97, 39–50. [CrossRef]
117. Tanner, L.H.; Wang, X.; Morabito, A.C. Fossil charcoal from the Middle Jurassic of the Ordos Basin, China
and its paleoatmospheric implications. Geosci. Front. 2012, 3, 493–502. [CrossRef]
Geosciences 2018, 8, 223 30 of 30

118. Belcher, C.M.; Yearsle, J.M.; Hadden, R.M.; McElwaine, J.C.; Rein, G. Baseline intrinsic flammability of
Earth’s exosystems estimated from paleoatmospheric oxygen over the past 350 million years. PNAS 2010,
107, 22448–22543. [CrossRef] [PubMed]
119. Uhl, D.; Dolezych, M.; Böhme, M. Taxodioxylon-like charcoal from the Late Miocene of western Bulgaraia.
Acta Palaeontol. 2014, 54, 101–111.
120. McPartland, L.C.; Collinson, M.E.; Scott, A.C.; Steart, D.C.; Grassineau, N.V.; Gibbins, S.J. Ferns and fires:
Experimental charring of ferns compared to wood and implications for paleobiology, paleoecology, coal
petrology, and isotope geochemistry. PALAIOS 2007, 22, 528–538. [CrossRef]
121. Byers, B.A.; Ash, S.R.; Chaney, D.; DeSoto, L. First known fire scar on a fossil free trunk provides evidence of
Late Triassic wildfire. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2014, 411, 180–187. [CrossRef]
122. Stout, S.A.; Brown, J.J.; Packman, W. Molecular aspects of the peatification and early coalification of
angiosperm and gymnosperm woods. Geochim. Cosmochim. Acta 1988, 52, 405–414. [CrossRef]
123. Rodgers, R.E.; Ramurthy, M.; Rodvelt, G.; Mullin, M. Coalbed Methane, Principles and Practices, 3rd ed.;
Oktibbeha Pub. Co.: Starkville, MS, USA, 2007; pp. 97–98.
124. Ravazzi, C.; Van der Burgh, J. Coniferous woods in the early Pleistocene brown coals of the Leffe Basin
(Lombardy, Italy). Riv. It. Paleont. Strat. 1995, 100.
125. Dolezych, M. Taxodiaceous woods in Lusatia (central Europe), including curiosities in their nomenclature
and taxonomy, with a focus on Taxodioxylon. Jpn. J. Hist. Bot. 2011, 19, 25–46.
126. Dolezych, M. A remarkable extinct wood from Lusatia (central Europe)—Juniperoxylon schneiderianum sp.
Nov. with affinity to Cupressospernum saxonicum MAI. Palaeontogr. Abt. B Palaeobot.-Paleopjhytol. 2016, 295,
5–31.
127. Nestler, K.; Deitrich, D.; Witke, K.; Röβler, R.; Marx, G. Thermogravimetric and Raman spectroscopic
investigations on different coals in comparison to dispersed anthracite found in permineralized tree fern
Psaronius sp. J. Mol. Struct. 2003, 661–662, 357–362. [CrossRef]
128. Hedges, J.I.; Cowie, G.L.; Ertel, J.R.; Barbour, R.J.; Hatcher, P.G. Degradation of carbohydrates and lignins in
buried woods. Geochim. Cosmochim. Acta 1985, 49, 700–711. [CrossRef]
129. Bestland, E. Volcanic stratigraphy of the Oligocene Colestin Formation in the Siskiyou Pass area of southern
Oregon. Or. Geol. 1987, 49, 79–86.
130. Williams, C.J.; Trostle, K.D.; Sunderlin, D. Fossil wood in coal-forming environments of the Late Paleocene
–Early Eocene Chickaloon Formation. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2010, 295, 363–375. [CrossRef]
131. Sweeney, I.J.; Chin, K.; Hower, J.C.; Budd, D.A.; Wolfe, D.G. Fossil wood from the middle Cretaceous Moreno
Hill Formation: Unique expressions of wood mineralization and implications for the processes of wood
preservation. Int. J. Coal Geol. 2009, 79, 1–17. [CrossRef]
132. Havelcová, M.; Sýkorová, I.; Bectel, A.; Mach, K.; Trejtnarová, H.; Žaloudková, M.; Matusová, P.; Blažek, J.;
Boudová, J.; Sakala, J. “Stump Horizon” Bílina Mine (Most Basin, Czech Republic)-GC-MS, optical and
electron microscopy in identification of wood of biological origin. Int. J. Coal Geol. 2013, 107, 62–77.
[CrossRef]

© 2018 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like