ArticlePDF Available

Abstract and Figures

The origins of South America’s exceptional plant diversity are poorly known from the fossil record. We report on unbiased quantitative collections of fossil floras from Laguna del Hunco (LH) and Río Pichileufú (RP) in Patagonia, Argentina. These sites represent a frost‐free humid biome in South American middle latitudes of the globally warm Eocene. At LH, from 4,303 identified specimens, we recognize 186 species of plant organs and 152 species of leaves. Adjusted for sample size, the LH flora is more diverse than comparable Eocene floras known from other continents. The RP flora shares several taxa with LH and appears to be as rich, although sampling is preliminary. The two floras were previously considered coeval. However, 40Ar/39Ar dating of three ash‐fall tuff beds in close stratigraphic association with the RP flora indicates an age of \documentclass{aastex} \usepackage{amsbsy} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{bm} \usepackage{mathrsfs} \usepackage{pifont} \usepackage{stmaryrd} \usepackage{textcomp} \usepackage{portland,xspace} \usepackage{amsmath,amsxtra} \usepackage[OT2,OT1]{fontenc} \newcommand\cyr{ \renewcommand\rmdefault{wncyr} \renewcommand\sfdefault{wncyss} \renewcommand\encodingdefault{OT2} \normalfont \selectfont} \DeclareTextFontCommand{\textcyr}{\cyr} \pagestyle{empty} \DeclareMathSizes{10}{9}{7}{6} \begin{document} \landscape $47.46\pm 0.05$ \end{document} Ma, 4.5 million years younger than LH, for which one tuff is reanalyzed here as \documentclass{aastex} \usepackage{amsbsy} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{bm} \usepackage{mathrsfs} \usepackage{pifont} \usepackage{stmaryrd} \usepackage{textcomp} \usepackage{portland,xspace} \usepackage{amsmath,amsxtra} \usepackage[OT2,OT1]{fontenc} \newcommand\cyr{ \renewcommand\rmdefault{wncyr} \renewcommand\sfdefault{wncyss} \renewcommand\encodingdefault{OT2} \normalfont \selectfont} \DeclareTextFontCommand{\textcyr}{\cyr} \pagestyle{empty} \DeclareMathSizes{10}{9}{7}{6} \begin{document} \landscape $51.91\pm 0.22$ \end{document} Ma. Thus, diverse floral associations in Patagonia evolved by the Eocene, possibly in response to global warming, and were persistent and areally extensive. This suggests extraordinary richness at low latitudes via the latitudinal diversity gradient, corroborated by published palynological data from the Eocene of Colombia.
Content may be subject to copyright.
The University of Chicago
(RFHQH3ODQW'LYHUVLW\DW/DJXQDGHO+XQFRDQG5¯R3LFKLOHXI¼3DWDJRQLD$UJHQWLQD
$XWKRUV3HWHUb:LOI.LUNb5b-RKQVRQ1b5XE«Qb&¼QHR0b(OOLRWb6PLWK%UDGOH\b6b6LQJHU
DQG0DULDb$b*DQGROIR
6RXUFH
7KH$PHULFDQ1DWXUDOLVW
9RO1R-XQHSS
3XEOLVKHGE\The University of Chicago PressIRUThe American Society of Naturalists
6WDEOH85/http://www.jstor.org/stable/10.1086/430055 .
$FFHVVHG
Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .
http://www.jstor.org/page/info/about/policies/terms.jsp
.
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of
content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms
of scholarship. For more information about JSTOR, please contact support@jstor.org.
.
The University of Chicago Press, The American Society of Naturalists, The University of Chicago are
collaborating with JSTOR to digitize, preserve and extend access to The American Naturalist.
http://www.jstor.org
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
vol. 165, no. 6 the american naturalist june 2005 !
Eocene Plant Diversity at Laguna del Hunco and
´o Pichileufu´ , Patagonia, Argentina
Peter Wilf,
1,
*
Kirk R. Johnson,
2,
N. Rube´n Cu´ neo,
3,
M. Elliot Smith,
4,§
Bradley S. Singer,
4,k
and
Maria A. Gandolfo
5,#
1. Department of Geosciences, Pennsylvania State University,
University Park, Pennsylvania 16802;
2. Department of Earth Sciences, Denver Museum of Nature and
Science, Denver, Colorado 80205;
3. Museo Paleontolo´gico Egidio Feruglio, Trelew, Chubut 9100,
Argentina;
4. Department of Geology and Geophysics, University of
Wisconsin, Madison, Wisconsin 53706;
5. L. H. Bailey Hortorium, Department of Plant Biology, Cornell
University, Ithaca, New York 14853
Submitted October 1, 2004; Accepted February 9, 2005;
Electronically published April 7, 2005
Online enhancement: appendix.
abstract: The origins of South America’s exceptional plant diver-
sity are poorly known from the fossil record. We report on unbiased
quantitative collections of fossil floras from Laguna del Hunco (LH)
and Rı´o Pichileufu´ (RP) in Patagonia, Argentina. These sites represent
a frost-free humid biome in South American middle latitudes of the
globally warm Eocene. At LH, from 4,303 identified specimens, we
recognize 186 species of plant organs and 152 species of leaves. Ad-
justed for sample size, the LH flora is more diverse than comparable
Eocene floras known from other continents. The RP flora shares
several taxa with LH and appears to be as rich, although sampling
is preliminary. The two floras were previously considered coeval.
However,
40
Ar/
39
Ar dating of three ash-fall tuff beds in close strati-
graphic association with the RP flora indicates an age of 47.46 !
Ma, 4.5 million years younger than LH, for which one tuff is0.05
reanalyzed here as Ma. Thus, diverse floral associations51.91 !0.22
in Patagonia evolved by the Eocene, possibly in response to global
warming, and were persistent and areally extensive. This suggests
*Correspondingauthor;e-mail:pwilf@psu.edu.
E-mail: kjohnson@dmns.org.
E-mail: rcuneo@mef.org.ar.
§
E-mail: msmith@geology.wisc.edu.
k
E-mail: bsinger@geology.wisc.edu.
#
E-mail: mag4@cornell.edu.
Am. Nat. 2005. Vol. 165, pp. 634–650. "2005 by The University of Chicago.
0003-0147/2005/16506-40660$15.00. All rights reserved.
extraordinary richness at low latitudes via the latitudinal diversity
gradient, corroborated by published palynological data from the Eo-
cene of Colombia.
Keywords: biodiversity, Eocene, geochronology, paleobotany, paleo-
climate, Patagonia.
South American floras today are marked by high diversity
and endemism. Most notably, Neotropical plant diversity
exceeds other tropical regions by factors of two to three
(Gentry 1988b;Davisetal.1997;PhillipsandMiller2002).
Elevated richness exists outside of the main equatorial belt,
including the Mata Atlaˆntica region of southeastern Brazil
(Mori and Boom 1981). Davis et al. (1997) designated 46
areas of the continent as centers of floral diversity and
endemism, including several extratropical locales such as
the central Chilean Mediterranean region and the Gran
Chaco of Brazil, Bolivia, Argentina, and Paraguay.
Because of a general lack of reliable, quantitative data
from plant macrofossils, remarkably little is known about
the origins, history, and geologic context of South Amer-
ican floral diversity, especially before the Pleistocene. Mod-
ern tropical weathering, vegetative cover, a paucity of in-
vestigators, and the obsolescence of historical reports
account for the scarcity of information from equatorial
latitudes for the Paleogene (Romero 1986; Mello et al.
2002; Burnham and Johnson 2004) and Neogene (Burn-
ham and Graham 1999). Most hypotheses for the Neo-
tropical region have emphasized geologically recent events.
Neogene uplift of the Andes is thought to have fostered
diversification (Van der Hammen and Hooghiemstra 2000;
Colinvaux and De Oliveira 2001), and a palynological
study indicates high floral richness during the Miocene in
Colombia (Hoorn 1994). The prevailing model for several
decades, that Pleistocene rain forest refugia catalyzed cur-
rent biodiversity as sites of allopatric speciation (Haffer
1969; Haffer and Prance 2001), has been strongly chal-
lenged (Nelson et al. 1990; Colinvaux et al. 1996, 2001;
Moritz et al. 2000; Kastner and Gon˜i 2003). Evidence for
timing of diversification in various lineages mostly comes
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
Eocene Plant Diversity in Patagonia 635
Figure 1: Locations of Laguna del Hunco, Rı´o Pichileufu´, and cities
mentioned in text, mapped onto 50-Ma positions of modern coastlines,
Mollweide Projection. Reconstruction made using the Plate Tectonic
Reconstruction Service of the Ocean Drilling Stratigraphic Network,
http://www.odsn.de/odsn/services/paleomap/paleomap.html, based on
data from Hay et al. (1999).
from molecular clock studies, constrained using the few
fossil data that are taxonomically reliable (Richardson et
al. 2001; Bremer 2002; Davis et al. 2004, 2005).
In contrast to conventional literature, two recent in-
vestigations indicate elevated plant diversity in South
America during the early Eocene, 30 million years before
significant Andean uplift, 50 million years before the Pleis-
tocene, and on a warm and humid continent without Pleis-
tocene analog refugia. First, palynological data from Co-
lombia quantify significant, apparently in situ plant
diversification in association with global warming and in-
creasing regional precipitation across the Paleocene/Eo-
cene boundary (Jaramillo and Dilcher 2000; Jaramillo
2002). Second, and antecedent to this article, is a prelim-
inary macrofloral study from Laguna del Hunco (LH) in
the northern Patagonian region of Argentina (Wilf et al.
2003; fig. 1). During the globally warm Eocene, when trop-
ical organisms reached middle and high latitudes of both
hemispheres (Estes and Hutchison 1980; Greenwood and
Wing 1995), LH was located near the southern limit of
Neotropical floral influence (Romero 1978, 1986, 1993;
Hinojosa and Villagra´n 1997). Today, LH is a desert area
with abundant exposures of fossiliferous strata. Analyses
of quantitative, stratigraphically controlled, unbiased col-
lections suggested that when adjusted for sample size, LH
is the most diverse fossil flora of Eocene age known any-
where in the world (Wilf et al. 2003). However, sample
sizes at individual quarries were small compared to North
American floras used for comparison (175–315 speci-
mens vs. typically more than 500 specimens). The report
also demonstrated that data from elsewhere in Patagonia
are needed to test the pattern of elevated floral richness
in time and space.
New results are presented here from significantly ex-
panded collections at Laguna del Hunco and initial col-
lections from Rı´o Pichileufu´ (RP; fig. 1), a rich Eocene
site about 160 km NNW of LH that has received negligible
investigation since the 1930s (Berry 1938). We test the
observation of elevated floral diversity at Laguna del
Hunco using a sample size approximately triple that of
the preliminary report (Wilf et al. 2003). We use paleo-
ecological data to ask whether the LH floras are compo-
sitionally stable through time, and we refine paleoclimatic
estimates and biome interpretation. Using the first quan-
titative paleobotanical data from Rı´o Pichileufu´, we ask
whether its flora had diversity comparable to LH. An ac-
curate geochronology for the floras, conventionally as-
sumed to be coeval, is critical for addressing several eco-
logical questions, including persistence. We analyze ashfall
tuffs intimately associated with the fossil plants at RP to
determine the first high-precision ages for the flora, and
we reanalyze a tuff from LH to resolve its age more finely.
We conclude by discussing the implications of our results
for understanding past and present South American
biodiversity.
Laguna del Hunco and Rı´o Pichileufu´Floras
Overview and Previous Work
The classic fossil floras from Laguna del Hunco (“Lake of
Reeds”) and Rı´o Pichileufu´ (“Little River”) have a long
history of investigation (Berry 1925, 1938), and they fea-
ture prominently in phytogeographic and paleoclimatic
literature about the continent (Mene´ndez 1971; Arago´n
and Romero 1984; Romero 1986, 1993; Hinojosa and Vil-
lagra´n 1997) and the Southern Hemisphere in general
(Christophel 1980; Hill 1994). However, these assemblages
are historically understudied and are mostly represented
by small museum collections made from limited numbers
of quarries collected near the surface without reliable strat-
igraphic or relative abundance data.
The two floras feature very good (RP) to outstanding
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
636 The American Naturalist
Figure 2: Stratigraphic section of the Tufolitas Laguna del Hunco, revised
and simplified from Wilf et al. (2003). Shown are the stratigraphic po-
sitions of fossil plant localities (LH-1 to LH-25), the samples yielding
radiometric ages (numbers in gray), including the reanalyzed age reported
here for the uppermost sample, and the paleomagnetic interpretation.
Asterisks indicate the four major sampling localities. Only 10 identifiable
fossils were found above the labeled densely sampled interval. Localities
19, 20, and 21 are stratigraphically isolated and not included in the figure.
See Wilf et al. (2003) for details of stratigraphy, lithology, and paleo-
magnetic and previous radioisotopic analyses.
(LH) preservation and are located at similar middle pa-
leolatitudes in the Patagonian region of Argentina (fig. 1).
Laguna del Hunco is a desert pond in Chubut Province,
located in the midst of the major fossil exposures. The
floras from LH are the major focus of this article because
of our extensive, recent collections; they are well exposed
and feature exceptional preservation equal to that of well-
known Eocene floras from North America such as Floris-
sant (MacGinitie 1953) and Republic (Wolfe and Wehr
1987). The Rı´o Pichileufu´floraislocatedneartheepon-
ymous stream in Rı´o Negro Province, 160 km northwest
of LH, and we report preliminary data based on initial
collections.
No significant uplift of the Southern Central Andes oc-
curred before the Miocene (Marshall and Salinas 1990);
Eocene Laguna del Hunco lay at a low elevation, and a
significant maritime influence moderated its climate. The
LH flora was deposited in tuffaceous mudstones and sand-
stones of the Tufolitas Laguna del Hunco, a lacustrine unit
of the middle Chubut River volcanic-pyroclastic complex
(Arago´n and Romero 1984; Arago´n and Mazzoni 1997)
and is inferred to represent lakeshore vegetation (Wilf et
al. 2003). The RP flora comes from volcanic lake beds of
the Ventana Formation and is probably similar in depo-
sitional setting to LH (fig. 1; Gonza´lez Dı´az 1979; Arago´n
and Romero 1984). The LH tuffs are well exposed over
an area of more than 25 km in diameter (Arago´n and
Mazzoni 1997), but fossiliferous horizons at Rı´o Pichileufu´
are restricted to a single small drainage (!300 m of map
distance separates the fossil quarries). Compressions of
flowers, fruits, seeds, and especially leaves are abundant
at both sites. At LH, fossil fish, insects (Fidalgo and Smith
1987; Petrulevicius and Nel 2003), caddis-fly cases (Genise
and Petrulevicius 2001), and pipoid frogs (Casamiquela
1961; Ba´ez and Trueb 1997) are present on the same bed-
ding planes as the plants. At RP, large-bodied ants, prob-
ably referable to Archimyrmex piatnitzkyi (Viana and
Haedo-Rossi 1957; Dlussky and Perfilieva 2003), co-occur
with plants, as do frogs (P. Wilf and K. R. Johnson, per-
sonal observation).
Fossil plants at Laguna del Hunco were discovered dur-
ing the 1920s (Clark 1923; Berry 1925), and occasional
new descriptions and taxonomic revisions have since been
published (Berry 1938; Frenguelli and Parodi 1941; Fren-
guelli 1943a,1943b;Traverso1964;RomeroandHickey
1976; Durango de Cabrera and Romero 1986; Gandolfo
et al. 1988; Romero et al. 1988; Gonza´ lez et al. 2002). The
total number of formally described entities referable to
surviving voucher specimens is about 44 species of leaves,
fruits, and seeds, nearly all in need of taxonomic revision:
only six species have been described or revised since 1945.
The age of the flora was originally reported as Miocene
(Berry 1925) and later, from
40
K/
40
Ar analyses of associated
volcanic rocks, was assigned ages ranging from late Pa-
leocene to middle Eocene (Archangelsky 1974; Mazzoni et
al. 1991).
The Rı´o Pichileufu´floraisthemostdiversefromCe-
nozoic South America, with more than 130 formally de-
scribed entities (Berry 1935a,1935b,1935c,1938);this
figure remains a useful starting estimate of richness. How-
ever, despite its diversity, very good to excellent preser-
vation, and accessibility from San Carlos de Bariloche (fig.
1), the flora has been little studied since Berry’s historic
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
Eocene Plant Diversity in Patagonia 637
Table 1 : Sampling and taphonomic data for the four principal quarries at Laguna del Hunco (LH)
LH-13 LH-2 LH-4 LH-6
Meter level (fig. 2) 40.30 68.18 75.63 99.28
Excavated fossiliferous volume (m
3
) 1.4 2.0 1.8 1.9
Plant fossil density (specimens/m
3
)1,008710929535
Plant specimens:
Total 1,451 1,406 1,672 995
Identifiable 1,123 (77%) 647 (46%) 1,268 (76%) 685 (69%)
Leaves, ID’d 1036 626 1260 642
Dicot leaves, ID’d 806 609 1248 530
Dicot reproductive 30 18 7 34
Conifer leaves 218 10 6 86
Conifer reproductive 57 3 0 7
Fern leaves 5 3 5 3
Dicot leaf specimens toothed (%) 38.3 56.7 76.3 47.4
Mean area of all dicot leaves, ln (mm
2
)6.78!.92 6.83 !.91 7.05 !.85 6.91 !.95
Helicopter fruits, abundance/species 0/0 0/0 1/1 4/2
Fish fossils Absent Absent Absent Articulated
Insect accumulations Present Absent Absent Present
Insect abundance
a
15% 5% 5% 76%
Interpretation: rank distance from shoreline 2 3 4 1
a
Number of body fossils in 2002 collections divided by the number from all four major localities combined.
efforts. It has been considered approximately coeval to LH
on the basis of shared plant species (Berry 1938; Petersen
1946; Arguijo and Romero 1981; Arago´n and Romero
1984; Markgraf et al. 1996; Hinojosa and Villagra´n 1997),
but until recently, neither flora had been reliably dated.
Biostratigraphic and
40
K/
40
Ar analyses of the Ventana For-
mation from sites with uncertain stratigraphic relation-
ships to the RP flora have suggested ages ranging from
late Paleocene to early Oligocene (Gonza´lez Dı´az 1979;
Rapela et al. 1988).
In November of 1999, our group completed the first
stratigraphically controlled, quantitative collection of the
LH floras (Wilf et al. 2003). From a 170-m stratigraphic
section (fig. 2), we excavated 25 quarries and collected or
systematically tallied 1,583 identifiable macrofossil speci-
mens without collection bias (totals and analyses for 1999
collections revised slightly in this article for recently iden-
tified material). Plant diversity at LH was compared to and
exceeded seven North American midlatitude assemblages
that have been collected using similar methods. Of these,
a collection of the lacustrine Republic flora from Wash-
ington, made by K. R. Johnson and colleagues from Denver
Museum of Nature and Science locality 2130 and prelim-
inarily analyzed (Passmore et al. 2002), was the most di-
verse. Three
40
Ar/
39
Ar analyses from tuffs discovered within
the LH fossiliferous sequence, coupled with six paleo-
magnetic reversals, calibrated the fossiliferous strata to ages
near 52 Ma and near the base of magnetic polarity Chron
23 (Wilf et al. 2003; fig. 2). These results date the LH flora
to the early Eocene climatic optimum, a 1–2-million-year
interval that featured the warmest sustained high temper-
atures of the Cenozoic (Zachos et al. 2001).
Biome Characterization and Biogeography
Several reconstructions of the warm early Eocene place the
RP and LH floras within tropical rain forest areas. Ac-
cording to Morley (2000), tropical rain forest usually re-
quires a minimum mean monthly temperature above
18#C, annual precipitation above 200 cm, and a dry season
with no more than 4 months below 10 cm of rainfall per
month. In early global reconstructions, tropical rain forests
cover most lowland areas up to about 50#north and south
paleolatitude (Wolfe 1985; Frakes et al. 1992). Morley
(2000) refined the mapping of past rain forest distribu-
tions, primarily from palynological data and inference
from modern climate systems, and included desert and
grassland belts centered on latitudes 30#north and south
resulting from Hadley cell circulation. His early Eocene
reconstruction shows American tropical rain forests ex-
pressed in three latitudinal bands, separated by the dry
belts. These are the Neotropical equatorial rain forest; the
Boreotropical rain forest, located at middle Northern lat-
itudes; and the Southern Megathermal rain forest at South-
ern middle latitudes, containing the Patagonian floras dis-
cussed here. The eastern margins of continents typically
receive significant rainfall from warm offshore currents
even in the desert latitudes, and Morley shows his Neo-
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
638 The American Naturalist
Table 2 : Diversity and paleoclimatic data, Laguna del Hunco (LH) flora
LH-13 LH-13 level LH-2 LH-2 level LH-4 LH-4 level LH-6 LH-6 level All
Meter level (fig. 2) 40.30 37.00, 40.30 68.18 68.18 75.63 75.63 99.28 99.28 All
Plant species:
All 81 88 75 75 64 72 74 83 186
Leaves 67 74 69 69 59 65 58 66 152
a
Dicot leaves 55 61 61 61 53 56 45 52 132
Nonleaves 14 14 6 6 5 7 16 17 34
Paleotemperature:
Species used 55 61 61 61 53 56 44 51 131
P.436 .426 .557 .557 .509 .482 .545 .569 .504
MAT (#C) 14.5 14.2 18.2 18. 2 16.7 15.9 17.8 18.6 16.6
a
j(MAT) (#C) 2.0 1.9 1.9 1.9 2.1 2.0 2.3 2.1 1.3
Paleoprecipitation:
Species used 51 57 59 59 48 51 43 49 119
MlnA, ln (mm
2
)7.177.15 7.097.097.387.407.247.207.24
MAP (cm) 110 109 104.8 105 123 125 114 112 114
a
SE !(MAP) (cm) 47.3 46.9 45.2 45.2 53.3 53.9 49.2 48.2 49.1
SE "(MAP) (cm) "33.1 "32.8 "31.6 "31.6 "37.2 "37.6 "34.3 "33.6 "34.3
Note: Levels are the minor localities grouped with major localities at the same stratigraphic levels (see fig. 2; LH-13 was grouped with LH-14, 15, and 23,
all 3.3 m below LH-13). Dicot leaf species used for mean annual temperature (MAT) and precipitation (MAP) estimates differ slightly because some species
suitable for leaf-margin analysis were too fragmentary for reliable measurement of leaf area. Gymnostoma was not used for climate estimates because of
unusual leaf morphology; sampling error, used as a minimum error estimate for MAT when 12.0#C; when , !2#C is thej(MAT) pbinomial j(MAT) !2#C
recommended minimum error estimate (Wilf 1997). of dicot leaf species with untoothed margins; natural log of species leafPpproportion MlnA pmean
areas (Wilf et al. 1998). error.SE pstandard
a
We consider these values to be the best single estimates for the flora.
tropical and Southern Megathermal Patagonian tropical
rain forests connected by a narrow strip of moist climatic
conditions along the Atlantic Coast. Morley did not pre-
sent direct evidence for Eocene tropical rain forests in
middle latitude South America, save for the presence of
diverse fossil floras with tropical affinities such as Rı´o Pi-
chileufu´. A dry or semiarid biome north of the LH and
RP floras is supported by evaporite occurrences (Ziegler
et al. 2003) and diverse evidence from the early Eocene
Gran Salitral Formation (Melchor et al. 2002).
Previous paleoclimatic analyses based specifically on the
paleobotany of the LH and RP floras have suggested some
type of moist to seasonally dry tropical or subtropical cli-
mate (Berry 1925, 1938; Arago´ n and Romero 1984; Rom-
ero 1986; Markgraf et al. 1996; Wilf et al. 2003). However,
none has indicated a tropical rain forest biome matching
the reconstructions discussed above.
Living relatives of taxa from the LH and RP floras are
dispersed widely, for the most part in tropical and tem-
perate areas of South America and Australasia; current
biogeographic knowledge of individual taxa is summarized
in the appendix in the online edition of the American
Naturalist. Gondwanic affinities of the floras are much
better understood than Neotropical relationships (e.g., Hill
and Carpenter 1991; Hill 1994). Although diverse tropical
elements are present in the fossil floras, including many
lineages that are diverse and abundant in today’s Neo-
tropics, few of these are confirmed to be Neotropical en-
demics (Durango de Cabrera and Romero 1986). Never-
theless, the proximity of the American tropics and the
presumed existence of direct interchange corridors (i.e.,
the Atlantic coast) support the argument for Neotropical
interchange with Patagonia. The warm Paleogene climate
presumably furthered the southward dispersal of Neo-
tropical elements, in concert with poleward migrations of
numerous tropical elements worldwide (e.g., Pole and
MacPhail 1996; Harrington 2004). The extant Chilean and
Argentine floras contain significant numbers of genera
with disjunct distributions in the Neotropics that may be
Paleogene relicts (Davis et al. 1997; Villagra´n and Hinojosa
1997).
The combined presence of taxa with tropical and Gond-
wanic/Antarctic affinities at LH, RP, and other Paleogene
floras in Patagonia led Romero (1978, 1986) to coin the
descriptive term “mixed paleoflora” for these associations,
which he considered to be stable vegetational units without
direct modern analog. Because both the LH and RP ma-
crofloras lack the cold-tolerant genus Nothofagus,Romero
considered them among the warmest type of mixed pa-
leoflora. However, Nothofagus fusca–type pollen was re-
cently discovered at RP, occurring at low abundance (W.
Volkheime r, p e rsonal communi c a t i o n , 2 0 0 4 ; palynologi c a l
preparations from LH so far have been unproductive).
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
Eocene Plant Diversity in Patagonia 639
Figure 3: The 30 most abundant plant species at Laguna del Hunco; all localities grouped
Collections and Methods
Major new collections from LH reported here were made
in December of 2002 and are combined with the material
reported by Wilf et al. (2003; collected in 1999); all were
acquired using identical procedures, as was the RP sample.
All LH voucher specimens, more than 3,500 in total, are
housed in the Paleobotanical Collections of the Museo
Paleontolo´ gico Egidio Feruglio, in Trelew, Chubut Prov-
ince. The repository for the RP fossils is the Museo Pa-
leontolo´gico Asociacio´n Paleontolo´gica Bariloche in Bari-
loche, Rı´o Negro Province. Historic collections from LH
and RP are not included in our analyses because previous
researchers did not use an unbiased, quantitative, strati-
graphic methodology. To improve the comparison to
North American plant diversity, we have also reanalyzed
the sample from Republic, mentioned earlier, and have
incorporated additional specimens. Although an exact an-
alog to LH does not exist in North America, the Republic
flora has the most similar combination of age (49–50 Ma)
and paleoenvironmental setting to the Patagonian floras,
and it is extremely diverse (Wolfe and Wehr 1987; Pigg et
al. 2001; Wilf et al. 2003).
At LH, our major effort was intensive sampling of the
four most productive quarries identified in 1999. These
sites (LH-2, LH-4, LH-6, and LH-13; table 1; fig. 2) are
hereafter referred to as the principal or major quarries and
the remaining sites as the minor quarries. The total for
all unbiased, identifiable material is 4,303 plant specimens,
of which 3,723 (87%) come from the four major quarries
(table 1). We also collected 57 float specimens, mostly not
assignable to specific horizons; these are not included in
analyses except where mentioned specifically below.
At Rı´o Pichileufu´, we found two sites, RP-1 and RP-3,
with very good preservation and diverse floras, occupying
approximately the same stratigraphic position (we could
not correlate them precisely because of covered and locally
slumped bedding). We made a preliminary, unbiased col-
lection of 341 specimens from RP-3, the most productive
quarry, of which 252 were identifiable and 213 were iden-
tifiable dicot leaves assigned to 39 species. Three tuffs were
found in close stratigraphic proximity to the fossil floras;
tuff RP1 lies immediately on top of the RP-1 fossil layer,
tuff 2 occurs approximately 6 m below RP-1 (although its
position may be slumped), and tuff RP3 is located about
6maboveRP-3.Inaddition,asanindependentlaboratory
test, we reanalyzed a tuff from LH, 2211A, which previ-
ously yielded an age from U.S. Geological Survey labo-
ratories of Ma (Wilf et al. 2003). More than
52.13 !0.32
140
40
Ar/
39
Ar laser fusion and incremental heating analyses
of sanidine crystals from the four tuffs, using the meth-
odology of Smith et al. (2003), were undertaken at the
University of Wisconsin–Madison Rare Gas Geochronol-
ogy Laboratory. Details of radioisotopic analyses are pro-
vided in the appendix.
From our inspection of the type and referred material,
many of Berry’s identifications are clearly incorrect, are
based on poorly preserved specimens that would not merit
formal nomenclature today, or are not supported with
sufficient diagnostic characters by modern standards (see
also Dilcher 1973; Hill and Brodribb 1999). For these rea-
sons, most of Berry’s generic names are enclosed here in
quotations. Both Berry’s taxa and the large numbers of
new species found in our collections will require a major
effort toward systematic analysis and revision, now un-
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
640 The American Naturalist
Figure 4: Rarefaction results for identified dicot leaves, withcomparisons
to the most diverse Eocene floras known from North America collected
using similar techniques. A,FourmajorquarriesatLagunadelHunco
combined, showing the effect of increased sampling from 1999 (reported
in Wilf et al. 2003) to 2002 collections, compared to the four census sites
from Florissant, 34 Ma (Evanoff et al. 2001), Colorado, reported by
MacGinitie (1953). B,EffectofincreasedsamplingateachoftheLaguna
del Hunco major quarries. Vertical dotted lines connect the maximum
values of the 1999 data to the current rarefaction curves for each locality.
C,Single-quarrycomparisons,includingselected95%condenceinter-
vals, for comparison to the Republic flora (49–50 Ma, Washington, Den-
ver Museum of Nature and Science locality 2130). Republic data are
reanalyzed and updated for new specimens from Wilf et al. (2003), with
a resulting increase in rarefied diversity (the previous curve was nearly
identical to the present curve for LH-4). Rarefactions and confidence
intervals computed from the formulas of Tipper (1979). LH pLaguna
del Hunco; RP p´o Pichileufu´.
derway and led by M. A. Gandolfo (Gandolfo et al. 1988,
2004; Romero et al. 1988; Gonza´ lez et al. 2002).
Lack of a comprehensive reliable taxonomy is a typical
problem in floras dominated by angiosperm leaves, which
tend to be very diverse; even in floras that are relatively
well understood, it is not unusual for a significant fraction
of species to remain undiagnosed even to the family level
(e.g., Johnson 2002). However, a standard procedure ex-
ists, which relies on detailed leaf architectural analyses to
discriminate distinct leaf morphotypes that represent
probable biological species (Hickey 1979; Johnson et al.
1989; Ash et al. 1999). This approach, which we use here,
and others like it have been applied successfully to many
fossil floras to analyze ecological, diversity-related, and
paleoclimatic variables that require presence-absence or
relative abundance data for all species in a flora (Hill 1982;
Johnson et al. 1989; Burnham 1994; Wing 1998; Smith et
al. 1998; Jacobs and Herendeen 2004; Wilf and Johnson
2004) and to provide an initial framework for systematic
descriptions (Crane et al. 1990; Johnson 1996).
We follo w e d a c o nservati v e “ l u mping” a p p roach w h ere
awell-preserved,voucheredexemplarspecimenwitha
unique, reproducible suite of diagnostic architectural char-
acters and accompanying description must represent each
morphotype. The common preservation of nearly com-
plete leaves and fine venation features, especially at LH,
facilitated this methodology: the morphotypes established
from the 1999 collections, even when based on single spec-
imens, proved reliable for identifying 2002 collections in
the field and laboratory. For convenience, we will use “spe-
cies” here to indicate both formally described species and
morphotypes; comprehensive descriptions of these entities
are in separate preparation. Existing paleobotanical no-
menclature, abundance data, and voucher specimen num-
bers are provided in tables A2 and A3 in the appendix.
Mean annual temperature (MAT) and precipitation
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
Eocene Plant Diversity in Patagonia 641
Figure 5: Detrended correspondence analysis of major quarries and taphonomic variables of the Laguna del Hunco (LH) flora (from table 1; most
values entered as proportions). Axes are labeled with their percentages of total variance explained. Computed using MVSP 3.1 (Kovach 2000).
(MAP) were estimated for LH, the better-sampled flora,
using the standard techniques of leaf-margin (Wolfe 1979;
Wing and Greenwood 1993; Wilf 1997) and leaf-area anal-
ysis (Wilf et al. 1998), respectively.
Ages of the Floras
All three tuff beds collected at Rı´o Pichileufu´produced
abundant sanidine phenocrysts that yielded
40
Ar/
39
Ar ages
with low analytical uncertainties of Ma for47.48 !0.10
tuff RP1, Ma for tuff 2, and Ma47.48 !0.47 47.45 !0.06
for tuff RP3 (!95% confidence about the mean). These
ages were calculated relative to 28.34 Ma for the Taylor
Creek rhyolite sanidine (Renne et al. 1998) using the decay
constants of Steiger and Ja¨ger (1977). Given the uncer-
tainties, the three
40
Ar/
39
Ar ages for RP are indistinguish-
able from one another, thereby indicating a common de-
positional age for these sediments of Ma, the47.46 !0.05
weighted mean age from all three tuffs (middle Eocene,
Lutetian). In terms of global climate, this age places the
RP flora within the initial phase of cooling subsequent to
the early Eocene climatic optimum, although temperatures
were high at this time compared to the remainder of the
Cenozoic (Zachos et al. 2001). Results from the reanalyzed
Laguna del Hunco 2211A ash gave a weighted mean age
of Ma, indistinguishable from the51.91 !0.22 52.13 !
Ma age reported by Wilf et al. (2003). Therefore, both0.32
floras are securely dated at high precision, and the RP flora
is about 4.5 million years younger than the Laguna del
Hunco flora, previously considered as nearly coeval. These
results demonstrate for the first time longevity of the di-
verse Patagonian paleofloras, discussed in more detail
below.
Floral Diversity
The total count of plant organ species from our Laguna
del Hunco collections is 186, including 152 leaves and 132
dicot leaves (table 2); these totals are more than four times
the sum of historic data. The nondicotyledonous foliar
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
642 The American Naturalist
Figure 6: Cluster analyses, derived from Bray-Curtis distances, of relative abundance data from Laguna del Hunco (LH) plant species. A, Major
quarries, proportional data. B, Minor quarries with at least 40 specimens, raw data (sample size is insufficient for reliable proportional data,although
a similar clustering results). Note the stratigraphic correspondence of clusters to the major localities in A(cf. fig. 2). Computed using MVSP 3.1
(Kovach 2000).
species are eight conifers, one cycad, one Ginkgo, three
monocots, and seven ferns. The remaining species include
27 angiosperm fruits, seeds, and flowers; four or possibly
five types of coniferous cone scales, seeds, and pollen
cones; and one bryophyte. The total richness increases
slightly with float specimens, which include three addi-
tional species each of dicot leaves and dicot fruits. All four
principal localities produced comparable raw values for
plant richness, for example, 58–69 leaf species (table 2),
and the maximum occurred at LH-2. The most abundant
species at all sites combined are shown in rank order in
figure 3; thermophilic families are abundant, such as Myr-
taceae (“Myrcia”), Sapindaceae (“Schmidelia,” “Cupania”),
Fabaceae (“Cassia”), Lauraceae, and Araucariaceae. Al-
though Lomatia is common, the prevalence of “temperate”
or tropical montane groups is hard to assess because many
common species are poorly understood taxonomically.
Rarefaction analyses of dicot leaf data demonstrate high
plant diversity at LH at robust sample sizes (fig. 4). In
general, increased sampling led to dramatic increases in
species recovered, and individual rarefaction trajectories
remained the same or increased. The four principal quar-
ries combined show increasing diversity past 3,000 spec-
imens, along the trajectory of the 1999 collections (fig.
4A). In contrast, MacGinitie’s (1953) classic collections
from Florissant, Colorado, the most comparable data set
available for quarries from several stratigraphic levels
within a short-lived lacustrine sequence, do not produce
new species after about 1,000 specimens (fig. 4A). At the
principal localities, numerous additional species accu-
mulated, mostly along the previous rarefaction trajectories
(fig. 4B). The most significant change occurred at LH-13,
which showed a dramatic jump in rarefied richness and a
total increase in dicot leaves from 24 to 55 species. The
maximum raw richness occurred at LH-2, which increased
from 43 to 61 dicot leaf species.
All four principal quarries have a rarefied richness com-
parable to the extremely diverse Republic flora, and two,
LH-2 and LH-13, are more rich (fig. 4C). The difference
is significant for LH-2, based on separation of 95% con-
fidence intervals, and at 600 specimens, LH-2 has one-
third more species than Republic (61 vs. 46). This is the
case even though the reanalyzed Republic sample has
higher rarefied diversity than reported earlier (Wilf et al.
2003). Although rarefaction curves are beginning to flatten
at the four principal quarries, there is great potential for
asignicantnumberofadditionalplantspeciestobe
found at LH, considering the large area of exposure
available.
The preliminary sample from RP is as diverse on a
rarefied basis as the most diverse LH locality, LH-2 (fig.
4C), supporting the richness described in the historic
monograph (Berry 1938) and our observations of corre-
sponding type material. This, along with our geochro-
nologic results, shows exceptional plant diversity at a dif-
ferent time and location in Eocene Patagonia from LH.
Moreover, the high diversity recovered at RP, despite hav-
ing a distinctly lower quality of preservation than LH,
diminishes the possible importance of preservation bias
on rarefaction results for both floras.
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
Eocene Plant Diversity in Patagonia 643
Figure 7: Percent abundance at each major quarry for the 10 most abundant species of dicot leaves at Laguna del Hunco (LH), calculated relative
to all dicot leaves from the quarry.
Stability of Floral Associations
To w h a t e x t e n t d o t h e L H a n d R P fl o r a s r e p r e s e n t a b r o a d l y
coherent, persistent floral association versus a suite of
mostly unrelated floras that share only high richness as a
common characteristic? Abundant evidence indicates the
former at LH, where the major source of variation in the
floras appears to be environment of deposition. The four
principal quarries show differences in taphonomic char-
acteristics that apparently reflect distance from shoreline
(table 1); LH-6 has the most distinct offshore signature,
seen especially in its thin, well-laminated bedding and the
presence of articulated fossil fish, symmetrically winged
(“helicopter”) dicot fruits, and relatively abundant insect
fossils (the latter two also at LH-13). Helicopter fruits and
insects are likely to be entrained in prevailing winds and
deposited far from their source, where they become better
represented in fossil deposits than material deposited solely
by gravity and water (Wilson 1980; Augspurger 1986). In
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
Eocene Plant Diversity in Patagonia 645
Figure 8: Paleoclimatic results from (A) leaf-margin and (B) leaf-area analysis, Laguna del Hunco bulk flora (see table 2), plotted using the method
of Burnham et al. (2001), where successive estimates are graphed from left to right as species are added in decreasing rank abundance order.
Abundances of the species as they are added are shown in the small accessory graphs; the two abundance graphs differ slightly because some
specimens suitable for leaf-margin analysis were too fragmentary for reliable measurement of leaf area. In A, the central horizontal line shows the
best value for the sample based on all species. Symmetrical curves show 1 SD (standard deviation) about the best value based on number of species
and a binomial probability model (Wilf 1997). Temperature estimates above or below these curves indicate leaf-margin percentages that deviate
significantly from a random order of margin types taken from the entire sample. The accessory horizontal lines show !2#C from the best value,
considered to be a minimum error for leaf-margin analysis when the binomial error is less than !2#C (Wilf 1997). Bincludes estimated mean
annual precipitation and the standard error of the estimate using the regression of Wilf et al. (1998); errors are asymmetrical because they are
transformed from logarithmic data. annual temperature; annual precipitation.MAT pmean MAP pmean
adetrendedcorrespondenceanalysis(DCA)ofquarries
and taphonomic variables (fig. 5), LH-6 is alone in its
position on the first axis, and its most strongly associated
variables, in decreasing order, are the presence of fish,
winged fruits, and insects.
Clustering based on Bray-Curtis distance, which is com-
puted using relative abundance data, grouped locality LH-
6withLH-13andlocalityLH-2withLH-4(g.6A). A
similar grouping of LH-6 and LH-13 along the first axis
results from a DCA of species relative abundance and quar-
ries (not shown). Although data from minor localities are
preliminary because of their small sample sizes, cluster
analysis based on Bray-Curtis distance also resolved the
stratigraphic groupings of six minor localities with at least
40 specimens each (fig. 6B).
The clustering of the stratigraphically lowest (LH-13)
and highest (LH-6) major and corresponding minor lo-
calities (fig. 6), combined with the prevalence of the same
dominant taxa at all localities (fig. 7), suggests that floral
differences within the LH sequence reflect paleoenviron-
ment of deposition more than compositional change
through time. These differences are mostly expressed in
varying relative abundance (figs. 6, 7) and not in com-
position. We therefore consider the LH floras to represent
a single major association that is preserved in a variety of
lacustrine environments.
The Laguna del Hunco and Rı´o Pichileufu´florashave
been considered to be similar in species composition since
Berry’s original treatments (Berry 1925, 1938; Arguijo and
Romero 1981; Markgraf et al. 1996; Hinojosa and Villagra´n
1997). Notably, Petersen (1946) listed 50 species shared
between the floras, but no surviving voucher specimens
are known that could be used to evaluate his identifica-
tions. Because the floras are shown here to be separated
by 4.5 million years, the assumption of species-level sim-
ilarity deserves a critical reevaluation, although significant
overlap of angiosperm families is indisputable (appendix).
A reevaluation of shared leaf species (table A4 in the ap-
pendix) produces a list of 10 angiosperms that is short in
comparison to the prodigious angiosperm diversity of the
two sites already discussed, although it is likely to grow
with increased collecting at RP. Significantly, only one of
the 10 most abundant dicot species at RP, “Cassiaargen-
tinensis, is among the top 10 at LH (fig. 7). Thus, the best
current support for close species similarity among the flo-
ras is in the less diverse gymnosperm fraction (table A4).
Paleoclimate
Our paleoclimatic results indicate that Eocene Laguna del
Hunco was neither warm nor humid enough to be a trop-
ical rain forest as suggested in previous global reconstruc-
tions (Wolfe 1985; Frakes et al. 1992; Morley 2000), al-
though moisture was abundant and the maritime climate
limited seasonality to a narrow range of temperatures that
precluded frost. Results for major localities and associated
sampling levels at LH are shown in table 2, along with
notes on methodology. The high numbers of species used
in the estimates help to counteract well-known biases from
undersampling, paleoenvironment, and taphonomy
(Burnham 1994; Wilf 1997; Burnham et al. 2001). For
MAT at individual sampling levels, there is less than 5#C
spread between the minimum estimate of C14.2#!2.0#
at the LH-13 level and the maximum estimate of
#C at the LH-6 level (table 2). Because there is18.6#!2.1
no evidence for significant change in floral composition
or available moisture, major shifts in temperature do not
seem likely. Accordingly, we place the greatest confidence
in the lumped estimate from all localities, #C,16.6#!2.0
based on 131 dicot leaf species. This result is within the
range (16#–17#C) of coeval sea surface temperatures an-
alyzed from several cores at bounding paleolatitudes of the
South Atlantic (Zachos et al. 1994).
Paleoprecipitation estimates for LH are nearly identical
at the various sampling levels (table 2). The 119 species
that were measurable for leaf size produce a lumped es-
timate of about 1.1 m/year. We consider this a minimum
estimate because transport into lakes generally selects
against larger leaves (Roth and Dilcher 1978). Although
several species are represented by very large leaves (area
110,000 mm
2
), their signal is diluted by numerous rare
species assumed to be undersampled for leaf size.
Sensitivity of climate estimates to sample size is tested
using the method of Burnham et al. (2001), where suc-
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
646 The American Naturalist
cessive estimates are graphed as species are added in de-
creasing rank abundance order (fig. 8A). For the LH bulk
flora, MAT estimates begin to converge near 60 species,
which suggests that the estimated paleotemperature is very
unlikely to change significantly with additional sampling,
although increased resolution of climate from individual
horizons may be possible. A similar plot of MAP estimates
shows instability below about 40 species but little variation
past about 60 species (fig. 8B).
The modern climatic ranges of several higher taxa in
the LH flora are well studied, helping to define climatic
seasonality using nearest living relatives. The combined
presence of palms, cycads, at least two species of araucarian
conifers, at least five species of podocarps, and possibly
three species of Gymnostoma (Gandolfo et al. 2004), along
with numerous additional tropical elements (appendix)
and the apparent absence of Nothofagus,providesstrong
evidence for an equable climate with winter mean tem-
peratures warmer than 10#Candabundantrainfall(Chris-
tophel 1980; Romero 1986; Hill 1994; Greenwood and
Wing 1995; Hill and Brodribb 1999; Kershaw and Wagstaff
2001).
From the RP-3 locality, 23 of the 39 dicot species were
untoothed, yielding a preliminary MAT estimate of
#C. However, greater species sampling at RP19.2#!2.4
will probably change this estimate (see fig. 8). Global ma-
rine data indicate cooling in the interval from 52 to 47
Ma (Zachos et al. 2001), which can be tested in Patagonia
with increased sampling at RP.
The nontropical but moist and equable climates may
explain the long-observed “mixing” (sensu Romero 1978,
1993) of temperate and tropical elements; the absence of
drought and frost fostered the growth of tropical lineages,
while the lack of high temperatures allowed the simulta-
neous presence of Gondwanic elements (see Axelrod et al.
1991). Except for moist corridors that probably existed to
the Neotropical region, the biome containing the LH and
RP floras was bounded by cooler winters to the south and
aridity to the north. This biome could be considered an
ancient biodiversity “hot spot,” although much better doc-
umentation of floral endemism than may ever be available
from fossils would be required to test a hot spot hypothesis
by comparison to modern definitions (e.g., Mittermeier et
al. 1998).
Discussion
The Laguna del Hunco and Rı´o Pichileufu´florasrepresent
ahighlybiodiverse,floristicallyandclimaticallydistinct
region. They provide the best macrofossil evidence for an
ancient history of high plant diversity in Cenozoic South
America. The 4.5-million-year age difference between
them shows that richness was both long-lived and wide-
spread in Patagonia. Their diversity exceeds that of any
comparable Eocene flora from the middle latitudes ofwest-
ern North America, the only region with quantitative data
derived using equivalent collection strategies, and exceeds
on an absolute basis all Eocene leaf floras known to us
from outside the Americas. Even without sample size cor-
rection, organ type equivalence, or stratigraphic control in
the comparisons, they rank among the most diverse fossil
macrofloras known from any time period (e.g., Knoll et
al. 1979). Eocene floras with equal or greater species rich-
ness, such as the Clarno nut beds (Manchester 1994) and
the London Clay flora (Reid and Chandler 1933; Collinson
1983), are fruit and seed assemblages, which are produced
via fundamentally different taphonomic pathways and in-
creased temporal averaging in comparison to assemblages
dominated by leaves (Behrensmeyer et al. 2000). They also
have been collected selectively over many decades from
unknown numbers of specimens, assumed to be in the
tens to hundreds of thousands. The Castle Rock flora,
which preserves forest floor litter from the early Paleocene
of Colorado (Johnson and Ellis 2002; Ellis et al. 2003), is
the only known assemblage, dominated by angiosperm
leaves and collected using comparable methods to LH and
RP, with similar species diversity. However, it represents a
true rain forest with a significantly warmer and wetter
environment than the Patagonian floras (Ellis et al. 2003);
this climate should allow the presence of elevated plant
diversity by analogy to living forests (Phillips and Miller
2002).
Diverse floras at Patagonian middle latitudes imply great
richness at tropical latitudes of Eocene South America via
the latitudinal diversity gradient, which appears to apply
to most present and past vegetation (Crane and Lidgard
1989; Willig et al. 2003; Harrington 2004; Hillebrand
2004). Palynological data from the Paleocene-Eocene in-
terval of Colombia strongly support a hyperdiverse Eocene
tropics (Jaramillo 2002), and a latitudinal diversity gra-
dient existed for Eocene vegetation in North America
(Harrington 2004). Patagonian diversity was probably
driven both by in situ evolution and the “spillover” of
tropical taxa from extremely diverse floras of lower lati-
tudes with Eocene warming. Immigrants from low lati-
tudes consequently mixed with resident, temperate, and
endemic taxa as well as tropical migrants from other con-
tinents. An analogous effect has been observed in Eocene
palynofloras from North America (Harrington 2004).
Reliable quantitative paleontological data for plant di-
versity from other areas of South America remain rare,
and diversity fluctuated between the Eocene and Recent
(Jaramillo 2003). Any explanation for past diversity or its
connection to modern floral richness is speculative, but
we make a few general points here. South America has
possessed immense areas of low-latitude land surface since
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
Eocene Plant Diversity in Patagonia 647
the Early Cretaceous and of tropical rain forest cover, dom-
inated by angiosperms, at low latitudes since at least the
late Paleocene (Morley 2000; Schettino and Scotese 2001;
Jaramillo 2002; Wing et al. 2004). Accordingly, there has
been great potential for diversification through time from
populations separated by geography, climate, or ecological
selection regime, as supported by studies in modern Ama-
zonia (Gentry 1988a;Schneideretal.1999;Moritzetal.
2000; Smith et al. 2001; Ogden and Thorpe 2002; Tuomisto
et al. 2003; Fine et al. 2004). The association of global
warming across the Paleocene-Eocene boundary and sig-
nificant in situ diversification of palynofloras in northern
South America (Rull 1999; Jaramillo 2002) suggests that
a major diversity increase occurred in Eocene South Amer-
ica, mediated by climate change. Testing this hypothesis
in Patagonia awaits quantitative analyses of suitable Pa-
leocene floras.
Acknowledgments
For generous support, we thank the National Geographic
Society (grant 7337-02), the National Science Foundation
(grants DEB-0345750, EAR-0230123, and EAR-0114055),
the University of Pennsylvania Research Foundation, the
Andrew W. Mellon Foundation, and the Petroleum Re-
search Fund (grant 35229-G2). For exceptional assistance
in the field and laboratory, we are grateful to M. Caffa, L.
Canessa, B. Cariglino, I. Escapa, C. Gonza´lez, R. Horwitt,
P. P u e r t a , E . R u i g o m e z , H . S m e k a l , X . Z h a n g , a n d t h e
Museo Paleontolo´gico Asociacio´n Paleontolo´gica Barilo-
che. Comments on drafts and reviews by E. Arago´n, R.
Horwitt, C. Jaramillo, D. Royer, and two anonymous re-
viewers greatly improved the manuscript. We are grateful
to the Nahueltripay family and INVAP (Instituto de In-
vestigaciones Aplicadas) for land access.
Literature Cited
Arago´ n, E., and M. M. Mazzoni. 1997. Geologı´a y estratigrafı´a del
complejo volca´ nico pirocla´ stico del Rı´o Chubut medio (Eoceno),
Chubut, Argentina. Revista de la Asociacio´n Geolo´gica Argentina
52:243–256.
Arago´ n, E., and E. J. Romero. 1984. Geologı´a, paleoambientes y
paleobota´nica de yacimientos Terciarios del occidente de Rı´o Ne-
gro, Neuque´n y Chubut. Actas del IX Congreso Geolo´gicoArgen-
tino, San Carlos de Bariloche 4:475–507.
Archangelsky, S. 1974. Sobre la edad de la tafoflora de la Laguna del
Hunco, Provincia de Chubut. Ameghiniana 11:413–417.
Arguijo, M. H., and E. J. Romero. 1981. Analisis bioestratigrafico de
formaciones portadores de tafofloras Terciarias. Actas del VIII
Congreso Geolo´ gico Argentino 4:691–717.
Ash, A. W., B. Ellis, L. J. Hickey, K. R. Johnson, P. Wilf, and S. L.
Wing. 1999. Manual of leaf architecture: morphologicaldescription
and categorization of dicotyledonous and net-veined monocoty-
ledonous angiosperms. Smithsonian Institution, Washington, DC.
Augspurger, C. K. 1986. Morphology and dispersal potential of wind-
dispersed diaspores of Neotropical trees. American Journal of Bot-
any 73:353–363.
Axelrod, D. I., M. K. Arroyo, and P. H. Raven. 1991. Historical
development of temperate vegetation in the Americas. Revista Chi-
lena de Historia Natural 64:413–446.
Ba´ ez, A. M., and L. Trueb. 1997. Redescription of the Paleogene
Shelania pascuali from Patagonia and its bearing on the relation-
ships of fossil and Recent pipoid frogs. Scientific Papers, Natural
History Museum, The University of Kansas 4:1–41.
Behrensmeyer, A. K., S. M. Kidwell, and R. A. Gastaldo. 2000. Ta-
phonomy and paleobiology. Paleobiology 26S:103–147.
Berry, E. W. 1925. A Miocene flora from Patagonia. Johns Hopkins
University Studies in Geology 6:183–251.
———. 1935a. A fossil Cochlospermum from northern Patagonia.
Bulletin of the Torrey Botanical Club 62:65–67.
———. 1935b. The Monimiaceae and a new Laurelia.Botanical
Gazette 96:751–754.
———. 1935c. A Tertiary Ginkgo from Patagonia. Torreya 35:11–13.
———. 1938. Tertiary flora from the Rı´o Pichileufu´,Argentina.Geo-
logical Society of America Special Paper 12:1–149.
Bremer, K. 2002. Gondwanan evolution of the grass alliance of fam-
ilies (Poales). Evolution 56:1374–1387.
Burnham, R. J. 1994. Paleoecological and floristic heterogeneity in
the plant-fossil record: an analysis based on the Eocene of Wash-
ington. U.S. Geological Survey Bulletin 2085-B:1–36.
Burnham, R. J., and A. Graham. 1999. The history of Neotropical
vegetation: new developments and status. Annals of the Missouri
Botanical Garden 86:546–589.
Burnham, R. J., and K. R. Johnson. 2004. South American palaeo-
botany and the origins of Neotropical rainforests. Philosophical
Tran s act i ons o f t h e R oya l S o cie t y of Lon d on B 359 : 159 5 161 0 .
Burnham, R. J., N. C. A. Pitman, K. R. Johnson, and P. Wilf. 2001.
Habitat-related error in estimating temperatures fromleaf margins
in a humid tropical forest. American Journal of Botany 88:1096–
1102.
Casamiquela, R. M. 1961. Un pipoideo fo´sil de Patagonia. Revista
del Museo de La Plata, Seccio´n Paleontologı´a 4:71–123.
Christophel, D. C. 1980. Occurrence of Casuarina megafossils in the
Tertiary of southeastern Australia. Australian Journal of Botany
28:249–259.
Clark, B. 1923. The economic deduction from an incidental general
reconnaissance of Patagonia between east central Santa Cruz and
northwestern Chubut. ESSO, Field Office Argentine No. 35, Office
Argentine No. 70, New York.
Colinvaux, P. A., and P. E. De Oliveira. 2001. Amazon plant diversity
and climate through the Cenozoic. Palaeogeography, Palaeocli-
matology, Palaeoecology 166:51–63.
Colinvaux, P. A., P. E. De Oliveira, J. E. Moreno, M. C. Miller, and
M. B. Bush. 1996. A long pollen record from lowland Amazonia:
forest and cooling in glacial times. Science 274:85–88.
Colinvaux, P. A., G. Irion, M. E. Rasanen, M. B. Bush, and J. A. S.
N. De Mello. 2001. A paradigm to be discarded: geological and
paleoecological data falsify the Haffer and Prance refuge hypothesis
of Amazonian speciation. Amazoniana 16:609–646.
Collinson, M. E. 1983. Fossil plants of the London Clay. Palaeon-
tological Association, London.
Crane, P. R., and S. Lidgard. 1989. Angiosperm diversification and
paleolatitudinal gradients in Cretaceous floristic diversity. Science
246:675–678.
Crane, P. R., S. R. Manchester, and D. L. Dilcher. 1990. A preliminary
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
648 The American Naturalist
survey of fossil leaves and well-preserved reproductive structures
from the Sentinel Butte Formation (Paleocene) near Almont,
North Dakota. Fieldiana Geology 20:1–63.
Davis, C. C., P. W. Fritsch, C. D. Bell, and S. Mathews. 2004. High-
latitude Tertiary migrations of an exclusively tropical clade: evi-
dence from Malpighiaceae. International Journal of Plant Sciences
165(suppl.):S107–S121.
Davis, C. C., C. O. Webb, K. J. Wurdack, C. A. Jaramillo, and M. J.
Donoghue. 2005. Explosive radiation of Malpighiales supports a
mid-Cretaceous origin of modern tropical rain forests. American
Naturalist 165:E36–E65.
Davis, S. D., V. H. Heywood, O. Herrera MacBryde, and A. C. Ham-
ilton. 1997. Centres of plant diversity: a guide and strategy for
their conservation. Vol. 3. The Americas. World Wide Fund for
Nature, London.
Dilcher, D. L. 1973. A revision of the Eocene flora of southeastern
North America. Palaeobotanist 20:7–18.
Dlussky, G. M., and K. S. Perfilieva. 2003. Paleogene antsof the genus
Archimyrmex Cockerell, 1923. Paleontological Journal 37:39–47.
Durango de Cabrera, J., and E. J. Romero. 1986. Roupala patagonica
n. sp. de Laguna del Hunco (Paleocene), Provincia Chubut, Ar-
gentina. Actas del V Congreso Argentino de Paleontolo´ gia y Bioes-
tratigrafı´a 3:121–124.
Ellis, B., K. R. Johnson, and R. E. Dunn. 2003. Evidence for an in
situ early Paleocene rainforest from Castle Rock, Colorado. Rocky
Mountain Geology 38:73–100.
Estes, R., and J. H. Hutchison. 1980. Eocene lower vertebrates from
Ellesmere Island, Canadian Arctic Archipelago. Palaeogeography,
Palaeoclimatology, Palaeoecology 30:325–347.
Evanoff, E., W. C. McIntosh, and P. C. Murphey. 2001. Stratigraphic
summary and
40
Ar/
39
Ar geochronology of the Florissant Formation,
Colorado. Pages 1–16 in E. Evanoff, K. M. Gregory-Wodzicki, and
K. R. Johnson, eds. Fossil flora and stratigraphy of the Florissant
Formation, Colorado. Denver Museum of Nature and Science,
Denver, CO.
Fidalgo, P., and D. R. Smith. 1987. A fossil Siricidae Hymenoptera
from Argentina. Entomological News 98:63–66.
Fine, P. V. A., I. Mesones, and P. D. Coley. 2004. Herbivores promote
habitat specialization by trees in Amazonian forests. Science 305:
663–665.
Frakes, L. A., J. E. Francis, and J. I. Syktus. 1992. Climate modes of
the Phanerozoic. Cambridge University Press, Cambridge.
Frenguelli, J. 1943a.ProteaceasdelCenozoicodePatagonia.Notas
del Museo de La Plata 8:201–213.
———. 1943b.RestosdeCasuarina en el Mioceno de El Mirador,
Patagonia central. Notas del Museo de La Plata 8:349–354.
Frenguelli, J., and L. R. Parodi. 1941. Una Chusquea fo´sil de El
Mirador (Chubut). Notas del Museo de La Plata 6:235–238.
Gandolfo, M. A., M. C. Dibbern, and E. J. Romero. 1988. Akania
patagonica n. sp. and additional material on Akania americana
Romero and Hickey (Akaniaceae), from Paleocene sediments of
Patagonia. Bulletin of the Torrey Botanical Club 115:83–88.
Gandolfo, M. A., M. C. Zamaloa, C. C. Gonza´ lez, N. R. Cu´neo, P.
Wilf, and E. J. Romero. 2004. Ea rly history of Casuarinaceae in
the Paleogene of Patagonia, Argentina. International Organization
of Paleobotany, Seventh Quadrennial Conference, Bariloche, Ar-
gentina, Abstracts, pp. 36–37.
Genise, J. F., and J. F. Petrulevicius. 2001. Caddisfly cases from the
early Eocene of Chubut, Patagonia, Argentina. Second Interna-
tional Congress on Paleoentomology, September 5–9, 2001, Kra-
ko´ w, Poland, Abstracts, pp. 12–13.
Gentry, A. H. 1988a.Changesinplantcommunitydiversityand
floristic composition on environmental and geographical gradi-
ents. Annals of the Missouri Botanical Garden 75:1–34.
———. 1988b. Tree species richness of upper Amazonian forests.
Proceedings of the National Academy of Sciences of the USA 85:
156–159.
Gonza´ lez, C. C., M. A. Gandolfo, N. R. Cu´ neo, and P. Wilf. 2002.
Revisio´n de las Myrtaceae de Laguna del Hunco y Rı´o Pichileufu´
(Eoceno Inferior), Patagonia, Argentina. Simposio Argentino de
Paleobotanica y Palinologı´a, Abstracts, p. 12.
Gonza´ lez Dı´az, E. F. 1979. La edad de la Formacio´n Ventana, en la
area al norte y al este del Lago Nahuel Huapi. Revista de la Aso-
ciacio´n Geolo´gica Argentina 34:113–124.
Greenwood, D. R., and S. L. Wing. 1995. Eocene continentalclimates
and latitudinal temperature gradients. Geology 23:1044–1048.
Haffer, J. 1969. Speciation in Amazonian forest birds. Science 165:
131–137.
Haffer, J., and G. T. Prance. 2001. Climatic forcing of evolution in
Amazonia during the Cenozoic: on the refuge theory of biotic
differentiation. Amazoniana 16:579–605.
Harrington, G. J. 2004. Structure of the North American vegetation
gradient during the late Paleocene/early Eocene warm climate.
Evolutionary Ecology Research 6:33–48.
Hay, W. W., R. DeConto, C. N. Wold, K. M. Wilson, S. Voigt, M.
Schulz, A. Wold-Rossby, et al. 1999. Alternative global Cretaceous
paleogeography. Geological Society of America Special Paper 332:
1–47.
Hickey, L. J. 1979. A revised classification of the architecture of
dicotyledonous leaves. Pages 25–39 in C. R. Metcalfe and L. Chalk,
eds. Anatomy of the dicotyledons. 2nd ed. Clarendon, Oxford.
Hill, R. S. 1982. The Eocene megafossil flora of Nerriga, New South
Wal e s, Aus tra lia . P ala eon tog r ap h ic a A bte ilu ng B: P ala e op h yt o lo g ie
181:44–77.
———. 1994. The history of selected Australian taxa. Pages 390–
419 in R. S. Hill, ed. History of the Australian vegetation: Cre-
taceous to Recent. Cambridge University Press, Cambridge.
Hill, R. S., and T. J. Brodribb. 1999. Southern conifers in time and
space. Australian Journal of Botany 47:639–696.
Hill, R. S., and R. J. Carpenter. 1991. Extensive past distributions for
major Gondwanic floral elements: macrofossil evidence. Pages
239–247 in M. R. Banks, S. J. Smith, A. E. Orchard, and G. Kant-
vilas, eds. Aspects of Tasmanian botany. Royal Society of Tasmania,
Hobart.
Hillebrand, H. 2004. On the generality of the latitudinal diversity
gradient. American Naturalist 163:192–211.
Hinojosa, L. F., and C. Villagra´n. 1997. Historyof thesouthern South
American forests. I. Paleobotanical, geological and climaticalback-
ground on Tertiary of southern South America. Revista Chilena
de Historia Natural 70:225–239.
Hoorn, C. 1994. Fluvial paleoenvironments in theintracratonic Ama-
zonas Basin (early Miocene–early middle Miocene, Colombia).
Palaeogeography, Palaeoclimatology, Palaeoecology 109:1–54.
Jacobs, B. F., and P. S. Herendeen. 2004. Eocene dry climate and
woodland vegetation in tropical Africa reconstructed from fossil
leaves from northern Tanzania. Palaeogeography, Palaeoclimatol-
ogy, Palaeoecology 213:115–123.
Jaramillo, C. A. 2002. Response of tropical vegetation to Paleogene
warming. Paleobiology 28:222–243.
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
Eocene Plant Diversity in Patagonia 649
———. 2003. Maastrichtian to early Miocene patterns of plant di-
versification in the Neotropics. Geological Society of America Ab-
stracts with Programs 35:58.
Jaramillo, C. A., and D. L. Dilcher. 2000. Microfloral diversity pat-
terns of the late Paleocene–Eocene interval in Colombia, northern
South America. Geology 28:815–818.
Johnson, K. R. 1996. Description of seven common plant megafossils
from the Hell Creek Formation (Late Cretaceous: late Maastrich-
tian), North Dakota, South Dakota, and Montana. Proceedings of
the Denver Museum of Natural History 3:1–48.
———. 2002. The megaflora of the Hell Creek and lower Fort Union
formations in the western Dakotas: vegetational response to cli-
mate change, the Cretaceous-Tertiary boundary event, and rapid
marine transgression. Geological Society of America Special Paper
361:329–391.
Johnson, K. R., and B. Ellis. 2002. A tropical rainforest in Colorado
1.4 million years after the Cretaceous-Tertiary boundary. Science
296:2379–2383.
Johnson, K. R., D. J. Nichols, M. Attrep Jr., and C. J. Orth. 1989.
High-resolution leaf-fossil record spanning the Cretaceous-
Tertiary boundary. Nature 340:708–711.
Kastner, T. P., and M. A. Gon˜ i. 2003. Constancy in the vegetation
of the Amazon Basin during the late Pleistocene: evidence from
the organic matter composition of Amazon deep sea fan sediments.
Geology 31:291–294.
Kershaw, P., and B. Wagstaff. 2001. The southern conifer family
Araucariaceae: history, status, and value for paleoenvironmental
reconstruction. Annual Review of Ecology and Systematics32:397–
414.
Knoll, A. H., K. J. Niklas, and B. H. Tiffney. 1979. Phanerozoic land-
plant diversity in North America. Science 206:1400–1402.
Kovach, W. L. 2000. MVSP: a multivariate statistical package for
Windows. Version 3.12c. Kovach Computing, Petraeth, Wales.
MacGinitie, H. D. 1953. Fossil plants of the Florissant beds,Colorado.
Carnegie Institution of Washington Publication 599:1–198.
Manchester, S. R. 1994. Fruits and seeds of the middle Eocene nut
beds flora, Clarno Formation, Oregon. Palaeontographica Amer-
icana 58:1–205.
Markgraf, V., E. J. Romero, and C. Villagra´n. 1996. History and
paleoecology of South American Nothofagus forest. Pages 354–386
in T. T. Veb l en , R . S. H i ll, a nd J. R ead , eds . The e col ogy a nd
biogeography of Nothofagus forests. Yale University Press, New
Haven, CT.
Marshall, L. G., and P. Salinas. 1990. Stratigraphy of the Rı´o Frı´as
Formation (Miocene), along the Alto Rı´o Cisnes, Aise´n, Chile.
Revista Geolo´gica de Chile 17:57–87.
Mazzoni, M. M., K. Kawashita, S. Harrison, and E. Arago´n. 1991.
Edades radime´ tricas Eocenas en el borde occidental del Macizo
Norpatago´ nico. Revista de la Asociacio´n Geolo´gica Argentina 46:
150–158.
Melchor, R. N., J. F. Genise, and S. E. Miquel. 2002. Ichnology,
sedimentology and paleontology of Eocene calcareous paleosols
from a palustrine sequence, Argentina. Palaios 17:16–35.
Mello, C. L., L. P. Bergqvist, and L. G. Sant’Anna. 2002. Fonseca,
MG: vegetais fo´sseis do Tercia´rio brasileiro. Pages 73–79 in C.
Schobbenhaus, D. A. Campos, E. T. Queiroz, M. Winge, and M.
L. C. Berbert-Bron, eds. Sı´tios geolo´gicos e paleontolo´ gicos do
Brasil. Comissa˜o Brasileira de Sı´tios Geolo´gicose Paleobiolo´ gicos,
Brasilia.
Mene´ndez, C.A. 1971. Floras Terciarias delaArgentina. Ameghiniana
8:357–368.
Mittermeier, R. A., N. Myers, J. B. Thomsen, G. A. B. da Fonseca,
and S. Olivieri. 1998. Biodiversity hotspots and major tropical
wilderness areas: approaches to setting conservation priorities.
Conservation Biology 12:516–520.
Mori, S. A., and B. B. Boom. 1981. Final report to the World Wildlife
Fund U.S. on the botanical survey of the endangered moist forests
of eastern Brazil. New York Botanical Garden, New York.
Moritz, C., J. L. Patton, C. J. Schneider, and T. B. Smith. 2000.
Diversification of rainforest faunas: an integrated molecular ap-
proach. Annual Review of Ecology and Systematics 31:533–563.
Morley, R. J. 2000. Origin and evolution oftropical rain forests. Wiley,
New York.
Nelson, B. W., C. A. C. Ferreira, M. F. da Silva, and M. L. Kawasaki.
1990. Endemism centers, refugia and botanical collection density
in Brazilian Amazonia. Nature 345:714–716.
Ogden, R., and R. S. Thorpe. 2002. Molecular evidence for ecological
speciation in tropical habitats. Proceedings of the National Acad-
emy of Sciences of the USA 99:13612–13615.
Passmore, S. M., K. R. Johnson, M. Reynolds, M. Scott, and D.
Meade-Hunter. 2002. Through the Quaternary looking glass: the
middle Eocene Republic flora over short timescales. Geological
Society of America Abstracts with Programs 34:556.
Petersen, C. S. 1946. Estudios geolo´ gicos en la regio´n del ´o Chubut
medio. Direccio´ n de Minas y Geologı´a Boletı´n 59:1–137.
Petrulevicius, J. F., and A. D. Nel. 2003. Frenguelliidae, a new family
of dragonflies from the earliest Eocene of Argentina (Insecta:
Odonata): phylogenetic relationships within Odonata. Journal of
Natural History 37:2909–2917.
Phillips, O. L., and J. S. Miller. 2002. Global patterns ofplant diversity:
Alwyn H. Gentry’s forest transect data set. Missouri Botanical
Garden Press, St. Louis.
Pigg, K. B., W. C. Wehr, and S. M. Ickert-Bond. 2001. Troc hod en dro n
and Nordenskioldia (Trochodendraceae) from the middle Eocene
of Washington State, USA. International Journal of Plant Sciences
162:1187–1198.
Pole, M. S., and M. K. MacPhail. 1996. Eocene Nypa from Regatta
Point, Tasmania. Review of Palaeobotany and Palynology 92:55–
67.
Rapela, C. W., L. A. Spalletti, J. C. Merodio, and E. Arago´ n. 1988.
Temporal evolution and spatial variation of early Tertiary volca-
nism in the Patagonian Andes (40#S-42#30!S). Journal of South
American Earth Sciences 1:75–88.
Reid, E. M., and M. E. J. Chandler. 1933. The London Clay flora.
British Museum (Natural History), London.
Renne, P. R., C. C. Swisher, A. L. Deino, D. B. Karner, T. L. Owens,
and D. J. DePaolo. 1998. Intercalibration of standards, absolute
ages and uncertainties in
40
Ar/
39
Ar dating. Chemical Geology 145:
117–152.
Richardson, J. E., R. T. Pennington, T. D. Pennington, and P. M.
Hollingsworth. 2001. Rapid diversification of a species-rich genus
of Neotropical rain forest trees. Science 293:2242–2245.
Romero, E. J. 1978. Paleoecologı´a y paleofitografı´a de las tafofloras
del Cenofitico de Argentina y areas vecinas. Ameghiniana 15:209–
227.
———. 1986. Paleogene phytogeography and climatology of South
America. Annals of the Missouri Botanical Garden 73:449–461.
———. 1993. South American paleofloras. Pages 62–85 in P. G o l d -
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
650 The American Naturalist
blatt, ed. Biological relationships between Africa and South Amer-
ica. Yale University Press, New Haven, CT.
Romero, E. J., and L. J. Hickey. 1976. Fossil leaf of Akaniaceae from
Paleocene beds in Argentina. Bulletin of the Torrey BotanicalClub
103:126–131.
Romero, E. J., M. C. Dibbern, and M. A. Gandolfo. 1988. Revisio´n
de Lomatia bivascularis (Berry) Frenguelli (Proteaceae) del yaci-
miento de la Laguna del Hunco (Paleoceno), Pcia. del Chubut.
Actas del IV Congreso Argentino de Paleontologı´a y Bioestrati-
grafı´a 3:125–130.
Roth, J. L., and D. L. Dilcher. 1978. Some considerations in leaf size
and leaf margin analysis of fossil leaves. Courier Forschungs-
lnstitut Senckenberg 30:165–171.
Rull, V. 1999. Palaeofloristic and palaeovegetational changes across
the Paleocene/Eocene boundary in northern South America. Re-
view of Palaeobotany and Palynology 107:83–95.
Schettino, A., and C. R. Scotese. 2001. New Internet software aids
paleomagnetic analysis and plate tectonic reconstructions. EOS
electronic supplement. Vol. 82. http://www.agu.org/eos_elec/
010181e.html.
Schneider, C. J., T. B. Smith, B. Larison, and C. Moritz. 1999. A test
of alternative models of diversification in tropical rainforests:eco-
logical gradients vs. rainforest refugia. Proceedings of the National
Academy of Sciences of the USA 96:13869–13873.
Smith, G. A., S. R. Manchester, M. Ashwill, W. C. McIntosh, and R.
M. Conrey. 1998. Late Eocene–early Oligocene tectonism, volca-
nism, and floristic change near Gray Butte, central Oregon. Geo-
logical Society of America Bulletin 110:759–778.
Smith, M. E., B. Singer, and A. Carroll. 2003.
40
Ar/
39
Ar geochronology
of the Eocene Green River Formation, Wyoming. Geological So-
ciety of America Bulletin 115:549–565.
Smith, T. B., C. J. Schneider, and K. Holder. 2001. Refugial isolation
versus ecological gradients. Genetica 112/113:383–398.
Steiger, R. H., and E. Ja¨ger. 1977. Subcommission on geochronology:
convention on the use of decay constants in geo- and cosmo-
chronology. Earth and Planetary Science Letters 36:359–362.
Tipper, J. C. 1979. Rarefaction and rarefiction: the use and abuse of
amethodinpaleontology.Paleobiology5:423434.
Trav e rso , N . E. 19 6 4 . La ep i der m is de Ginkgo patagonica Berry, del
Terciario de El Mirador, Provincia del Chubut. Ameghiniana 3:
163–166.
Tuom ist o , H., K . R uok o la i nen , a nd M. Y li- H all a . 20 0 3. Di spe r sa l ,
environment, and floristic variation of western Amazonian forests.
Science 299:241–244.
Van der Hammen, T., and H. Hooghiemstra. 2000. Neogene and
Quaternary history of vegetation, climate, and plant diversity in
Amazonia. Quaternary Science Reviews 19:725–742.
Viana, M. J., and J. A. Haedo-Rossi. 1957. Primer hallazgo en el
hemisferio sur de Formicidae extinguidos y ca´ talogo mundial de
los Formicidae fo´siles. Primera parte. Ameghiniana 1:108–113.
Villagra´n, C., and L. F. Hinojosa. 1997. History of the forests of
southern South America. II. Phytogeographical analysis. Revista
Chilena de Historia Natural 70:241–267.
Wilf, P. 1997. When are leaves good thermometers? a new case for
leaf margin analysis. Paleobiology 23:373–390.
Wilf, P., and K. R. Johnson. 2004. Land plant extinction at the end
of the Cretaceous: a quantitative analysis of the North Dakota
megafloral record. Paleobiology 30:347–368.
Wilf, P., S. L. Wing, D. R. Greenwood, and C. L. Greenwood. 1998.
Using fossil leaves as paleoprecipitation indicators: an Eocene ex-
ample. Geology 26:203–206.
Wilf, P., N. R. Cu´ neo, K. R. Johnson, J. F. Hicks, S. L. Wing, and J.
D. Obradovich. 2003. High plant diversity in Eocene South Amer-
ica: evidence from Patagonia. Science 300:122–125.
Willig, M. R., D. M. Kaufm an, and R. D. Stevens. 2003. Latitud inal
gradients of biodiversity: pattern, process, scale, and synthesis.
Annual Review of Ecology, Evolution, and Systematics 34:273–
309.
Wilson, M. V. H. 1980. Eocene lake environmen ts: depth and distance-
from-shore variation in fish, insect, and plant assemblages. Pa-
laeogeography, Palaeoclimatology, Palaeoecology 32:21–44.
Wing, S. L. 1998. Late Paleocene–early Eocene floral and climatic
change in the Bighorn Basin, Wyoming. Pages 380–400 in M.-P.
Aubry, S. Lucas, and W. A. Berggren, eds. Late Paleocene–early
Eocene climatic and biotic events in the marine and terrestrial
records. Columbia University Press, New York.
Wing, S. L., and D. R. Greenwood. 1993. Fossils and fossil climate:
the case for equable continental interiors in the Eocene. Philo-
sophical Transactions of the Royal Society of London B 341:243–
252.
Wing, S. L., F. Herrera, and C. A. Jaramillo. 2004. A Paleocene flora
from the Cerrejon Formation, Guajira Peninsula, northeastern Co-
lombia. International Organization of Paleobotany, Seventh Qua-
drennial Conference, Bariloche, Argentina, Abstracts, pp. 146–147.
Wolfe, J. A. 1979. Temperature parameters of humid to mesic forests
of eastern Asia and relation to forests of other regions in the
Northern Hemisphere and Australasia. U.S. Geological Survey Pro-
fessional Paper 1106:1–37.
———. 1985. Distribution of major vegetational types during the
Tertiary. Pages 357–375 in E. T. Sundquist and W. S. Broeker, eds.
The carbon cycle and atmospheric CO
2
:naturalvariationsArchean
to present. Geophysical Monograph 32. American Geophysical
Union, Washington, DC.
Wolfe, J. A., and W. C. Wehr. 1987. Middle Eocene dicotyledonous
plants from Republic, northeastern Washington. U.S. Geological
Survey Bulletin 1597:1–25.
Zachos, J. C., L. D. Stott, and K. C. Lohmann. 1994. Evolution of
early Cenozoic marine temperatures. Paleoceanography9:353–387.
Zachos, J. C., M. Pagani, L. C. Sloan, E. Thomas, and K. Billups.
2001. Trends, rhythms, and aberrations in global climate 65 Ma
to present. Science 292:686–693.
Ziegler, A. M., G. Eshel, P. M. Rees, T. A. Rothfus, D. B. Rowley,
and D. Sunderlin. 2003. Tracing the tropics across land and sea:
Permian to present. Lethaia 36:227–254.
Editor: Jonathan B. Losos
Associate Editor: Mark Westoby
This content downloaded from 128.253.177.74 on Thu, 25 Apr 2013 09:52:06 AM
All use subject to JSTOR Terms and Conditions
... The depositional settings for the modern leaf assemblages are as follows: 1 p wet lakeshore; 2 p understory leaf litter; 3 p grab sample of leaf pack from point bar; 4 p litter collection from a 1-m 2 area on the shore of a major channel bar; 5 p litter collection from a 0.5-m 2 area on a forested island; 6 p forest floor leaf litter collection from a 1-m 2 area. CURRANO ET AL.-AFRICAN PLANT DIVERSITY 557 fossil leaf floras ever studied (Wilf et al. 2003(Wilf et al. , 2005. Wilf et al. (2005) conducted macrofossil censuses at five stratigraphic levels, and rarefaction analyses show that Chilga is within error of the two most diverse levels ( fig. 6). ...
... CURRANO ET AL.-AFRICAN PLANT DIVERSITY 557 fossil leaf floras ever studied (Wilf et al. 2003(Wilf et al. , 2005. Wilf et al. (2005) conducted macrofossil censuses at five stratigraphic levels, and rarefaction analyses show that Chilga is within error of the two most diverse levels ( fig. 6). Nevertheless, Laguna del Hunco is a high-latitude lacustrine site with lower mean annual paleotemperature and paleoprecipitation than estimated for Chilga; a comparison between Chilga and a contemporaneous macroflora from the tropical low latitudes of South America (Burgess et al. 2004). ...
... Thus, the vicissitudes of environmental change and the presence of landscape heterogeneity likely led to both extinction and speciation, as they certainly did in other tropical regions. (Wilf et al. 2005). Chilga and Mush are denoted using solid lines and the Laguna del Hunco stratigraphic levels using dashed lines. ...
... Correlation scheme integrating geochronological information for PAT and AP Gondwanan Episode (Mazzoni et al., 1991;Gosses et al., 2006;Dunn et al., 2013;Montes et al., 2012Montes et al., , 2013Clyde et al., 2014;Krause et al., 2017). We include lithostratigraphic units (Raigemborn et al., 2010;Krause and Piña, 2012;Bellosi and Krause, 2014), Patagonian fossil plant associations (Wilf et al., 2003(Wilf et al., , 2005(Wilf et al., , 2010, vertebrate assemblages (Bond et al., 1995;Goin et al., 2012a;Reguero et al., 2013;Woodburne et al., 2014aWoodburne et al., , 2014bGelfo et al., 2015), and the global deep-sea oxygen isotope and paleoclimatic record (Zachos et al., 2001(Zachos et al., , 2008. Abbreviations: AP, Antarctic Peninsula; CH, Chubut; MR, Magallanes Region; PAT, Patagonia; PETM, Paleocene-Eocene Thermal Maximum; RN, Río Negro. ...
... We emphasize subsidence south of South America in the Tierra del Fuego region followed by narrowing in response to closure of former seaways due to tectonic uplift of the North Scotia Ridge and of the Fuegian and Patagonian Cordillera. warm and probably moderately humid climate (Wilf et al., 2005(Wilf et al., , 2009). The faunas from near Paso del Sapo in Chubut (~49-47 Ma), which are temporally and geographically close to the Laguna del Hunco and Río Pichileufú rainforest floras, contain metatherians, placental mammals, as well as one of the last gondwanatheres from SouthAmerica (Goin et al., 2012b). ...
Article
The Mesozoic plate tectonic and paleogeographic history of the final break up of West Gondwana had a profound effect on the distribution of terrestrial vertebrates in South America. As the supercontinent fragmented into a series of large landmasses (South America, Antarctica, Australia, New Zealand, the Indian subcontinent, and Madagascar), particularly during the Late Jurassic and Cretaceous, its terrestrial vertebrates became progressively isolated, evolving into unique faunal assemblages. The episodic nature of South American mammalian Cenozoic faunas became apparent in its modern formulation after George Gaylord Simpson’s seminal works on this topic. Two aspects add complexity to this generally accepted scheme: first, the fact that South America is not (and was not) a biogeographic unit, as the Neotropical Region does not include its southernmost tip (the Andean Region, including Patagonia and the southern Andes). Second, and intimately linked with the first one, that South America was not an island continent during the Late Cretaceous and the beginning of the Cenozoic, being its southernmost portion closely linked with West Antarctica up to the late Paleocene at least. Here we stress on this second aspect; we summarize a series of recent, detailed paleogeographical analyses of the continental breakup between Patagonia (including the Magallanes Region) and the Antarctic Peninsula crustal block, beginning with the opening of the Atlantic Ocean in the Early Cretaceous and running up to the Early Paleogene with the expansion of the Scotia Basin. In second place, we comment on the implications of these distinct paleogeographic and paleobiogeographic scenarios (before and after their geographic and faunistic isolation) for the evolution of South American terrestrial mammalian faunas. Summarizing, (1) we recognize a West Weddellian terrestrial biogeographic unit with the assemblage of the southern part of South America (Patagonia and the Magallanes Region) and the Antarctic Peninsula (and probably Thurston Island) crustal block of West Antarctica, spanning from the Late Cretaceous (Campanian) through the Early Paleogene (Paleocene); (2) we suggest that the Antarctic Peninsula acted as a double "Noah’s Ark” regarding, first, the probable migration of some non-therian lineages into southern South America; later, the migration of metatherians to Australasia.
... Consequently, significant transitions between high-temperature environments, such as greenhouse periods, and colder conditions, like glacial periods, have profoundly impacted global diversity patterns throughout history (Jaramillo et al., 2006;Provan & Bennett, 2008). For example, the early Paleogene greenhouse conditions supported highly diverse communities in both tropic and temperate regions (Fernández et al., 2021;Jaramillo et al., 2006), particularly during the Early Eocene Climatic Optimum (52-47 Ma) of Colombia and Patagonia (Jaramillo et al., 2006(Jaramillo et al., , 2010Wilf et al., 2003Wilf et al., , 2005Zachos et al., 2001). The shift toward colder conditions by the end of the Paleogene, associated with the onset of glaciation in Antarctica (∼34-33 Ma), meant a decrease in species richness in conjunction with a sharpening of the equator-pole climatic gradient (Coxall et al., 2005;Hinojosa & Villagrán, 1997;Jaramillo et al., 2006;Westerhold et al., 2020). ...
Article
Full-text available
The changing climate during the Cenozoic affected the diversity of plants in Patagonia, as species richness tends to increase during warm periods and decrease during cold periods. Precipitation is a significant factor shaping diversity, as shown in the case of central Chile during the Miocene. This study presents a reconstruction of the climate and vegetation in Tierra del Fuego Island, located approximately 52°S, using fossil flora recovered from the Filaret Formation to understand the Miocene epoch, characterized by contrasting global climatic changes. Filaret flora comprises twenty‐seven morpho‐taxa, including nine Nothofagus species and other Gondwanan and Neotropical families, such as Atherospermataceae and Anacardiaceae, in agreement with a forest habitat. Leaf physiognomy climate reconstruction suggests microthermal conditions, with a mean annual temperature of 9.4–11°C and annual precipitation ranging from 985 to 1,130 mm. These conditions are warmer and wetter than the modern record of the area, with a MAT of 6°C and mean annual precipitation of 300 mm. As the Filaret fossil record suggests, the forest habitat under a microthermal climate is consistent with the global climatic reconstruction of the Early Miocene. This Miocene landscape on Tierra del Fuego was possible because the Andes could not rain‐shadow humid westerly winds by this timeframe.
... Early Eocene sedimentation includes restricted regions with calderalake systems (for geographic and stratigraphic details see Gosses et al., 2021) that preserved diverse floras (e.g. Romero and Hickey, 1976;Gandolfo et al., 1988Gandolfo et al., , 2011Gandolfo and Hermsen, 2017;Wilf et al., 2005Wilf et al., , 2009Wilf, 2012;Zamaloa et al., 2006Zamaloa et al., , 2020Jud et al., 2018). Mid-late Eocene Atlantic incursions recorded mainly in southern Patagonia include shallow marine and highly fossiliferous sediments that have been constrained to the middle to late Eocene ages (Griffin, 1991;Malumián andCaramés, 1997,2000a;Camacho et al., 1998Camacho et al., , 2000aMalumián, 1999;Olivero and Malumián, 1999;Olivero et al., 2020;Guerstein and Junciel, 2001;Guerstein et al., 2014;González Estebenet et al., 2016;del Río, 2021). ...
Article
The fossil record from Cenozoic sediments provides a great deal of information that has direct bearing on the early assembling of modern Patagonian ecosystems. In this synthesis, we revise selected fossil marine and terrestrial records from the last 66 Ma with the aim of understanding major shifts of Patagonian biotas. From the Paleocene to the mid Eocene this region supported outstandingly diverse terrestrial assemblages that show strong connections to modern-day Australasia (e.g. gum trees, casuarinas, monotremes). Nearshore marine biotas confirm peak warmth conditions, with tropical species with Tethyan affinities. The late Eocene and early Oligocene marks the onset of a period of overall regional cooling, drying, and increasingly variable ecological conditions. The rise of palm-dominated flammable biomes in hinterlands and the prevalence of Gondwanan gallery forest (e.g. southern beeches and podocarps) along river-sides supported the existence of mosaic habitats maintained by edaphic and regional climatic conditions. This shift in landscapes reflects the evolution of a wide range of herbivorous mammals (e.g. Notoungulata, Litopterna, and Astrapotheria). The late Oligocene and early-to-mid Miocene witnessed a dramatic modification of landscapes including the incursion of high sea-level episodes, the emergence of specialized coastal (i.e. salt-marsh) plant taxa and the expansion of large herbivorous mammals with predominantly high-crowned teeth (e.g. Notoungulata: Hegetotheriidae, Interatheriidae, and Mesotheriidae). The cooling trend of this interval was interrupted by a mid-Miocene transient warming event, with the dispersion of terrestrial (e.g. platyrrhine monkeys, palms) and marine (e.g. Tuberculodinium vancampoae) elements with tropical affinity into southernmost South American regions. Seasonally-dry conditions increased towards the end of the Miocene, yet subtropical species persisted either in terrestrial (e.g. malpighs, passion vines, capybaras), and marine (e.g. Subtropical and Caribbean molluscs) environments. The increasing aridity caused by the Andean uplift wiped out most of the forest species and promoted the diversification of open-habitat species; the emergence of the current grass-dominated Patagonian Steppe occurred later on, probably during the Quaternary.
... The predominant biome developed in Patagonia in the Jurassic was the Subtropical Optimum, and a Thermal Maximum at the Paleocene/ Eocene boundary (Zachos et al., 2001;westerhold et al., 2020; and references therein). For Patagonia, temperatures between 14-18°C and mean annual precipitations over 2000 mm were estimated (wilf et al., 2005Hinojosa et al., 2010;Krause et al., 2010Krause et al., , 2014. From the Early Cretaceous until the mid-Eocene, chelid and meiolaniform turtles dominated the continental cheloniofauna in today's Chubut Province. ...
Article
Full-text available
In this work we focus on the fossil record of turtles and tortoises from the Chubut Province, in Patagonia, Argentina. This record is the richest, most diverse, the longest and continuous in the country and one of the most important in the continent. In this work, we present and study new fossils from all known clades of turtles from the province, coming from targeted field campaigns, as well as from past investigations, placed in a comprehensive and detailed chronostratigraphic context. In Chubut, more than 241 occurrences of turtles from at least 223 collection points are known, spanning from the Toarcian (Jurassic) to the Tortonian (Miocene). we manage to fill some of the gaps in the fossil record, complete the anatomical knowledge of many taxa, explore new geographical areas, and present some important highlights. Among these, the record of the oldest pan-chelid turtles indicates the significant potential of the sedimentary deposits of Chubut. Furthermore, we discuss the changes in the diversity and faunal turnovers of the various turtle clades in this region during the last 180 Ma and across important events during the Cretaceous/Paleocene and Oligocene/Miocene boundaries.
... The Paleocene-Eocene transition is characterized by a generalized marine regression in Patagonia and consequently, early Eocene to middle Eocene sediments are rarely represented outside the Austral Basin (Malumián, 1999). Nonetheless, from northern to southern Patagonia some Eocene floras are preserved (Hünicken 1955;Troncoso et al. 2002;Wilf et al. 2005;Panti 2018). These Eocene floras had a higher diversity of Gondwanan components (Podocarpaceae, Casuarinaceae, Myrtaceae, Nothofagaceae, Proteaceae) than today's southern We present a new wood assemblage from the Eocene Huitrera Formation, near Corcovado, Patagonia. ...
Article
Eocene paleofloras of Patagonia are diverse and increasingly known. A new assemblage of fossil wood has been recovered from Eocene sediments in Corcovado, western Argentinean Patagonia. The lithological succession (formerly Arroyo Lyn Formation sensu Pesce 1979) is correlated with the Huitrera Formation. The specimens were mostly found embedded in sandstones and conglomerates. One-third of the assemblage are conifers and studied herein. We found four taxonomic types: Agathoxylon cf. antarcticum (Araucariaceae), Phyllocladoxylon antarcticum (Podocarpaceae), Podocarpoxylon dusenii (Podocarpaceae), and Cupressinoxylon hallei (Cupressaceae or Podocarpaceae). The presence of four taxonomic units among only 7 specimens suggests a significant conifer species richness in the assemblage. Araucariaceae and dominant Podocarpaceae are usually found in previously described conifer wood assemblages from the Eocene of Patagonia andAntarctica. The diversity of the conifer assemblage in Corcovado is very similar to that found at Laguna del Hunco (these two localities are 170 km distant), also from the Huitrera Formation. This is consistent with the proposal that the bearing sediments of both localities are from the same stratigraphic unit.
... During the early Eocene, megathermal and diverse floras composed mainly of angiosperms dominated Patagonia (Wilf et al., 2003(Wilf et al., , 2005Barreda & Palazzesi, 2007). Some of these floras like that of Pichileufú or Laguna del Hunco lack Nothofagaceae or they are marginal components (Berry, 1938;Wilf et al., 2003;VD Barreda pers. ...
Article
We compiled the numerous fossil records (486) assigned to Nothofagaceae including pollen grains (from surface sediments and continental and oceanic borehole cores), leaves, woods and reproductive structures from South America. All the records are revised and the latest systematic treatments and ages of the bearing strata of each record are followed. When possible, we proposed a subgeneric affinity to each record based on updated bibliography. Fossils of three (Nothofagus, Fuscospora and Lophozonia) of the four subgenera are found in similar proportions through time since the Late Cretaceous. Fossils with reliable affinity with subgenus Brassospora were not found in South America. Most of the records are concentrated in the southern tip of South America (Patagonia Region) and nearby areas. After a significant presence of Nothofagaceae in the Cretaceous, the family declined in diversity and abundance in the Palaeocene and then increased from the Eocene to the Miocene. In the Miocene, the records reach their maximum diversity and abundance, and Nothofagaceae usually dominate the assemblages of pollen, leaves and woods from Patagonia. Pliocene Nothofagaceae records are virtually absent, probably because sedimentary rocks of that age are rare in Patagonia. The fossil record for Nothofagaceae varies according to environmental turnover; when tropical/subtropical floras were present in Patagonia in the Palaeocene–early Eocene, Nothofagaceae contracted southwards and when open steppes developed in the Miocene of east Patagonia, Nothofagaceae contracted westward.
... However, only the fossil record provides empirical evidence on how biodiversity is affected by long-term climatic transitions, even during global warming events. For example, fossil floras are known to have peaked in diversity during earlier hyperthermal episodes either at low 4 or high 5 paleo-latitudes of the American continent. The MECO may have also influenced terrestrial biotas, yet the magnitude of this response remains largely unknown as most published data have traditionally focused on the marine realm; it is still unclear whether biotic diversity increased, whether turnovers were gradual or step-like or whether tropical immigrants were frequent at the highest latitudes during the MECO. ...
Article
Full-text available
A major climate shift took place about 40 Myr ago-the Middle Eocene Climatic Optimum or MECO-triggered by a significant rise of atmospheric CO 2 concentrations. The biotic response to this MECO is well documented in the marine realm, but poorly explored in adjacent landmasses. Here, we quantify the response of the floras from America's south-ernmost latitudes based on the analysis of terrestrially derived spores and pollen grains from the mid-late Eocene (~46-34 Myr) of southern Patagonia. Robust nonparametric estimators indicate that floras in southern Patagonia were in average~40% more diverse during the MECO than pre-MECO and post-MECO intervals. The high atmospheric CO 2 and increasing temperatures may have favored the combination of neotropical migrants with Gondwanan species, explaining in part the high diversity that we observed during the MECO. Our reconstructed biota reflects a greenhouse world and offers a climatic and ecological deep time scenario of an ice-free sub-Antarctic realm.
... 2) at a position above the highest productive levels for compression fossils (i.e., quarry LH6; Wilf et al. 2003). The Tufolitas LH are well dated from 40 Ar/ 39 Ar analyses of minerals in three different tuffs, all of them below the resting level of the trunk, and the presence of two paleomagnetic reversals (Wilf et al. 2003(Wilf et al. , 2005. A maximum sanidine 40 Ar/ 39 Ar age of 52:22 5 0:22 Ma for one tuff is considered most reliable (M. ...
Article
Full-text available
Premise of research. Winteraceae, a family within the Canellales, is composed of tropical trees and shrubs broadly distributed in the Southern Hemisphere. The family is found today in eastern Australia, New Zealand, Malesia, Oceania, Madagascar, and the Neotropics across a range of dry to wet tropical to temperate climate regions. The fossil record of woods related to the Winteraceae in the Southern Hemisphere is limited to the Late Cretaceous of the Antarctic Peninsula. Here, we present a detailed anatomical description of the secondary xylem of a well-preserved trunk from the early Eocene Laguna del Hunco site, Huitrera Formation, Patagonia (Chubut Province, Argentina), that is referable to a new species of the genus Winteroxylon (Gottwald) Poole and Francis. Methodology. The wood is preserved as a siliceous permineralization; it was sectioned using standard petrographic techniques and observed under both light and scanning electron microscopy. The anatomy was compared with that of extant and fossil species of Winteraceae. Pivotal results. The diagnostic anatomical features of Winteraceae preserved in the fossil include an absence of growth rings, a lack of vessels, tracheids that are rectangular in cross section with circular bordered pits, diffuse axial parenchyma, rays showing two distinct size ranges (uniseriate-biseriate or multiseriate, 3–15 cells wide), and the presence of heterocellular rays containing sclerotic nests, cells with dark contents, and oil cells. The new fossil species most resembles extant genera within the Zygogynum s.l. clade from Australasian and Malesian rainforests; its anatomy is very similar to that of the extant genus Bubbia. The new Patagonian Winteraceae fossil wood is characterized by the presence of sclerotic nests and oil cells in the rays, which differ from those of previously described species of Winteroxylon. Conclusions. On the basis of the distinctive characters preserved, we erect Winteroxylon oleiferum sp. nov. The new fossil is the first reliable macrofossil record of Winteraceae from South America, supporting the abundant palynological record of the family from the continent, and it is the oldest record of the Zygogynum s.l. clade, adding to the long list of southern biogeographic connections between South America and Australasia via Antarctica during the warm early Cenozoic. Keywords: wood anatomy, early Eocene, Huitrera Formation, Winteraceae, Winteroxylon, Zygogynum s.l. clade.
Article
Full-text available
The most common macrofossils in the highly diverse flora from Laguna del Hunco (early Eocene of Chubut, Argentina) are "Celtis" ameghinoi leaves, whose true affinities have remained enigmatic for a century. The species accounts for 14% of all plant fossils in unbiased field counts and bears diverse insect-feeding damage, suggesting its high biomass and paleoecological importance. The leaves have well-preserved architecture but lack cuticles or reproductive attachments. We find that the fossils only superficially resemble Celtis and comparable taxa in Cannabaceae, Ulmaceae, Rhamnaceae, Malvaceae, and many other families. However, the distinctive foliar morphology conforms in detail to Dobinea (Anacardiaceae), a genus with two species of shrubs and large herbs ranging from India's Far East and Tibet to Myanmar and central China, and we propose Dobineaites ameghinoi (E.W. Berry) gen et. comb. nov. for the fossils. This discovery strengthens the extensive biogeographic links between Eocene Patagonia and mainland Asia, provides the first fossil record related to Dobinea, and represents a rare Gondwanan macrofossil occurrence of Anacardiaceae, which was widespread and diversified in the Northern Hemisphere at the time. The diverse leaf architecture of Anacardiaceae includes several patterns usually associated with other taxa, and many other leaf fossils in this family may remain misidentified.
Article
The Lower Eocene floras of southeastern North America were first published as a comprehensive flora by E. W. Berry in 1916 and later revised by him in 1930. This flora is one of the largest and most completely studied Eocene floras in North America. However, reinvestigation, presently in progress, has resulted in several revisions. Improved methods of research, more detailed and inclusive study expanding into new areas of research, and the increased understanding that has resulted from continued study in previously established areas of research, as well as the recent remapping of important deposits and the continued collection of fossil material, have provided the tools and information necessary for this revision. The age of these deposits has been revised from Wilcox group (Lower Eocene) to Claiborne group (Middle Eocene) and the nature of the deposits may now best be considered to be leaf-bearing clays laid down in ancient oxbow lakes. Recent studies of the cuticular remains of some leaves and pollen indicate that the clay pits are not isochronous but span the range of time of the Middle Eocene. Work which has been completed indicates that at least 60% of the taxonomic relationships of fossil forms to modern families and genera published by Berry are incorrect. When the fine venation and cuticular remains of the leaves are examined the presence of several extant taxa which Berry reported cannot be substantiated. Some taxonomic revisions are proposed; some of the fossil leaves studied could not be assigned to any known taxa and may represent extinct forms. The evolution of the angiosperms in the early Tertiary is often misunderstood because of the excessive number of extant generic and family names that are applied to fossil leaves with little or no detailed analysis of the fine venation or cuticular features of either modern or fossil angiosperms. The use of taxonomic affinities, community structure, and foliar physiognomy in making palaeoecological interpretations of early Tertiary floras is mentioned and the climate during Middle Eocene time in western Kentucky and Tennessee is reevaluated. The climate appears to have been dryer and somewhat cooler than previous investigators indicated.
Article
Precise estimates of past temperatures are critical for understanding the evolution of organisms and the physical biosphere, and data from continental areas are an indispensable complement to the marine record of stable isotopes. Climate is considered to be a primary selective force on leaf morphology, and two widely used methods exist for estimating past mean annual temperatures from assemblages of fossil leaves. The first approach, Leaf Margin Analysis, is univariate, based on the positive correlation in modern forests between mean annual temperature and the proportion of species in a flora with untoothed leaf margins. The second approach, known as the Climate-Leaf Analysis Multivariate Program, is based on a modern data set that is multivariate. I argue here that the simpler, univariate approach will give paleotemperature estimates at least as precise as the multivariate method because (1) the temperature signal in the multivariate data set is dominated by the leaf-margin character; (2) the additional characters add minimal statistical precision and in practical use do not appear to improve the quality of the estimate; (3) the predictor samples in the univariate data set contain at least twice as many species as those in the multivariate data set; and (4) the presence of numerous sites in the multivariate data set that are both dry and extremely cold depresses temperature estimates for moist and nonfrigid paleofloras by about 2°C, unless the dry and cold sites are excluded from the predictor set. New data from Western Hemisphere forests are used to test the univariate and multivariate methods and to compare observed vs. predicted error distributions for temperature estimates as a function of species richness. Leaf Margin Analysis provides excellent estimates of mean annual temperature for nine floral samples. Estimated temperatures given by 16 floral subsamples are very close both to actual temperatures and to the estimates from the samples. Temperature estimates based on the multivariate data set for four of the subsamples were generally less accurate than the estimates from Leaf Margin Analysis. Leaf-margin data from 45 transect collections demonstrate that sampling of low-diversity floras at extremely local scales can result in biased leaf-margin percentages because species abundance patterns are uneven. For climate analysis, both modern and fossil floras should be sampled over an area sufficient to minimize this bias and to maximize recovered species richness within a given climate.
Article
A classification of the architectural features of dicot leaves—i.e., the placement and form of those elements constituting the outward expression of leaf structure, including shape, marginal configuration, venation, and gland position—has been developed as the result of an extensive survey of both living and fossil leaves. This system partially incorporates modifications of two earlier classifications: that of Turrill for leaf shape and that of Von Ettingshausen for venation pattern. After categorization of such features as shape of the whole leaf and of the apex and base, leaves are separated into a number of classes depending on the course of their principal venation. Identification of order of venation, which is fundamental to the application of the classification, is determined by size of a vein at its point of origin and to a lesser extent by its behavior in relation to that of other orders. The classification concludes by describing features of the areoles, i.e., the smallest areas of leaf tissue surrounded by veins which form a contiguous field over most of the leaf. Because most taxa of dicots possess consistent patterns of leaf architecture, this rigorous method of describing the features of leaves is of immediate usefulness in both modern and fossil taxonomic studies. In addition, as a result of this method, it is anticipated that leaves will play an increasingly important part in phylogenetic and ecological studies.
Article
Rarefaction is a method for comparing community diversities that has consistently been abused by paleoecologists: here its assumptions are clarified and advice given on its application. Rarefaction should be restricted to comparison of collections from communities that are taxonomically similar and from similar habitats: the collections should have been obtained by using standardised procedures. The rarefaction curve is a graph of the estimated species richness of sub-samples drawn from a collection, plotted against the size of sub-sample: it is a deterministic transform of the collection's species-abundance distribution. Although rarefaction curves can be compared statistically, it may be more efficient to compare the species-abundance distributions directly. Both types of comparison are discussed in detail.
Article
Late Paleocene/early Eocene pollen and spore data taken from the US Gulf Coast (paleolatitude 32°N), western interior basins (Wyoming, North Dakota; paleolatitude 44-47°N) and Canadian Arctic (paleolatitude > 68°N) represent a vegetation proxy for ancient paratropical, subtropical and temperate biomes. These data provide information on the latitudinal diversity gradient of plants during an ancient greenhouse climate with non-freezing winters at polar latitudes. Comparing pollen data from the early Paleogene with a pollen data set compiled at the same latitudes from the late Holocene (3000 years B.P. to present) reveals that the diversity gradient between middle to high latitudes was steeper than today at the same sampling intensity. The gradient is a step-like decrease of about 50% in taxonomic diversity with increasing latitude between regions. The diversity gradient is formed by the 'spillover' of paratropical taxa into other regions of North America, which reflects the modern pattern of plant ranges. Taxa present in the Arctic, therefore, have great geographic ranges with endemism greatest in the paratropical biome. Paleogene diversity gradients show that decreasing diversity with increasing latitude is ancient and not dependent upon freezing temperatures.