+ All Categories
Home > Documents > Late Miocene climate and life on land in Oregon within a ...€¦ · Clarendonian (12 Ma) fossil...

Late Miocene climate and life on land in Oregon within a ...€¦ · Clarendonian (12 Ma) fossil...

Date post: 20-Oct-2020
Category:
Upload: others
View: 27 times
Download: 0 times
Share this document with a friend
27
Late Miocene climate and life on land in Oregon within a context of Neogene global change Gregory J. Retallack * Department of Geological Sciences University of Oregon, Eugene, OR 97403-1272, United States Received 13 November 2003; received in revised form 2 July 2004; accepted 20 July 2004 Abstract Clarendonian (12 Ma) fossil soils, plants, molluscs, fish, and mammals of eastern Oregon allow reconstruction of late Miocene paleogeography, paleoclimate, and paleoecology on land between the global thermal maximum of the middle Miocene (16 Ma) and global cooling and drying of the late Miocene (7 Ma). Six different pedotypes of paleosols recognized near Unity and Juntura allow reinterpretation of local mammalian paleoecology. Fossil beavers dominated gleyed Entisols of riparian forest. Abundant camels and common hipparionine horses dominated Alfisols of wooded grassland and grassy woodland. Diatomites overlying mammal-bearing beds have bullhead catfish [Ictalurus (Ameiurus ) vespertinus ], as well as fossil leaves dominated by live oak (Quercus pollardiana). Fossil plants and soils of Unity and Juntura are most like those of grassy live oak woodland and savanna on the western slopes of the Sierra Nevada in northern California today. Fossil plants and soils indicate mean annual temperature of 12.9 (7.7–17.7) 8C and mean annual precipitation of 879 (604–1098) mm. Miocene paleoclimatic changes in eastern Oregon show no relationship to changes in oxygen isotopic composition of marine foraminifera, usually taken as an index of global paleoclimatic change. Mismatch between land and sea paleoclimatic records is most likely an artefact of global ice volume perturbation of oxygen isotopic values. Instead, Miocene paleoclimatic change in eastern Oregon parallels changes in carbon isotopic composition of marine foraminifera, presumably through fluctuations in greenhouse gases. D 2004 Elsevier B.V. All rights reserved. Keywords: Miocene; Paleosols; Plants; Mammals; Oregon 1. Introduction Because Miocene soils, plants, and animals were close to our own time and modern in many respects, the surprisingly large amplitude of Miocene paleo- climatic change is of interest for understanding future global change. Evidence of global middle Miocene (16 Ma) warmth includes lateritic paleosols as far north and south as South Australia, Japan, Oregon, and Germany (Schwarz, 1997). Middle Miocene thermophilic trees such as sweetgum (Liquidambar pachyphylla ) grew as far north as Alaska (Wolfe and Tanai, 1980), and coconuts (Cocos zeylanica ) dispersed as far south as New Zealand (Fleming, 1975). Middle Miocene thermo- 0031-0182/$ - see front matter D 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.palaeo.2004.07.024 * Tel.: +1 541 3464558; fax: +1 541 3464692. E-mail address: [email protected]. Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97– 123 www.elsevier.com/locate/palaeo
Transcript
  • www.elsevier.com/locate/palaeo

    Palaeogeography, Palaeoclimatology, P

    Late Miocene climate and life on land in Oregon

    within a context of Neogene global change

    Gregory J. Retallack*

    Department of Geological Sciences University of Oregon, Eugene, OR 97403-1272, United States

    Received 13 November 2003; received in revised form 2 July 2004; accepted 20 July 2004

    Abstract

    Clarendonian (12 Ma) fossil soils, plants, molluscs, fish, and mammals of eastern Oregon allow reconstruction of late

    Miocene paleogeography, paleoclimate, and paleoecology on land between the global thermal maximum of the middle Miocene

    (16 Ma) and global cooling and drying of the late Miocene (7 Ma). Six different pedotypes of paleosols recognized near Unity

    and Juntura allow reinterpretation of local mammalian paleoecology. Fossil beavers dominated gleyed Entisols of riparian

    forest. Abundant camels and common hipparionine horses dominated Alfisols of wooded grassland and grassy woodland.

    Diatomites overlying mammal-bearing beds have bullhead catfish [Ictalurus (Ameiurus) vespertinus], as well as fossil leaves

    dominated by live oak (Quercus pollardiana). Fossil plants and soils of Unity and Juntura are most like those of grassy live oak

    woodland and savanna on the western slopes of the Sierra Nevada in northern California today. Fossil plants and soils indicate

    mean annual temperature of 12.9 (7.7–17.7) 8C and mean annual precipitation of 879 (604–1098) mm. Miocene paleoclimaticchanges in eastern Oregon show no relationship to changes in oxygen isotopic composition of marine foraminifera, usually

    taken as an index of global paleoclimatic change. Mismatch between land and sea paleoclimatic records is most likely an

    artefact of global ice volume perturbation of oxygen isotopic values. Instead, Miocene paleoclimatic change in eastern Oregon

    parallels changes in carbon isotopic composition of marine foraminifera, presumably through fluctuations in greenhouse gases.

    D 2004 Elsevier B.V. All rights reserved.

    Keywords: Miocene; Paleosols; Plants; Mammals; Oregon

    1. Introduction

    Because Miocene soils, plants, and animals were

    close to our own time and modern in many respects,

    the surprisingly large amplitude of Miocene paleo-

    climatic change is of interest for understanding

    0031-0182/$ - see front matter D 2004 Elsevier B.V. All rights reserved.

    doi:10.1016/j.palaeo.2004.07.024

    * Tel.: +1 541 3464558; fax: +1 541 3464692.

    E-mail address: [email protected].

    future global change. Evidence of global middle

    Miocene (16 Ma) warmth includes lateritic paleosols

    as far north and south as South Australia, Japan,

    Oregon, and Germany (Schwarz, 1997). Middle

    Miocene thermophilic trees such as sweetgum

    (Liquidambar pachyphylla) grew as far north as

    Alaska (Wolfe and Tanai, 1980), and coconuts

    (Cocos zeylanica) dispersed as far south as New

    Zealand (Fleming, 1975). Middle Miocene thermo-

    alaeoecology 214 (2004) 97–123

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–12398

    philic large foraminifera, corals, and molluscs

    extended as far north as Oregon, Alaska, Japan,

    and Kamchatkta (Moore, 1963; Itoigawa and Yama-

    noi, 1990; Oleinik and Marincovich, 2001), and as

    far south as South Australia and New Zealand (Li

    and McGowran, 2000; Fleming, 1975). These

    indicators of warmth disappeared from such high

    latitudes by later middle Miocene (15 Ma), when

    cold water minerals (glendonite pseudomorphs of

    ikaite) formed in marine rocks of Oregon (Boggs,

    1972). Benthic foraminiferal oxygen isotopic values

    also indicate a global middle Miocene marine

    thermal maximum, followed by late Miocene global

    cooling (Zachos et al., 2001).

    Fig. 1. Late Miocene fossil localities of Juntura, Ironside, and Unity, east

    Unity (Thayer, 1957; Thayer and Brown, 1973; Reef, 1983; Walker and M

    Tables 4–5; unnumbered localities have not yet yielded identifiable specim

    These global trends also were felt in eastern

    Oregon, where fossil leaves indicate marked cooling

    and drying between about 16 and 15 Ma (Wolfe et al.,

    1997; Graham, 1999), and a second cooling and

    drying to a modern flora by latest Miocene (7 Ma;

    Chaney, 1948). Paleosols of eastern Oregon show

    similar changes between middle Miocene mesic

    forests and late Miocene sagebrush steppe (Bestland

    and Krull, 1997; Retallack et al., 2002a; Sheldon,

    2003). Associated mammal faunas include a transition

    from three-toed, mixed-feeding horses of the middle

    Miocene (15 Ma, Barstovian) to modern monodactyl

    grazing horses by late Miocene (7 Ma, Hemphillian;

    Downs, 1956; Shotwell, 1970).

    ern Oregon, with detailed geological map and cross-section around

    acLeod, 1991). Numbered fossil localities and fossils are listed in

    ens.

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 99

    This study examines paleoclimate and paleo-

    ecology of the intervening part of the late

    Miocene within the context of Miocene global

    paleoclimatic extremes, using evidence from fossil

    soils, plants, molluscs, fish, and mammals in the

    eastern Oregon. Previous study of mammalian

    communities of this age from near Juntura

    (Shotwell, 1963; Shotwell and Russell, 1963),

    Ironside (Merriam, 1916), and Baker City (Downs,

    1952) are here supplemented with additional

    collections from near Unity, including reappraisal

    of a locality for a previously collected gompho-

    there skull and jaws (Orr and Orr, 1999). New

    Fig. 2. Measured section of the Ironside Formation in Windlass Gulch, 3

    and their development, calcareousness, and hue (using conventions of Re

    collections of fossil plants and fish are also

    reported, as well as new mapping of the area

    around Unity (Fig. 1) and an assessment of the

    paleoenvironmental significance of paleosols asso-

    ciated with these fossils (Figs. 2 and 3). The

    record of Neogene paleoclimatic fluctuation from

    these and other paleosols in eastern Oregon and

    Washington (Retallack, 1997b) does not correlate

    with oxygen isotopic records from deep-sea

    foraminifera but is in phase with foraminiferal

    variations in carbon isotopic composition (Zachos

    et al., 2001). This may indicate a closer relation-

    ship with the carbon cycle and greenhouse gases,

    miles northeast of Unity, Baker County, Oregon, showing paleosols

    tallack, 1997a, 2001b).

  • Fig. 3. Geological sections of fossil quarries excavated by Shotwell

    (1963) near Juntura, Oregon, showing paleosols and their develop-

    ment, calcareousness, and hue (using conventions of Retallack,

    1997a, 2001b).

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123100

    such as CO2 and CH4 (Retallack, 2002), than with

    deep-sea temperature or polar ice volume tracked

    by oxygen isotopic composition.

    2. Materials and methods

    Mapping and measurement of two stratigraphic

    sections near Unity, Baker County, Oregon (Figs.

    2 and 4) involved collection of 95 rock specimens

    and 62 fossils, which have been catalogued in

    collections of John Day Fossil Beds National

    Monument, Kimberly, Oregon (Scott Foss curator:

    catalog on line at http://www.museum.nps.gov/).

    Paleocurrents were sighted along trough cross bed

    axes using a Brunton compass. Two stratigraphic

    sections also were measured near Juntura, Mail-

    heur County, Oregon (Fig. 3) at fossil quarries of

    Shotwell (1963). Petrographic thin sections were

    cut of all the collected rocks in order to support

    field identifications of lithologies, and 36 thin

    sections were point-counted to determine paleosol

    textures and mineral composition (following the

    methods of Retallack, 1997a, 2001b). Clays and

    evaporite minerals were identified by X-ray

    diffraction using a Rigaku computer-automated

    goniometer.

    Six distinct kinds of paleosols, or pedotypes, were

    recognized (Table 1) throughout the mapped area

    (Figs. 1 and 2) and also near Juntura (Fig. 3) and

    Ironside. Each pedotype represents a distinct ancient

    environment (Table 2). The pedotype names are from

    the Sahaptin Native American language (Rigsby,

    1965; Delancey et al., 1988) and are part of a wider

    classification of Oregon Cenozoic paleosols (Retal-

    lack et al., 2000). Descriptive and analytical data for

    these paleosols are presented in Fig. 5 and Table 1.

    Chemical analyses of 21 paleosol specimens by

    Bondar Clegg of Vancouver, Canada, used X-ray

    fluorescence of a borate-fused bead for major ele-

    ments, titration for ferrous iron, and gravimetry for

    loss on ignition. Bulk density for each specimen was

    determined using the clod method. Of 109 paleosols

    logged near Unity and 42 near Juntura, 45 had

    pedogenic carbonate, and 5 were chemically analyzed

    for paleoclimatic interpretation. These data are added

    to 754 pedogenic carbonate measurements and 92

    chemical analyses of other paleosols from Oregon and

    Washington ranging back in age to 28.7 Ma (Gus-

    tafson, 1978; Bestland and Krull, 1997; Tate, 1998;

    Retallack, 2004), in order to assess local Neogene

    paleoclimate and its correspondence with global

    paleoclimatic change (Zachos et al., 2001).

    3. Geological background

    Only rocks of the Clarendonian North Ameri-

    can Land Mammal bAgeQ around Unity aredescribed and mapped in detail here (Figs. 1

    and 2). Few fossils have been recovered from

    poor exposures of the Ironside Formation near

    Ironside (Merriam, 1916) and Baker City, Oregon

    (Downs, 1952). The Unity, Ironside, and Baker

    City sequences and local faunas are similar to

    http://www.museum.nps.gov/

  • Fig. 4. Badlands of the late Miocene Ironside Formation in central Windlass Gulch (95–135 m of measured section Fig. 2), northeast of Unity. A

    gray airfall tuff is visible to the lower right and two thin white tuffs in the middle of these badlands.

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 101

    better-known sequences and faunas of the Juntura

    Formation near Juntura, Malheur County (Fig. 3;

    Bowen et al., 1963; Greene, 1973; Fiebelkorn et

    al., 1982; Johnson et al., 1996).

    3.1. John Day Formation (late Eocene–early

    Miocene)

    The oldest rocks exposed near Unity are 300 m of

    rhyodacitic tuff breccias, interpreted as lahars by Reef

    (1983). He considered them Eocene in age, based on

    fossil leaves (Rhamnidium chaneyi) and permineral-

    Table 1

    Paleosols of the Ironside and Juntura Formations and their classification

    Pedotype Sahaptin

    meaning

    Diagnosis

    Abiaxi Bitterroot Near-mollic surface (A) horizon over

    weakly weathered subsurface (Bw)

    Cmti New Shale and siltstone with root traces

    Monana Underneath Thin impure lignite (O) on hale and

    siltstone (A) with shallow root traces

    Skaw Scare Thin dark gray to black spotted siltstone

    with root traces (A) over bedded siltstone

    Tutanik Hair Near-mollic brown thick surface (A) with

    deep calcareous and siliceous

    rhizoconcretions and nodules (Bk)

    Xaus Root Sandstone with root traces (A) often

    ferruginized over bedded sandstone

    ized palm wood (Palmoxylon sp. indet.). However,

    Walker (1990) reports K–Ar ages from hornblende and

    plagioclase of 19.6F0.8 and 19.5F0.6 Ma, respec-tively (early Miocene), and regarded these rocks as

    outliers of the late Eocene to early Miocene John Day

    Formation. The laharic tuff breccia is overlain by a 100-

    m-thick rhyodacitic flow of similar petrographic and

    geochemical composition to the tuff breccia, additional

    tuff breccia, and local basalt flows (Reef, 1983;Walker,

    1990). These volcanic rocks are overlain by about 20 m

    of red claystone, including numerous paleosols of the

    Luca pedotype, which are late middle Eocene to early

    Australian

    classification

    (Stace et al., 1968)

    F.A.O.

    World Map

    (F.A.O., 1974)

    U.S. Soil

    Taxonomy

    (Soil Survey

    Staff, 1999)

    Brown clay Eutric Cambisol Haplosalid

    Alluvial soil Eutric Fluvisol Fluvent

    Humic gley Eutric Histosol Fibrist

    Wiesenboden Eutric Gleysol Mollic

    Endoaquent

    Brown earth Chromic Luvisol Typic

    Natrixeralf

    Alluvial soil Eutric Fluvisol Psamment

  • Table 2

    Paleosols of the Ironside and Juntura Formations and their interpretation

    Pedotype Paleoclimate Past vegetation Past animal life Paleotopographic

    setting

    Parent

    material

    Time of

    formation

    Abiaxi No indication Early successional

    wooded grassland

    cf. Cormohipparion

    sphenodus, Prosthennops,

    Gomphotherium osborni

    Riparian fringe

    and low terraces

    of floodplain

    Volcaniclastic

    silt

    1000–2000

    years

    Cmti No indication Riparian early

    successional

    woodland

    Proboscidea indet.,

    Merycodus

    Streamside levee

    deposits

    Volcaniclastic

    silt

    5–100

    years

    Monana No indication Swamp woodland None found Lake margin and

    floodplain swamp

    Volcaniclastic

    silt and

    diatomite

    100–1000

    years

    Skaw No indication Riparian floodway

    woodland:

    Gramineae,

    Equisetum,

    Populus, Salix

    Meterix, Scapanus,

    Hypolagus, Spermophilus

    junturensis, Diprionomys,

    Eucastor malheurensis,

    Peromyscus, Vulpes,

    Mammut furlongi,

    Prosthennops, Merychyus

    Seasonally

    waterlogged

    streamside swale

    Volcaniclastic

    silt

    50–1000

    years

    Tutanik Subhumid

    (800–1000 mm

    mean annual

    precipitation),

    summer-dry

    Grassy woodland Mylagaulus, Mammut

    furlongi, cf.

    Cormohipparion

    sphenodus, Aphelops,

    Procamelus grandis,

    Megatylopus

    Well-drained

    floodplain

    Volcaniclastic

    silt

    2000–5000

    years

    Xaus No indication Riparian early

    successional

    woodland

    Merycodus Streamside sand

    bars

    Volcaniclastic

    sand

    5–100

    years

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123102

    Oligocene (Uintan to Orellan) in age near Clarno and

    the Painted Hills, Oregon (Bestland et al., 1999;

    Retallack et al., 2000). Comparable red paleosols also

    are found in the upper John Day Formation east of

    Kimberly, Oregon, where they formed on well-drained

    uplands (Retallack et al., 2000).

    3.2. Strawberry and Dooley Volcanics (middle

    Miocene)

    Platy, 2-pyroxene, basaltic andesites are character-

    istic of middle Miocene (radiometric ages 15–10 Ma),

    Strawberry Volcanics, with smaller amounts of basalt,

    rhyolite, rhyolitic tuff, vent breccia, and micronorite

    (Robyn, 1977). The Strawberry Volcanics are calcalka-

    line and were not comagmatic with voluminous

    tholeiitic basalts of the Columbia River Basalt Group,

    which also are middle Miocene (radiometric age 17–15

    Ma; Sheldon, 2003). Calcalkaline volcanism is pro-

    duced by melting above subduction zones, but Straw-

    berry Volcanics and other coeval andesitic volcanoes

    (Walker, 1990) were far inland from subduction of the

    Farallon Plate during the middle Miocene (Orr et al.,

    1992). Middle Miocene Strawberry stratovolcanoes

    rose at least 1000 m above valley floors of the

    Paleozoic and Mesozoic basement (Robyn, 1977).

    Dooley Volcanics near Unity are 150 m thick and

    include three distinct units of rhyolitic, welded, ash

    flow tuffs (Reef, 1983). The ash flow tuffs are light

    gray with abundant pumice, rock fragments, and

    phenocrysts of plagioclase, biotite, and quartz.

    Dooley Volcanics reach a thickness of 2400 m

    around volcanic centers only a few kilometers to the

    north and northwest of the mapped area. One of the

    last eruptive events of the Dooley Volcanics was a

    rhyolite dome dated by K–Ar at 14.7F0.4 Ma(Evans, 1992).

    3.3. Ironside Formation (late Miocene)

    The Ironside Formation is at least 250 m of

    conglomerates, siltstones, diatomites and lignites

  • Fig. 5. Petrographic and chemical data for newly described paleosols of Tutanik and Xaus paleosols from 4 to 6 m, Tutanik, Cmti, Xaus, Skaw,

    and Abiaxi paleosols from 41 to 46 m, and Abiaxi, Cmti and Skaw paleosols from 229 to 232 m in reference section of Ironside Formation in

    Windlass Gulch (Fig. 2).

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 103

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123104

    named for the small village of Ironside, 15 km east

    on U.S. highway 26 from Unity (Thayer, 1957;

    Thayer and Brown, 1973). Sediments and paleosols

    along the highway east of Ironside are identical with

    those near Unity and particularly with the lower part

    of a measured section in Windlass Gulch (0–7 m in

    Fig. 2). Diatomites like those in the upper part of

    the Ironside Formation near Unity, also crop out in

    the hills north of Ironside. A fossil locality southeast

    of Baker City also exposes tuffs and conglomerates

    like those of the lower Ironside Formation (95–100

    m in Fig. 2) in Windlass Gulch (Downs, 1952;

    Evans, 1992).

    These three areas are linked by late Miocene

    fossil mammals of the Clarendonian North American

    Land Mammal "Age" (8.4–12.3 Ma of Prothero,

    Fig. 6. Late Miocene (Clarendonian) fossil mammals from Unity: (A)–(B

    and 9017, respectively); (C)–(D) Gomphotherium osborni, skull and mandi

    indet., molar with chipped crown (JODA9000). Scale bars as indicated.

    1998), although nomenclatural difficulties remain.

    For example, a fossil proboscidean skull from Unity

    on display at the Oregon Museum of Science and

    Industry in Portland (Fig. 6C–D) has remained

    undescribed beyond a photograph and account of

    its collection (Baldwin, 1964; Orr and Orr, 1999).

    The excavation is still visible within an Abiaxi

    paleosol on a low knoll southeast of a prominent

    bluff, east of Unity (locality 8 of Table 4; Walker,

    1990, figs. 5 and 9). George Gaylord Simpson

    examined the specimen in 1953 and suggested it was

    bMiomastodon merriamiQ (Orr and Orr, 1999). A moreappropriate name is bGomphotherium cingulatumQof Downs (1952), based on a lower jaw discovered

    by Albert Werner at a locality 15 km east of Baker

    City and about 80 km northeast of Unity. This and

    ) Hipparionini gen. indet., molar fragments (JODA specimens 9004

    ble (Oregon Museum of Science and Industry); (E) Prosthennops sp.

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 105

    other validly established taxa were not recognized

    by Lambert and Shoshani (1998), who considered

    the genus oversplit. Tobien (1972) went so far as to

    include all North American Gomphotherium in

    bGomphotherium productumQ. Here, I follow Mad-den and Storer (1985) in using their broadly defined

    Gomphotherium osborni (Barbour) for the Unity

    skull. Another gomphothere tooth, probably con-

    specific, was found by Elmer Molthan in the

    Ironside Formation 400 m south of Ironside Post

    Office, which is only 15 km east of Unity

    (bTetrabelodonQ sp. of Merriam, 1916). Gomphothe-

    Fig. 7. Fossil plants and fish from the Ironside Formation near Unity, Orego

    box elder, Acer negundoides, winged seed (JODA9409); (C) white o

    pollardiana leaf (JODA9058a); (E) shagbark hickory, Carya bendirei, l

    (JODA9050a); (G) bullhead catfish, Ictalurus (Ameiurus) vespertinus, sku

    rium and two additional proboscidean species are

    recognized from the Clarendonian Juntura Formation

    to the south in Malheur County (Shotwell and

    Russell, 1963): a two-tusker bMammut furlongiQ (aspecies not recognized by Lambert and Shoshani,

    1998), and a shovel-tusker similar to Platybelodon

    barnumbrowni (but regarded as a new genus by

    Lambert and Shoshani, 1998).

    There are also nomenclatural problems with

    Clarendonian horse teeth from Unity and Juntura,

    because among the numerous teeth and jaws, there is

    no skull with facial fossae or other diagnostic

    n: (A) sycamore, Platanus dissecta, leaf (specimen JODA9408); (B)

    ak, Quercus prelobata leaf (JODA9053); (D) live oak, Quercus

    eaf (JODA9055); (F) swamp ash, Fraxinus dayana, winged seed

    ll and partial skeleton (JODA9043a). Scale bars all 1 cm.

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123106

    features. The names given to these teeth (bHipparionanthonyiQ of Merriam, 1916 and bHipparion con-doniQ of Shotwell and Russell, 1963) are consideredby MacFadden (1984) likely junior synonyms of

    Cormohipparion occidentale or Cormohipparion

    sphenodus, the latter most likely on the basis of

    size. Shotwell and Russell (1963) considered that

    the large collection of teeth and jaws from

    Juntura represented only one species.

    Late Miocene geological age is also indicated by

    fossil plants from near Unity and Juntura. A new

    fossil plant assemblage in Juniper Gulch near Unity

    (Fig. 7; locality 6 of Table 4, at 160 m in Fig. 2) is

    here designated the Unity flora. Entire to weakly

    dentate oak leaves like those dominating this

    assemblage have been identified as bQuercushannibaliQ, but this name has been included withinQuercus pollardiana (Knowlton) by Axelrod (1995)

    or identified with the similar, but more dentate,

    living Quercus chrysolepis (Wolfe for Leopold and

    Wright, 1985). The Unity flora is taxonomically

    most like the late Miocene Stinkingwater flora in the

    Juntura Formation near Juntura (Chaney and Axel-

    rod, 1959) and the Weiser flora from the Poison

    Creek Formation of Idaho (Dorf, 1936; Leopold and

    Wright, 1985).

    The thick gray volcanic ash of the Ironside

    Formation at 171 m in Windlass Gulch is minera-

    logically most similar to the 11.31 Ma CPTXI Ash of

    the Great Basin (Perkins et al., 1998), and another

    gray tuff at 101 m is most like the 11.59 Ma CPTIX

    Ash (Figs. 2 and 3). A geological age of 11–12 Ma is

    similar to the age of the Juntura Formation, 78 km to

    the south in Malheur County. A whole-rock K–Ar

    date for 12-m-thick basalt 152 m stratigraphically

    below a fossil mammal quarry (OU locality 2337 of

    Fig. 3) and 85 m above the Stinkingwater flora

    (Chaney and Axelrod, 1959) near Juntura is 12.4 Ma

    (Evernden and James, 1964; corrected using Dal-

    rymple, 1979). A tuff 76 m stratigraphically above

    the mammal quarry has been dated by K–Ar on

    shards at 11.5F0.6 Ma (Fiebelkorn et al., 1982), andthe Devine Canyon ashflow tuff at the base of the

    Drewsey Formation, 107 m above the quarry, has

    been dated by 39Ar/40Ar of sanidine as 9.7 Ma

    (Johnson et al., 1996; five similar K–Ar dates are

    tabulated by Greene, 1973). Also comparable in age

    and lithology are volcanic ashes, conglomerates, and

    diatomites of the Coal Valley Formation of Nevada

    dated by the K–Ar technique at 11.4–9.4 Ma (Golia

    and Stewart, 1984).

    4. Late Miocene paleogeography of eastern Oregon

    4.1. Sedimentological evidence

    The depositional basin of the Ironside Forma-

    tion was bounded by volcanic terranes in the

    Strawberry and Dooley Volcanics to the north and

    west, and fault blocks of Paleozoic and Mesozoic

    basement to the north and southwest. There was

    erosional paleorelief on the underlying Dooley

    Volcanics and Clarno Formation, as indicated by

    several inliers cropping out south of Juniper

    Gulch. The stepped system of northeast-extensional

    faults (Fig. 1) did not exist during deposition,

    when this broad alluvial and lacustrine basin

    extended southeast from highlands to the north

    and west (Fig. 8).

    Basin-marginal boulder breccias and conglomer-

    ates interfinger with diatomites and tuffaceous

    siltstones of the Ironside Formation in road cuts

    over Dooley Mountain and north of Whited

    Reservoir (Fig. 1). Paleocurrent directions measured

    from trough cross bedding in conglomeratic paleo-

    channels in the Windlass Gulch section (Fig. 9)

    indicate that streams flowed southeast. This early

    phase of river channels and floodplains was

    followed by extensive oligotrophic lakes, which

    eventually accumulated diatomites out to the hilly

    margins of the depositional basin. A final deposi-

    tional phase of diminished lake area is recorded by

    fluvial gravels and lignites at the top of the exposed

    sequence near Unity (Fig. 2).

    4.2. Paleosol evidence

    Various paleosols represent different sedimentary

    environments and communities within these river

    valleys and lake shores. Weakly developed paleosols

    (Abiaxi, Cmti, Xaus) represent ephemeral commun-

    ities on the landscape, such as vegetation colonizing

    stream and lake margins after flooding. Weakly

    developed carbonaceous paleosols (Skaw) represent

    riparian or lake-margin marsh, and moderately

  • Fig. 8. Reconstructed late Miocene (Clarendonian) landscape ecology of the area near Unity, eastern Oregon.

    Fig. 9. Paleocurrents measured from trough cross bedding in

    conglomerates at 123–131 m (left) and 33–39 m (right) in the

    measured section in Windlass Gulch (Fig. 2).

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 107

    developed peaty paleosols (Monana) represent

    swamp woodlands, where decay of organic debris

    was suppressed by seasonally high water table.

    In contrast, Tutanik paleosols represent stable

    floodplains and alluvial terraces. Tutanik paleosols

    are thick, brown, and have little relict bedding due to

    bioturbation (including siliceous rhizoconcretions;

    Fig. 10A–B) and to development of soil structure

    (fine subangular blocky peds) and microfabric (skel-

    masepic porphyroskelic plasmic fabric of Brewer,

    1976). They are also more aluminous and less rich in

  • Fig. 10. Chalcedony nodules and rhizoconcretions from Tutanik paleosols of the Ironside Formation near Unity, Oregon: (A) petrographic thin

    section under crossed nicols of three rhizoconcretions (black holes with bright banded haloes) in oblique (lower left) and cross-sections

    (specimen JODA9225); (B) rhizoconcretion (JODA9311); (C) near-spherical nodule (JODA9310a); (D) lobed nodule (JODA9310b); (E) cigar-

    shaped nodule (JODA9309a); (F) tear-shaped nodule (JODA9310c); (G) cross-section of septarian nodule (JODA9310d); (H)–(I) petrographic

    thin sections under crossed nicols of dendritic chalcedony and void-filling dolomite (JODA9310d). Scale bars as indicated.

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123108

    alkalis and alkaline earths than associated paleosols

    (Fig. 5). Although volcanic ash was an important part

    of Tutanik paleosol parent material, few recognizable

    shards have escaped weathering. This degree of

    weathering is found after 8–27 ka in tropical highland

    soils of New Guinea (Ruxton, 1968). Tutanik

    paleosols have common clay skins in thin section

    and a higher proportion of fine pedogenic clay than

    associated paleosols (Fig. 5). The degree of enrich-

    ment of clay is intermediate between that seen in

    upper Modesto soils dated at 10 Ka and lower

    Modesto soils dated at 40 Ka in alluvium of the

    Merced River, San Joaquin Valley, California

    (Harden, 1982, 1990).

    4.3. Paleontological evidence

    Diatomites near Unity have not been studied in

    detail, but of 108 diatom taxa described from coeval

    diatomites of Juntura, only two are found in estuaries

    and the sea: bCoscinodiscus miocenicusQ andbCoscinodiscus subtilisQ of Van Landingham (1967;perhaps better referred to Actinocyclus according to

    Bradbury et al., 1985). Melosira granulata, Melosira

    distans, and Fragilaria construens dominate individ-

    ual collections (58–88%) and indicate fresh water of

    low alkalinity, low in dissolved solids, neutral to

    slightly acidic in pH, and low in nutrient levels

    (Bradbury et al., 1985). Significant (8%) amounts of

    C. subtilis in one sample (Van Landingham, 1967)

    may indicate connection to the sea by way of low-

    gradient streams (Bradbury et al., 1985). Fossil

    aquatic molluscs from Juntura and Unity (Tables 4–

    5; Taylor, 1985) are evidence for a Miocene course of

    the Snake River southwest into the Pit River of

    northern California, before Basin and Range faulting

    and northward capture by the Columbia River (Van

    TassellL et al., 2001). Lakes were also needed by

    many of the fossil birds (Brodkorb, 1961): cormor-

    ants, coots, mergansers, teals, and extinct straight-

    billed flamingos (Table 5).

    5. Late Miocene paleoclimate of eastern Oregon

    5.1. Evidence from paleosols

    Only Tutanik paleosols are useful paleoclimatic

    proxies, because other pedotypes were too weakly

    developed to show a paleoclimatic signature. Differ-

    ences between Tutanik and other comparably devel-

    oped Cenozoic paleosols from Oregon may reflect

    changing Neogene paleoclimate. Tutanik paleosols

    are generally similar in their brown color and clayey

    subsurface (argillic horizons) to Skwiskwi paleosols

    of the Oligocene John Day and latest Miocene

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 109

    Rattlesnake Formations (Retallack et al., 2000,

    2002a), and to early Miocene Tima paleosols of

    the John Day Formation (Retallack, 2004). Tutanik

    paleosols lack thick silcretes of Tima paleosols and

    have silica–micrite rhizoconcretions (Fig. 8A–B) and

    chalcedony septarian nodules (Fig. 8C–I) unknown

    in Skwiskwi paleosols. The large amount of silt

    (never less than 27% by volume) and sodium (soda/

    potash ratiosN1) also ally Tutanik paleosols with

    Tima rather than Skwiskwi paleosols. Tutanik

    paleosols also differ from paleosols of the late

    Miocene (7 Ma) Rattlesnake Formation, such as

    the Tatas pedotype, which was crumb-structured,

    moderately leached, and calcareous (Retallack et al.,

    2002a). Tutanik paleosols thus show greater affin-

    ities with interpreted summer-dry (xeric) paleosols,

    such as Tima, than summer-wet (ustic) paleosols,

    such as Swkiskwi and Tatas.

    Modern soils like Tutanik paleosols are found

    near Redding, northern California (map unit Lc3-2a

    with duric phase of F.A.O., 1975), where mean

    annual temperature is 17.7 8C, and mean annualprecipitation is 1040 mm, January and July temper-

    atures are 8 and 29 8C, respectively, and precip-itation is 216 mm in January, but only 5 mm in July

    (Ruffner, 1985). In contrast, for Baker City near

    Unity, mean annual temperature is 7.6 8C, meanannual rainfall is 270 mm, temperature is more

    seasonal (�4 8C in January, 19 8C in July), butprecipitation is less seasonal (35 mm in January, 12

    mm in July; Ruffner, 1985).

    Independent estimates of former precipitation can

    be gained from chemical composition of clayey

    horizons and from carbonate in rhizoconcretions of

    Tutanik paleosols at depths of 79–91 cm. This

    depth (D in cm) can be used to estimate mean

    annual rainfall (P in mmF156) using a transferfunction derived from study of modern soils

    (Retallack, 1994), once allowance is made for burial

    compaction (Sheldon and Retallack, 2001) due to at

    least 60 m of overlying Pleistocene fanglomerate

    (Evans, 1992). A large database of chemical

    analyses of North American soil B horizons has

    been used to derive yet other transfer functions for

    mean annual precipitation (P in mmF182) andmean annual temperature (T in 8CF4.4) from twochemical measures: S, the molar ratio Na2O+K2O/

    Al2O3; and C, molar Al2O3/(Al2O3+CaO+Na2O)

    times 100 (following Sheldon et al., 2002), as

    follows:

    P ¼ 139:6þ 6:388D� 0:01303D2

    P ¼ 221e0:02C

    T ¼ � 18:516 Sð Þ þ 17:298:

    Paleoclimatic results of these calculations (Table 3)

    are 450–800 mm mean annual precipitation from

    depth to carbonate and 600–1100 mm from chemical

    composition of clayey horizons. Because the carbo-

    nate-depth transfer-function was designed for use with

    nodules (Retallack, 1994, 2000), rather than rhizo-

    concretions, these estimates are less likely to be

    reliable than the subhumid regime estimated from

    chemical composition of B horizons. Geochemical

    composition of paleosol parent rhyodacitic and

    andesitic airfall ash shows insignificant change

    through time compared with paleosol modification

    (Perkins et al., 1998; Retallack et al., 2000).

    A highly seasonal climate is indicated by concen-

    trically banded rhizoconcretions with silica and

    micrite in Tutanik paleosols (Fig. 10A–B). Season-

    ality is also indicated by the unusually abundant silt

    and soda in Tutanik paleosols, especially considering

    their likely mean annual precipitation (Table 3). In

    such rainy climates, there would not be so much salt

    or dust unless summers were very dry, and most rain

    and snow were during winter and spring. Furthermore,

    the scarce carbonate and distinctive silica–micritic

    rhizoconcretions (Fig. 10A–B) are like those found in

    modern soils of Mediterranean (summer-dry) climates

    (Chadwick et al., 1987, 1995). Siliceous rhizoconcre-

    tions from a few Tutanik paleosols (45, 59, and 69 m

    in Fig. 2) grade insensibly through cigar- and tear-

    shaped forms into ellipsoidal septarian nodules of a

    form not previously reported from paleosols (Fig.

    10C–F). Some of these contain fossil wood or lumpy

    root traces morphologically comparable to rhizobially

    nodulated roots. Others are hollow and cracked, with

    void-filling dolomite rhombs (Fig. 10G,I). Micro-

    fabric of the nodules is radially oriented, dendritic

    chalcedony, creating an external fabric reminiscent of

    cone-in-cone structure. The ellipsoidal nodules are

    superficially similar to chalcedony nodules of the

    bbutton bedsQ, a lacustrine facies of the middle

  • Table 3

    Mean annual temperature and precipitation for Tutanik paleosols of the Ironside Formation

    Estimated geological

    age (Ma in Fig. 5)

    Stratigraphic level

    (m in Fig. 3)

    Mean annual precipitation

    (mm) from Bk

    Mean annual precipitation

    (mm) from chemical

    composition

    Mean annual temperature

    (8C) from chemicalcomposition

    11.59 100.5 644F156 (503–785)11.60 Unity Reservoir 657F156 (516–798)11.69 83 588F156 (447–729) 916F182 (734–1098) 13.3F4.4 (8.9–17.7)11.83 45.1 601F156 (460–742) 934F182 (752–1116) 13.2F4.4 (8.8–17.6)11.95 16.6 643F156 (502–784)11.96 14 617F156 (476–758)11.97 12.5 626F156 (485–767)11.00 Ironside 640F156 (499–781)12.00 5.2 627F156 (486–768) 786F182 (604–968) 12.1F4.4 (7.7–16.5)12.01 3.3 636F156 (495–777)12.02 1.7 649F156 (508–790)12.10 Denny Flat 571F156 (430–712)12.11 Denny Flat 594F156 (453–735)Means All listed 623F156 879F182 12.9F4.4

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123110

    Miocene Barstow Formation of California (Palmer,

    1957). My own examination of the Barstow nodules

    found conspicuous relict bedding and microcrystalline

    textures, unlike the septarian and radial-dendritic

    crystal form of the Unity nodules. Nevertheless,

    highly alkaline late-summer aridity could have played

    a role in the formation of both kinds of siliceous

    nodules.

    5.2. Evidence from fossil plants

    Close floristic composition of Unity (Table 5)

    and Stinkingwater floras (Chaney and Axelrod,

    1959) with the modern flora of California supports

    evidence from associated paleosols for extension of

    Mediterranean climate into eastern Oregon during

    the late Miocene. The paleoclimatic affinities of

    these floras are quite distinct from more humid and

    warm climates evident during the middle Miocene

    thermal maximum, represented locally by the

    Austin, Tipton, Baker, and Mascall floras (Oliver,

    1934; Brown for Gilluly, 1937; Chaney and

    Axelrod, 1959), and before the advent of modern

    sagebrush and riparian communities during the

    Pliocene (Chaney, 1948; Ashwill, 1983, Retallack

    et al., 2002a). Mean annual precipitation of around

    900–1000 mm and mean annual temperature of 13

    (8–18) 8C have been estimated from other lateMiocene fossil floras of the Great Basin (Wolfe,

    1994; Wolfe et al., 1997; Graham, 1999). A

    summer-dry paleoclimatic regime for the Unity flora

    is also compatible with frequent wildfires indicated

    by charcoal associated with middle to late Miocene

    fossil floras of southeastern Oregon (Taggart et al.,

    1982; Taggart and Cross, 1990).

    5.3. Evidence from fossil animals

    Abundance of bullhead catfish in siltstones with

    the Unity flora (Fig. 7; Table 4) can be contrasted

    with the overlying diatomites, which have a sparse

    fish fauna of other kinds of fish. The leaves and

    catfish probably represent a nearshore lacustrine

    habitat distinct from that of the open lake. Bullhead

    catfish lived in western North America from the

    Eocene to Pliocene, but after that time became

    restricted to eastern North America. They have been

    artificially reintroduced to Oregon and Idaho within

    the last 120 years. Today, they live in low-gradient

    streams and lakes at elevations less than 1000 m in

    warm temperate climates with at least 230 frost-free

    days and at least 400 mm mean annual precipitation

    (Van Tassell et al., 2001).

    Brodkorb (1961) invokes Bergmann’s rule (Brown

    and Lomolino, 1998) to argue for a warmer than

    present climate from the small size of Juntura

    cormorants, coots, and teals compared with living

    relatives. Oxygen isotopic studies of fossil mammal

    teeth from Juntura, Oregon, indicate mean annual

    temperatures of about 18 8C, and a mean annual range

  • Table 4

    Late Miocene (Clarendonian) megafossils of the Ironside Formation, near Unity, Oregon

    Fossil taxon Similar living taxon Common name Loc.

    Equisetum sp. indet. Equisetum arvense Common horsetail 1–2

    Gramineae sp. indet. Gramineae Grass 1–2

    Sagittaria sp. indet. Sagittaria cuneata Arum-leaved arrowhead 3

    Salix sp. indet. Salix scouleriana Scouler’s willow 3

    Populus sp. indet. Populus balsamifera Black cottonwood 3

    Trapa americana Knowlton (1898) Trapa natans Water chestnut 3–5

    Platanus dissecta Lesquereux (1878) Platanus occidentalis Sycamore 6

    Quercus pollardiana (Knowlton) Axelrod (1995) Quercus chrysolepis Maul oak 6

    Quercus prelobata Condit (1944) Quercus garryana Oregon white oak 6

    Quercus simulata Knowlton (1898) Quercus myrsinaefolia Chinese oak 6

    Cercocarpus ovatifolius Axelrod (1985) Cercocarpus betuloides

    var. blancheae

    Birchleaf mountain

    mahogany

    6

    Acer negundoides MacGinitie (1934) Acer negundo Box elder 6

    Fraxinus dayana Chaney and Axelrod (1959) Fraxinus caroliniana Swamp ash 6

    Carya bendirei (Lesquereux) Chaney and Axelrod (1959) Carya ovata Shagbark hickory 6

    Ptelea miocenica Berry (1931) Ptelea trifoliata Carolina hop tree 7

    Radix junturae Taylor (1963) Radix auriculata Aquatic big-ear snail 7

    Ictalurus (Ameiurus) vespertinus Miller and Smith (1967) Ictalurus melas Black bullhead catfish 6

    Gomphotherium osborni (Barbour) Madden and Storer (1985) None Extinct four-tusker 8

    Prosthennops sp. indet. None Extinct large peccary 8

    Rhinocerotidae sp. indet. None Large extinct rhino 9

    cf. Cormohipparion sphenodus (Cope)

    MacFadden and Skinner (1977)

    None Extinct cursorial

    three-toed horse

    10

    cf. Megatylopus gigas Matthew and Cook (1909) None Extinct large camel 11

    cf. Merycodus sp. Antilocapra americana Pronghorn 11–12

    Note: these are identifcations of newly collected material in collections of John Day Fossil Beds National Monument. Legal and NAD27 UTM

    coordinates to localities are 1—Windlass Gulch section 60 m NWH NEH SEH SEH S33 T12S R38E 11T 0416253 4925216; 2—Windlass

    Gulch section 44 m NWH NEH SEH SEH S33 T12S R38E 11T 0416262 4925216; 3—Windlass Gulch section 61 m NWH NEH SEH SEH

    S33 T12S R38E 11T 0416262 4925216; 4—Denny Flat southeast SEH SEH SEH SWH S2 T13S R37E 11T 0409723 492339; 5—Denny flat

    southeast NEH NEH NEH NEH S2 T13S R37E 11T 0409761 492353; 6—Windlass Gulch section 158 m NWH NWH SWH SWH S34 T12S

    R38E 11T 0416526 495228; 7—Windlass Gulch section 225 m NEH NWH SWH S34 T12S R38E 11T 0416665 4925404; 8—east of Unity

    SEH NEH SEH NWH S11 T13S R37E 11T 0408804 4922516; 9—Windlass Gulch section 100 m NEH SWH SWH SWH S34 T12S R38E

    11T 0416596 4925185; 10—Windlass Gulch section 5 m SEH NWH SEH SEH S33 T12S R38E 11T 0416227 4925153; 11—Denny Flat

    northwest NWH SWH SWH SWH S35 T12S R37E 11T 0408416 4925224; 12—Windlass Gulch section 89 m NEH NEH SEH SEH S33 T12S

    R38M 11T 0416315 4925249.

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 111

    of temperature of some 25 8C (Kohn et al., 2002),compared with 7.6 and 23 8C, respectively, for BakerCity today (Ruffner, 1985).

    Indications of a late Miocene climate at Juntura,

    wetter than at present, come from the sewellel

    (Tardontia in Table 5). Similar aplodontids are

    now restricted to Oregon and Washington, west of

    the Cascades (Shotwell, 1958). The western Amer-

    ican mole (Scapanus) is also now found well west

    of Juntura and Unity (Hutchison, 1968). On the

    other hand, subhumid to semiarid rangeland con-

    ditions are indicated by ground squirrels, pocket

    gophers, and extinct mylagaulines (Shotwell, 1958,

    1970).

    6. Late Miocene palaeoecology of eastern Oregon

    6.1. Evidence from paleosols

    A detailed local mosaic of vegetation (Fig. 8)

    can be inferred from the relationship of paleosols to

    local sedimentary facies and from plant-induced

    features of the paleosols (Figs. 2 and 3; Table 3).

    Tutanik paleosols are finely structured (granular

    peds about 1 cm across) with abundant fine root

    traces as well as scattered rhizoconcretions, as in

    soils supporting grass with scattered woody shrubs

    or trees (Retallack, 1997b, 2001b). The paleosols

    lack fine crumb peds found in sod-forming grass-

  • Table 5

    Late Miocene (Clarendonian) megafossils of the Juntura Formation

    near Juntura

    Fossil taxon Common name

    Sphaerium sp. cf. S. lavernense

    Herrington (Herrington and

    Taylor, 1958)

    Aquatic lake orb

    mussel

    Pisidium sp. cf. P. clessini

    Neumayr and Paul (1875)

    Aquatic pea clam

    Pisidium sp. indet. Aquatic pea clam

    Fluminicola junturae Taylor, 1963 Aquatic pebble snail

    Hydrobiidae indet. Small aquatic snail

    Viviparus turneri Hannibal, 1912 River snail

    Radix junturae Taylor, 1963 Aquatic big-ear snail

    Carinifex shotwelli Taylor, 1963 Keeled ramshorn snail

    Promenetus sp. indet. Aquatic ramshorn snail

    Ictalurus peregrinus Lundberg, 1975 Catfish

    Cyprinidae gen. indet. Squawfish

    Catastomidae gen. indet. Sucker

    Phalacrocorax leptopus

    Brodkorb, 1961

    Small cormorant

    Ciconiidae gen. indet. Extinct stork

    Megapaloelodus opsigonus

    Brodkorb, 1961

    Straight-billed

    early flamingo

    Eremochen russelli

    Brodkorb, 1961

    Small goose

    Querquedula pullulans

    Brodkorb, 1961

    Small teal

    Ocyplonessa shotwelli

    Brodkorb, 1961

    Extinct merganser

    Neophrontops dakotensis

    Compton, 1935

    Extinct Old

    World vulture

    Fulica infelix Brodkorb, 1961 Small coot

    Meterix sp. indet. Extinct large shrew

    Anouroneomys minimus Hutchison

    & Bown (Bown, 1980)

    Extinct small shrew

    Mystipterus sp indet. Extinct nonburrowing

    mole

    Scalopoides sp. indet. A Extinct weakly

    burrowing mole

    Scalopoides sp. indet. A Extinct weakly

    burrowing mole

    Scapanus sp. cf. S. shultzi

    Tedford, 1961

    Western American

    mole

    Chiroptera gen et sp. indet. Extinct bat, like big

    brown bat

    Hesperolagomys sp. cf. H. galbreathi

    Dawson (Clark et al., 1964)

    Extinct pika

    Hypolagus sp. cf. H. fontinalis

    Dawson, 1958

    Extinct rabbit

    Tardontia sp. cf. T occidentale

    (Macdonald) Shotwell, 1958

    Extinct sewellel

    Mylagaulus sp. indet. Extinct horned rodent

    Epigaulus minor Hibbard and

    Phillis (1945)

    Extinct horned rodent

    Fossil taxon Common name

    Ammospermophilus junturensis

    (Shotwell & Russell) Korth, 1994

    Ground squirrel

    Ammospermophilus sp. indet. Ground squirrel

    Eutamias sp. indet. Extinct chipmunk

    Hystricops sp. indet. Extinct large beaver

    Eucastor malheurensis Shotwell

    and Russell (1963)

    Extinct beaver

    Microtoscoptes sp indet. Extinct vole

    Copemys dentalis (Hall)

    Lindsay, 1972

    Extinct small field

    mouse

    Copemys sp. cf. C. esmeraldensis

    (Wood) Lindsay, 1972

    Extinct large field

    mouse

    Copemys sp indet. Extinct large field

    mouse

    Leptodontomys sp. indet. Extinct pocket mouse

    Perognathus sp. indet. Pocket mouse

    Diprionomys sp. indet. Extinct pocket mouse

    Pliosaccomys sp. indet. Extinct pocket gopher

    Macrognathomys nanus Hall, 1930 Extinct birch mouse

    Borophagus sp. indet. Short-face bone-

    crushing dog

    Aelurodon sp. indet. Long-face bone-

    crushing dog

    Vulpes sp. indet. Fox

    Leptarctus sp. indet. Extinct badgerlike

    mustelid

    Hoplictis sp. indet. Extinct like honey

    badger

    Sthenictis junturensis Shotwell and

    Russell (1963)

    Extinct otterlike

    musteline

    Pseudailurus sp. indet. Extinct small cat

    Gomphotherium sp. indet. Extinct four-tusk

    proboscidean

    Mammut furlongi Shotwell and

    Russell (1963)

    Extinct two-tusk

    mastodon

    cf. Platybelodon barnumbrowni

    (Barbour) Barbour, 1932

    Extinct shovel-tusker

    cf. Cormohipparion sphenodus (Cope)

    MacFadden and Skinner (1977)

    Extinct three-toed

    horse

    Tapiridae gen. indet. Tapir

    Aphelops sp. indet. Extinct hornless

    rhinoceros

    Prosthennops sp. indet. Extinct large peccary

    Merychyus major Leidy, 1858 Extinct piglike oreodon

    Procamelus grandis Gregory, 1942 Extinct camel

    Megatylopus sp. indet. Extinct large camel

    Note: this is a compilation of the Black Butte local fauna of

    Shotwell (1963, 1967, 1970), Hutchison (1968), and Taylor (1963)

    as revised by Honey et al. (1998), Lambert and Shoshani (1998),

    Baskin (1998), Martin (1998), Lander (1998), and Wright (1998).

    Table 5 (continued)

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123112

    land soils (Mollisols, Phaeozems, Chernozems;

    Retallack et al., 2002a), so that grasses were most

    likely bunch grasses.

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 113

    Other paleosols with prominent relict bedding

    and large woody root traces supported plant

    formations early in ecological succession after

    disturbance. For example, Xaus paleosols of in-

    channel sandy bars probably supported riparian

    woodland, and Cmti and Abiaxi paleosols of levee

    siltstones probably supported riparian gallery wood-

    land. Cmti paleosols are thinner, less sodic, and

    have more relict bedding than Abiaxi paleosols (Fig.

    7), indicating a shorter period of vegetation growth

    and successional development. Early successional

    marsh ecosystems are represented by carbonaceous

    clayey paleosols with small root traces and relict

    bedding (Skaw pedotype). Swamp ecosystems are

    represented by lignitic paleosols with large woody

    root traces (Monana pedotype). The swamp paleo-

    sols have impure coals and charcoal and were

    probably seasonally dry, but there is no comparable

    evidence of seasonal exposure of marsh paleosols,

    which may have been associated with perennial

    watercourses.

    6.2. Evidence from fossil plants

    Only one kind of paleosol contained preserved

    fossil plants: weakly developed paleosols of the

    Skaw pedotype (localities 1–3 of Table 4). The best

    preserved fossil plants were found within laminated

    diatomites and siltstones (localities 4–7 of Table 4).

    These are interpreted as leaf litter of marsh soils

    (Skaw pedotype) and leaves blown into an oligo-

    trophic lake, respectively. In all cases, there is a bias

    toward aquatic plants, such as water chestnut,

    arrowhead, and horsetail.

    Despite these taphonomic limitations, the Stin-

    kingwater, Weiser, and Unity floras include locally

    abundant fossil grasses and live oak comparable to

    those of summer-dry parts of California today. The

    Unity flora is most similar to interior live oak

    woodland of Barbour and Major (1988) and canyon

    live oak woodland of Sawyer and Keeler-Wolf

    (1995). These oak woodlands are at moderate

    elevation below the zone of abundant conifers on

    the southwestern slopes of the Sierra Nevada

    flanking the northern California Great Valley. The

    oak trees are largely evergreen and up to 15 m high

    in both closed and open formation. Their grassy

    understorey is brown and dry for most of the

    summer. Mesic elements in the Unity flora (Carya,

    Ptelea) are not found in modern Californian live oak

    woodlands. These middle Miocene holdover taxa are

    evidence that late Miocene dry woodlands of Unity

    were not exactly like Californian vegetation today.

    No taxodiaceous conifers were found near Unity,

    where they would be expected considering charcoal

    and woody root traces in Monana paleosols. These

    swamp conifers are common in the coeval Stinking-

    water flora of Juntura (Chaney and Axelrod, 1959)

    and the Weiser flora of Idaho (Dorf, 1936).

    Fossil grasses are represented by sheathing leaves

    near Unity (Table 4) and by abundant pollen in coeval

    rocks of Idaho (Leopold and Wright, 1985). Miocene

    grasses of the Great Plains and Oregon are unlikely to

    have been floristically similar, given differences

    between fossil dicot floras of Oregon and the Great

    Plains (Thomasson, 1979; Graham, 1999; Strfmberg,2002) and differences between sodic–silicic Miocene

    paleosols of Oregon (Retallack, 2004) and calcareous

    Miocene paleosols of the Great Plains (Retallack,

    1997b, 2001b). A comparable modern vegetation

    distinction is the western wheat grass (Agropyron

    spicatum) rangeland province of northern California

    and the Great Basin, and blue grama (Bouteloua

    gracilis) grasslands of the western Great Plains

    (Leopold and Denton, 1987).

    6.3. Evidence from fossil animals

    Quarry collections from Juntura formed the basis

    for Shotwell’s (1963) pioneering study of Clarendo-

    nian mammalian paleocommunities (Fig. 11), and

    these quarries were revisited in order to examine

    their paleosols (Fig. 3). Shotwell’s (1963) bpondbank communityQ came from Skaw paleosols (theonly pedotype found), whereas his bsavannacommunityQ came largely from sandstones immedi-ately overlying a Tutanik paleosol. His bsavannacommunityQ was mainly camels, hipparioninehorses, two-tusker mastodon, and rhinoceros. The

    bpond bank communityQ, which from paleosol andsedimentological evidence appears more likely to

    have been riparian woodland, was mainly fish and

    turtles, but also had many beavers, as well as rabbits

    and a variety of rodents and insectivores (Table 5).

    The weak development and carbonaceous and

    manganiferous composition of Skaw paleosols are

  • Fig. 11. Specimen numbers of fossil mammals of Clarendonian

    bsavannaQ and bpond bankQ communities by Shotwell (1963), fromquarries of Fig. 3. Numerous fish and turtle bones from locality

    2337 are not included in these figures.

    G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123114

    evidence of seasonally inundated streamside swales

    (Table 2), where fish were trapped and devoured by

    predators.

    Contrary to Shotwell (1963), Bernor et al. (1988)

    concluded that coeval late Miocene (11 Ma)

    hipparionine horses of Austria were forest browsers,

    because forest leaves were found in associated lake

    beds. This view is not supported by the abundance

    of hipparionines in Tutanik paleosols and rarity in

    Skaw paleosols of Juntura and Unity (Fig. 11).

    Later studies of hipparionine tooth wear (Hayek et

    al., 1992) and associated mammalian fauna (Webb,

    1983) also support Shotwell’s (1963) conclusion that

    hipparionines were grassland and grassy woodland

    mixed feeders. Open grassy terrane was also

    required by an extinct species of Old World vulture

    from Juntura (Table 5).

    Proboscideans are now associated with mosaics of

    grassy and woody vegetation, which they maintain

    by destructive feeding (Owen-Smith, 1988). A

    diversity of proboscidean habitats may be expected

    from the high diversity of Clarendonian probosci-

    deans, including eight genera in North America and

    three in the Unity and Juntura faunas (Tables 4–5).

    Shovel-tuskers (Platybelodon) are commonly por-

    trayed as aquatic shovelers, although tusk wear

    patterns indicate scraping against bark, twigs, peb-

    bles, or other hard debris (Lambert, 1992). Four-

    tuskers (Gomphotherium) also show tusk wear from

    bark stripping and digging, and two-tusk mastodons

    (Mammut) are known from Pleistocene stomach

    contents to have eaten conifer needles, cones, and

    grass (Lambert and Shoshani, 1998). Associated

    paleosols are Abiaxi for Gomphotherium, Tutanik

    for Mammut, and probably Skaw for cf. Platybelo-

    don (uncertainty comes from poor exposure and

    inexact location). The inferred paleoenvironments of

    these paleosols (Table 3) cannot be claimed to

    represent habitat preferences based on only one or

    two specimens of each species.

    Paleosols also support the idea that late Miocene

    rangelands of the Pacific Northwest were distinct from

    those of the Great Plains (Leopold and Denton, 1987).

    Late Miocene summer-dry siliceous alfisol and aridisol

    paleosols of Oregon (Retallack et al., 2002a) can be

    contrasted with summer-wet Mollisols of Nebraska

    and Montana (Retallack, 1997b, 2001b). The hippar-

    ionine province of the Pacific Northwest was distinct

    from a Pliohippus province of southern California and

    the Great Plains during the Clarendonian (Shotwell,

    1963). Late Miocene (7 Ma) grazing Pliohippus in

    Oregon appeared in calcareous Mollisol paleosols

    reflecting an incursion of summer-wet Gulf Coast air

    masses into eastern Oregon at that time (7.3 Ma;

    Retallack et al., 2002a). The tridactyl mixed-feeding

    hipparionines may reflect rigors of summer-dry range-

    land vegetation botanically and structurally distinct

    from summer-wet rangeland of monodactyl grazing

    Pliohippus. Comparable differences in vegetation and

    seasonality may also explain the numerical dominance

    of foregut-fermenting ruminant camels over hindgut-

    fermenting horses at Juntura (Fig. 11) and other

    western Clarendonian faunas (Tedford and Barghoorn,

    1993), whereas horses dominated Clarendonian

    assemblages of Nebraska and Texas (Webb, 1983).

    Muzzle morphology of the camels Procamelus and

    Megatylopus has been taken to indicate that they were

    browsers, perhaps mixed feeders, but not grazers

    (Dompierre and Churcher, 1996). Similar dominance

    of ruminants over perissodactyls is found in Kenyan

    (Retallack et al., 2002b) and Greek (Solounias, 1981)

    Miocene faunas. By reprocessing their cud, ruminants

    are sustained by less-nutritious forage than perisso-

    dactyls of comparable body sizes (Janis et al., 1994).

    Summer-dry seasons in Kenya, Greece, and western

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 115

    North America selected for ruminant-dominated fau-

    nas able to cope with seasonal shortages of forage,

    whereas equid-dominated faunas of the Great Plains

    and Gulf Coast were supported by year-round forage.

    7. Implications for global change

    7.1. Neogene paleoclimate in eastern Oregon

    A detailed Neogene paleoclimatic time series (Fig.

    12A–C) can be constructed by extending the results

    from paleosols presented here to other paleosols,

    including paleosols from the early Miocene upper

    John Day Formation near Kimberly, Oregon (Retal-

    lack, 2004) dated by extrapolation from 40Ar/39Ar

    radiometry (Fremd et al., 1994), the middle Miocene

    Mascall Formation and Columbia River Basalts near

    Dayville, Oregon (Bestland and Krull, 1997; Sheldon,

    2003) dated by magnetostratigraphy and radiometry

    (Draus and Prothero, 2002), the latest Miocene

    Rattlesnake Formation near Dayville, Oregon (Retal-

    lack et al., 2002a) dated by magnetostratigraphy

    (Hoffman and Prothero, 2002), the Mio–Pliocene

    Ringold Formation near Pasco and Taunton, Wash-

    ington (Gustafson, 1978; Smith et al., 2000) dated by

    magnetostratigraphy (Gustafson, 1985; Morgan and

    Morgan, 1995), and the Palouse Loess near Helix,

    Oregon, and Washtucna, Washington dated by mag-

    netostratigraphy, tephrochronology, and thermolumi-

    nescence dating (Busacca, 1989, 1991; Tate, 1998;

    Blinnikov et al., 2002).

    Fig. 12. Miocene paleoclimate of eastern Oregon follows the

    oceanic foraminiferal carbon isotopic record but not the oceanic

    oxygen isotopic record nor the record of volcanism in Oregon: (A)

    mean annual precipitation from depth to Bk horizon of 799

    paleosols (heavy line and 1r error envelope): (B) mean annualprecipitation from chemical composition of 97 paleosol Bt horizons

    (heavy line and 1r error envelope); (C) mean annual temperaturefrom chemical composition of paleosols (heavy line and 1j errorenvelope): (D) frequency of lavas erupted from Cascades volcanoes

    (from McBirney et al., 1974); (E) oxygen isotopic data (y18Oapatite)from fossil horse teeth in Oregon and Nebraska (Kohn et al., 2002;

    Passy et al., 2002); (F) oxygen isotopic (y18Ocarb) and carbonisotopic values (y13Ccarb) of 8959 samples of marine foraminiferafrom deep-sea cores (from Zachos et al., 2001). Abbreviations are

    Geringian (G.), Monroecreekian (MO.), Harrisonian (HA.), Hemi-

    ngfordian (HEMI.), Barstovian (BAR.), Clarendonian (CLAR.),

    Hemphillian (HEM.), and Blancan (BLA.). Most recent Irvingto-

    nian and Rancholabrean are unlabelled.

    The resulting curve shows great variation in both

    precipitation (Fig. 12A–B) and temperature (Fig. 12C)

    through time and is comparable with variation in

    carbon isotopic composition of benthic marine fora-

    minifera (Fig. 12F). Also comparable are time series of

    North American mammalian diversity, which was

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123116

    higher during warmer and wetter times than during

    cooler and drier times (Alroy et al., 2000; Janis, 2002).

    7.2. Mismatch of Neogene terrestrial and marine

    oxygen isotopic records

    Neither paleoclimatic fluctuations from paleosols

    (Fig. 12A–C) nor oxygen isotopic fluctuations from

    mammal teeth (Fig. 12E) correspond with fluctua-

    tions in oxygen isotopic composition of benthic

    marine foraminifera (Fig. 12F). Marine and terrestrial

    carbonate oxygen isotopic records do not correspond

    during the late Paleocene (Koch et al., 2003) or late

    Permian (MacLeod et al., 2000) either. Although

    marine oxygen isotopic records have been regarded

    as a global paleoclimatic standard (Zachos et al.,

    2001), their mismatch with continental records is a

    fundamental problem for global change studies.

    The mismatch between North American mamma-

    lian evolutionary dynamics and marine oxygen iso-

    topic records has been discussed at length by Alroy et

    al. (2000), who attribute the mismatch to biotic

    insensitivity to climatic forcing. This is contrary to

    the observation of iterative evolution of mammal

    ecomorphs within North American land mammal

    bagesQ which average 2.3 Myr in duration (Martin,1994; Meehan and Martin, 1994; Meehan, 1996,

    1999). The sediments of each mammal age in the

    Great Plains are bounded at the base by erosional

    downcutting, interpreted as evidence of a wet climatic

    phase, and terminated by shallow calcic paleosols or a

    caprock caliche, interpreted as evidence of a dry

    climatic phase (Schultz and Stout, 1980). I have made

    comparable observations in Montana and Oregon

    (Retallack, 1997b, 2001a; Retallack et al., 2002a).

    Late Oligocene mammalian communities of Oregon

    also show alternation between sagebrush steppe with

    Hypertragulus and wooded grassland communities

    with Nanotragulus on Milankovitch time scales (41–

    100 ka) within paleosols with differing depth to calcic

    horizons reflecting alternating semiarid and subhumid

    paleoclimates (Retallack, 2004; Retallack et al., 2004).

    Finally, there is annual variation in oxygen isotopic

    composition of growth bands within Miocene mam-

    malian tooth enamel (Fox, 2001; Kohn et al., 2002).

    North American mammals show clear sensitivity to

    climate on time scales ranging from evolutionary to

    ecological contrary to Alroy et al. (2000).

    A possible reason for mismatch of marine oxygen

    and continental paleoclimatic records is the Cenozoic

    evolution of C4 photosynthetic pathways in land

    plants. The C4 (Hatch–Slack) pathways of tropical

    grasses result in isotopically heavier plant carbon than

    C3 (Calvin–Benson) pathways (Cerling et al., 1997),

    and this pathway increases oxygen isotopic values of

    plant as well because operating on CO2 (Farquhar et

    al., 1993). The antiquity of the C4 pathway has been

    estimated from phylogenetic analyses at 25 Ma

    (Kellogg, 1999), from isotopic analyses of paleosols

    at 15 Ma (Kingston et al., 1994; Morgan et al., 1994),

    and from isotopic and anatomical studies of permin-

    eralized fossil grasses at 12.5 Ma (Nambudiri et al.,

    1978; age revised by Whistler and Burbank, 1992).

    Isotopic studies of paleosol carbonate and mammalian

    teeth show that C4 grasses became widespread in the

    southern Great Plains after 6.6 Ma but were never

    common in Oregon or the northern Great Plains

    (Cerling et al., 1997; Passy et al., 2002; Fox and

    Koch, 2003). C4 grasses avoid cool and summer-dry

    climates today (Sage et al., 1999). The spread of C4grasses after 6.6 Ma is too late to explain the earlier

    divergences of marine carbon and oxygen isotopic

    records (Fig. 12F), mismatch of marine and terrestrial

    oxygen isotopic records (Fig. 12A–C,F), and the

    generally upward trend of marine oxygen isotopic

    values (Fig. 12F).

    A more likely explanation for mismatch of the

    oceanic oxygen isotopic and continental paleocli-

    matic proxies is isotopic depletion as rain shadows

    spread and intensified with uplift of the Cascade

    volcanic and other western mountain ranges (Kohn

    et al., 2002). The idea of Cenozoic drying by means

    of an orographic rain shadow has been a traditional

    explanation for pronounced Neogene aridity evident

    from fossil plants (Chaney, 1948; Ashwill, 1983).

    The growth of the Oregon Cascades can be inferred

    from numerous radiometrically dated eruptions (Fig.

    12D; McBirney et al., 1974; Priest, 1990), but this

    does not match either marine or continental records

    (Fig. 12A–F). The Cascade and Klamath Mountain

    rain shadow has been important in creating generally

    dry long-term climate in eastern Oregon since the

    late Oligocene (30 Ma; Retallack et al., 2000).

    Rocky Mountain barriers isolated eastern Oregon

    from Gulf Coast cyclonic circulation since at least

    early Miocene (19 Ma), as indicated by the

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 117

    appearance of sodic–silicic rather than calcic paleo-

    sols (Retallack, 2004) and very different oxygen

    isotopic composition of fossil horse teeth in Oregon

    and Nebraska (Fig. 12E). Fossil floras from the

    northern Great Basin indicate high altitudes since at

    least the early Miocene, for example, 2100 m for the

    middle Miocene 49 Camp flora of northwestern

    Nevada (Wolfe et al., 1997). Eastern Oregon was

    elevated during Miocene initial eruptions of the

    Yellowstone hot spot (Humphreys et al., 2000).

    The most important reason for a mismatch of

    oceanic and continental records is isotopic enrichment

    of oceanic oxygen by growth of isotopically depleted

    polar ice caps, which has long been recognized to

    compromise direct paleotemperature interpretation of

    marine oxygen isotopic values (Zachos et al., 2001).

    This factor is not unrelated to the growth of mountains

    with their expanding rain shadows, because montane

    ice caps enrich the isotopic composition of atmos-

    pheric oxygen as well. Ice cap fluctuation in volume

    introduces a highly variable bias into paleotemper-

    ature interpreted from the oxygen isotopic record of

    marine benthic foraminifera. This bias can be cor-

    rected using independent estimates of paleotemper-

    ature from Mg/Ca ratios of foraminifera (Lear et al.,

    2000), but this is compromised by dissolution and

    other forms of diagenesis (de Villiers, 2003), and by

    crustal recycling and other long-term influences on

    ocean chemistry (Veizer et al., 2000).

    7.3. Neogene greenhouse–icehouse fluctuation and its

    causes

    Greenhouse mechanisms for Neogene paleocli-

    matic variation are suggested by correspondence of

    the Oregon paleosol sequence (Fig. 12A–C) with the

    marine signal of carbon sequestration inferred from

    benthic foraminiferal y13C decrease and carbon oxi-dation inferred from benthic foraminiferal y13Cincrease (Fig. 12F). High carbon isotopic values in

    the middle Miocene ocean correspond to warm-wet

    conditions in Oregon, whereas low carbon isotopic

    values in the Plio–Pleistocene ocean corresponded to

    cool-dry conditions in Oregon. High carbon isotopic

    values in marine foraminifera reflect diminished burial

    of chemically reduced carbon and higher atmospheric

    CO2 levels due to some combination of oceanic

    respiration, volcanic emission, or impact destruction

    of biomass. A middle Miocene high in atmospheric

    CO2 has been confirmed by stomatal index estimates

    of atmospheric CO2 (Kürschner et al., 1996; Retallack,

    2001c, 2002). Oceanic proxies of atmospheric CO2 do

    not show such variation (Pagani et al., 1999; Pearson

    and Palmer, 2000) but may be compromised by

    volatility of isotopic composition and runoff (Retal-

    lack, 2002). Despite these problems, a useful working

    hypothesis is control of paleoclimate by a varying

    greenhouse effect from changing atmospheric con-

    centrations of CO2.

    Plausible mechanisms for transient climatic warm-

    ing include impact, volcanism (Coutillot, 2002) and

    methane clathrate release (Kennett et al., 2002). There

    are numerous large (N10 km diameter) Neogene

    craters: Haughton, Canada (24 km, 23F1 Ma), Ries,Germany (24 km, 15F0.1 Ma), Karla, Russia (10 km,5F1 Ma), ElTgygytygn, Russia (18 km, 3.5F0.5 Ma),Bosumtwi, Ghana (10.5 km, 1.03F0.02 Ma), andZhamanshin, Kazachstan (14 km, 0.9F0.1 Ma;Dressler and Reimold, 2001; website http://

    www.unb.ca/passc/impactdatabase accessed Jan. 10,

    2003). Time series of volcanic eruptions are also

    strongly episodic (McBirney et al., 1974), with some

    exceptionally large eruptions, such as the middle

    Miocene Columbia River flood basalts of Oregon

    (Wignall, 2001). Columbia River Basalts and Ries

    Crater impact may have been involved in the middle

    Miocene peak of global warmth, but why was there

    not a comparable warm spike after the 23-Ma

    Haughton Crater or a better correlation between

    North American paleoclimate and Cordilleran volcan-

    ism (Fig. 12D). Because of methane’s distinctively

    low carbon isotopic values, such methane outbursts

    are suspected when carbon isotopic compositions fall

    by more than 4x (Jahren et al., 2001). There are nonegative excursions of this magnitude in Neogene

    carbon isotopic records (Fig. 12F), thus methane is

    unlikely to be the whole explanation either.

    A counterpoint to transient carbon-oxidizing events

    was progressive long-term Neogene global cooling,

    perhaps related to changing oceanic current config-

    uration (Broecker, 1997), mountain uplift (Raymo and

    Ruddiman, 1992), or biotic intensification of weath-

    ering with sod–grassland coevolution (Retallack,

    2001a). Both mountain uplift and increased oceanic

    thermohaline circulation create oligotrophic alpine and

    polar marine communities, with less carbon sequestra-

    http://www.unb.ca/passc/impactdatabase

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123118

    tion potential, but grasslands provided more organic

    and nutrient-rich runoff to oceanic phytoplankton

    (Retallack, 2001b). Grasslands in addition had greater

    amounts of soil carbon, higher albedo, and lower rates

    of transpiration than the woodlands and shrublands

    they replaced in subhumid to semiarid regions of the

    world (Nepstad et al., 1994; Jackson et al., 2002), and

    this also could have contributed to global drying and

    cooling (Retallack, 2001b). This mechanism of global

    change is biological, because it is driven by coevolu-

    tion of grasses and grazers, with crumb-textured,

    organic, fertile soils nourishing hypsodont ungulates

    capable of withstanding abrasion from dusty, siliceous

    grasses (Retallack, 2001a). The progressive evolution

    and spread of crumb-textured soils, siliceous grasses,

    and hypsodont ungulates proceeded against a backdrop

    of volatile climatic change, and both short-sod and tall-

    sod grassland paleosols appearing at times of relatively

    high temperature and humidity (Fig. 12A). Grasslands

    did not merely fill in dry times and places (Kohn et al.,

    2002) but were a biological force for climatic cooling

    and drying in their own right (Retallack, 2001b).

    Paleoclimatic warm spikes due to volcanism, bolide

    impact, and methane clathrate release may also have

    been exacerbated by newly coevolved Cenozoic

    grasslands, because ruminants liberate CO2 and CH4.

    The middle Miocene thermal optimum was a time of

    unusually high diversity of North American plants and

    mammals, in part due to evolution within high-

    productivity grassland–woodland mosaics in a warm-

    wet CO2 greenhouse (Janis, 2002), and in part due to

    immigration from Asia through high-latitude land

    bridges (Wolfe and Tanai, 1980; Kohno, 1997;

    Lambert and Shoshani, 1998; Webb, 1998). There

    were more grazing mammal species then ever before,

    and also more browsers than before or after (Janis et

    al., 2000). The middle Miocene greenhouse peak was

    preceded and followed by several other episodes of

    greenhouse paleoclimate of lesser magnitude, includ-

    ing early Clarendonian communities described here.

    Warm-wet conditions of greenhouse transients, even if

    initiated by tectonic or impact forcing, would be

    enhanced by increased mammalian diversity and

    cropping of vegetation. Cold-dry conditions follow

    carbon and water sequestration by grassland soils, their

    high albedo and burial of carbon-rich grassland soil

    crumbs. Paleoclimatic cycles on million-year time

    scales could reflect alternating evolutionary advances

    by plants and animals. Such reciprocating coevolution

    is apparent from current estimates of the times for

    evolution of bunch grassland paleosols (33 Ma), short-

    sod grassland paleosols (19 Ma), and tall-sod grass-

    land paleosols (7.3 Ma; Retallack, 1997b, 2001b,

    2004; Retallack et al., 2002a), which alternate with

    evolution of horse tridactly (36 Ma), cursoriality (25

    Ma), hypsodonty and springing gait (17 Ma), mono-

    dactyly, wide muzzles, and knee-locking (12 Ma), and

    large size and passive-stay shoulder (5 Ma; MacFad-

    den, 1992; Janis and Wilheim, 1993; Hermanson and

    MacFadden, 1992, 1996). This schedule is compatible

    with a coevolutionary process long envisioned for

    grassland ecosystems (Jacobs et al., 1999). Global

    change involves more than just impact or volcanic

    forcing, but a complex interplay of factors for which

    detailed proxies are now becoming available (Fig. 12).

    These various records will have to become even more

    detailed before answers become clear.

    Acknowledgements

    For introduction to the fossils of this area, I

    thank Ted Fremd and his collecting crew, Scott

    Foss, Lia Vella, Delda Findieson, Alois Pajak,

    Daniella Gould, Matthew, and La Rue Rogers. Kent

    Nelson allowed access to his land near Hereford.

    Mary Oman of the Baker City Bureau of Land

    Management helped with geological reconnaissance,

    funding, and research clearance. Megan Smith, Ben

    Cook, and Russell Hunt helped with fieldwork and

    fossil collection. Alicia Duncan prepared numerous

    difficult petrographic thin sections, and Nathan

    Sheldon provided bulk density measurements.

    References

    Alroy, J., Koch, P.L., Zachos, J.C., 2000. Global climate change and

    North American mammalian evolution. Paleobiol. Suppl. 26,

    259–288.

    Ashwill, M., 1983. Seven fossil floras in the rain shadow of the

    Cascades Mountains. Or. Geol. 45, 107–117.

    Axelrod, D.I., 1985. Miocene floras from the Middlegate Basin,

    west-central Nevada. Univ. Calif. Publ. Geol. Sci. 129 (279 pp.).

    Axelrod, D.I., 1995. The Miocene Purple Mountain Flora of

    western Nevada. Univ. Calif. Publ. Geol. Sci. 139 (62 pp.).

    Baldwin, E.M., 1964. Geology of Oregon. Cooperative Book Store,

    Eugene. 165 pp.

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123 119

    Barbour, E.H., 1932. The mandible of Platybelodon barnumbrowni.

    Bull. Nebr. State Mus. 1, 155–158.

    Barbour, M.G., Major, J. (Eds.), Terrestrial Vegetation of California.

    Spec. Publ. Calif. Native Plant Soc. Sacram, vol. 9, 1030 pp.

    Baskin, J.A., 1998. Procyonidae. In: Janis, C.M., Scott, K.M., Jacobs,

    L.L. (Eds.), Evolution of Tertiary Mammals of North America,

    Terrestrial Carnivores, Ungulates and Ungulatelike mammals,

    vol. 1. Cambridge University Press, Cambridge, pp. 144–173.

    Bernor, R.L., Kovar-Eder, J., Lipscomb, P., Rfgl, F., Sen, S.,Tobien, H., 1988. Systematics, stratigraphic, and paleoenvir-

    onmental contexts of first-appearing Hipparion in the Vienna

    Basin, Austria. J. Vertebr. Paleontol. 8, 427–452.

    Berry, E.W., 1931. A Miocene flora from Grand Coulee, Wash-

    ington. U. S. Geol. Surv. Prof. Pap. 170-C, 31–42.

    Bestland, E.A., Krull, E.S., 1997. Mid-Miocene climatic optimum

    recorded in paleosols from the Mascall Formation (Oregon).

    Abstr. Geol. Soc. Amer. 29 (6), 5.

    Bestland, E.A., Hammond, P.E., Blackwell, D.L.S., Kays, M.A.,

    Retallack, G.J., Stimac, J., 1999. Geologic framework of the

    Clarno Unit, John Day Fossil Beds National Monument, central

    Oregon. Or. Geol. 61, 3–19.

    Blinnikov, M., Busacca, A., Whitlock, C., 2002. Reconstruction of

    the late Pleistocene grassland of the Columbia Basin, Wash-

    ington, USA, based on phytolith records in loess. Palaeogeogr.

    Palaeoclimatol. Palaeoecol. 177, 77–101.

    Boggs, S., 1972. Petrography and geochemistry of rhombic, calcite

    pseudomorphs from mid-Tertiary mudstones. Sedimentology

    19, 204–219.

    Bowen, R.G., Gray, W.L., Gregory, D.C., 1963. General geology of

    the northern Juntura Basin. Trans. Am. Philos. Soc. 53, 22–34.

    Bown, T.M., 1980. The fossil insectivora of Lemoyne Quarry (Ash

    Hollow Formation, Hemphillian), Keith County, Nebraska.

    Trans. Nebr. Acad. Sci. 8, 99–121.

    Bradbury, J.P., Dieterich, K.V., Williams, J.L., 1985. Diatom flora of

    the Miocene lake beds near Clarkia in northern Idaho. In:

    Smiley, C.J. (Ed.), Late Cenozoic History of the Pacific

    Northwest. American Association for the Advancement of

    Science, pp. 33–59.

    Brewer, R., 1976. Fabric and Mineral Analysis of Soils. Krieger,

    New York. 482 pp.

    Brodkorb, P., 1961. Birds from the Pliocene of Juntura, Oregon. Fla.

    Sci. 24, 169–184.

    Broecker, W.S., 1997. Thermohaline circulation, the Achilles heel

    of our climate system: will man-made CO2 upset the current

    balance? Science 2878, 1582–1588.

    Brown, J.H., Lomolino, M.V., 1998. Biogeography. Sinauer,

    Sunderland, Massachussetts. 691 pp.

    Busacca, A.J., 1989. Long Quaternary record in eastern Wash-

    ington, USA, interpreted from multiple paleosols in loess.

    Geoderma 45, 105–122.

    Busacca, A.J., 1991. Loess deposits and soils of the Palouse and

    vicinity. In: Baker, V.R., Bjornstad, B.N., Busacca, A.J., Fecht,

    K.R., Kiver, E.P., Moody, U.L., Rigby, J.G., Stradling, D.F.,

    Tallman, A.M. (Eds.), The Columbia Plateau;

    In: Morrison, R.B. (Ed.), Quaternary Geology of the United

    States. Geological Society of America, Boulder, pp. 216–228.

    v. K2.

    Cerling, T.E., Harris, J.M., MacFadden, B.J., Leaker, M.G., Quade,

    J., Eisenmann, V., Ehleringer, J.R., 1997. Global vegetation

    change through the Miocene/Pliocene boundary. Nature 389,

    153–158.

    Chadwick, O.A., Hendricks, D.M., Nettleton, W.D., 1987. Silica in

    duric soils: I. A depositional model. Soil Sci. Soc. Am. J. 51,

    975–982.

    Chadwick, O.A., Nettleton, W.D., Staidle, G.J., 1995. Soil poly-

    genesis as a function of Quaternary climate change, northern

    Great Basin, USA. Geoderma 68, 10–26.

    Chaney, R.W., 1948. The Ancient Forests of Oregon. University of

    Oregon, Eugene. 56 pp.

    Chaney, R.W., Axelrod, D.I., 1959. Miocene floras of the Columbia

    Plateau, vol. 617. Carnegie Inst., Washington Publ. 237 pp.

    Clark, J.M., Dawson, M.R., Wood, A.E., 1964. Fossil mammals

    from the lower Pliocene of Fish Lake Valley. Nev. Bull. Mus.

    Comp. Zool. 131, 29–63.

    Compton, L.V., 1935. Two avian fossils from the lower Pliocene of

    South Dakota. Am. J. Sci. 30, 343–348.

    Condit, C., 1944. The Remington Hill flora. Carnegie Institute.

    Washington Publ. 553, 21-55.

    Coutillot, V., 2002. Evolutionary Catastrophism. Cambridge Uni-

    versity Press, Cambridge.

    Dalrymple, G.B., 1979. Critical tables for conversion of K–Ar ages

    from old to new constants. Geology 7, 558–560.

    Dawson, M.R., 1958. Later Tertiary Leporidae of North

    America. Univ. Kans. Paleontol. Contrib., Pap. 22 (Vertebrata

    6, 75 pp.).

    Delancey, S., Genetti, C., Rude, N., 1988. Some Sahaptian–

    Klamath–Tsimshianic lexical sets. In: Shipley, W. (Ed.), In

    Honor of Mary Haas. Mouton de Gruyter, Berlin, pp. 195–224.

    de Villiers, S., 2003. Dissolution effects on foraminiferal Mg/Ca

    records of sea surface temperature in the western equatorial

    Pacific. Paleoceanography 18, 1070.

    Dorf, E., 1936. A late Tertiary flora from southwestern Idaho.

    Carnegie Inst. Wash. Contr. Paleontol. 426, 185–216.

    Dompierre, H., Churcher, C.S., 1996. Premaxillary shape as an

    indicator of the diet of seven late Cenozoic New World camels.

    J. Vertebr. Paleontol. 16, 141–148.

    Downs, T., 1952. A new mastodont from the Miocene of Oregon.

    Univ. Calif. Publ. Geol. Sci. 29, 1–20.

    Downs, T., 1956. The Mascall fauna from the Miocene of Oregon.

    Univ. Calif. Publ. Geol. Sci. 31, 199–354.

    Draus, E., Prothero, D.R., 2002. Magnetic stratigraphy of the

    middle Miocene (early Barstovian) Mascall Formation, central

    Oregon. Abstr. Geol. Soc. Amer. 36 (6), 135–136.

    Dressler, B.D., Reimold, W.J., 2001. Terrestrial impact melt rocks

    and glasses. Earth Sci. Rev. 56, 205–284.

    Evans, J.G., 1992. Geologic map of the Dooley Mountain 7OVQuadrangle, Baker County, Oregon. U.S. Geol. Surv. Geol.

    Quad. (Map GQ-1694, 9 pp.).

    Evernden, J.F., James, G.T., 1964. Potassium–argon dates and the

    Tertiary floras of North America. Am. J. Sci. 262, 945–974.

    F.A.O. (Food and Agriculture Organization), 1974. Soil Map of the

    World. Legend, vol. I. U.N.E.S.C.O, Paris. 59 pp.

    F.A.O. (Food and Agriculture Organization), 1975. Soil Map of the

    World. North America, vol. II. U.N.E.S.C.O, Paris. 210 pp.

  • G.J. Retallack / Palaeogeography, Palaeoclimatology, Palaeoecology 214 (2004) 97–123120

    Farquhar, G.D., Lloyd, D.J., Taylor, J.A., Flanagan, L.B., Syvert-

    son, J.P., Hubick, K.T., Wong, C.S., Ehleringer, J.R., 1993.

    Vegetation effects on the isotopic composition of oxygen in

    atmospheric CO2. Nature 363, 439–443.

    Fiebelkorn, R.B., Walker, G.W., MacLeod, N.S., McKee, E.H.,

    Smith, J.G., 1982. Index to K–Ar determinations for the State of

    Oregon. Isochron-West, vol. 37. 60 pp.

    Fleming, C.A., 1975. Biogeography and ecology in New Zealand.

    In: Kuschel, G. (Ed.), The Geological History of New Zealand

    and its Biota. W. Junk, The Hague, pp. 1–86.

    Fox, D.L., 2001. Growth increments in Gomphotherium tusks

    and implications for late Miocene climate change in

    North America. Palaeogeogr. Palaeoclimatol. Palaeoecol. 156,

    327–348.

    Fox, D.L., Koch, P.L., 2003. Tertiary history of C4 biomass in the

    Great Plains, USA. Geology 31, 809–812.

    Fremd, T., Bestland, E.A., Retallack, G.J., 1994. John Day Basin

    Fieldtrip Guide and Road Log for the Society of Vertebrate

    Paleontology Annual Meeting. Northwest Interpretive Associ-

    ation, Seattle. 90 pp.

    Gilluly, J., 1937. Geology and mineral resources of the Baker

    Quadrangle. U.S. Geol. Surv. Bull. 879 (119 pp.).

    Golia, R.T., Stewart, J.H., 1984. Depositional environments and

    paleogeography of the upper Miocene Wassuk Group, west

    central Nevada. Sediment. Geol. 38, 159–180.

    Graham, A., 1999. Late Cretaceous and Cenozoic History of

    North American Vegetation. Oxford University Press, New

    York. 350 pp.

    Greene, R.C., 1973. Petrology of the welded tuff of Devine

    Canyon, southeastern Oregon. U. S. Geol. Surv. Prof. Pap. 797

    (26 pp.).

    Gregory, J.T., 1942. Pliocene vertebrates from Big Spring Canyon,

    South Dakota. Univ. Calif. Publ. Geol. Sci. 26, 307–442.

    Gustafson, E.P., 1978. The vertebrate fauna of the Pliocene Ringold

    Formation, south-central Washington. Bull. Mus. Nat. Hist.

    Univ. Oregon 23 (62 pp.).

    Gustafson, E.P., 1985. Soricids (Mammalian: Insectivora) from

    the Blufftop local fauna, Blancan Ringold Formation of

    central Washington and correlation of the Ringold Formation.

    J. Vertebr. Paleontol. 5, 88–92.

    2Hall, E.R., 1930. Rodents and lagomorphs from the Late Tertiary

    of Fish Lake Valley, Nevada. Univ. Calif. Publ. Geol. Sci. 19,

    295–312.

    Harden, J., 1982. A quantitative index of soil development from

    field descriptions: examples from a chronosequence in central

    California. Geoderma 28, 1–28.

    Harden, J., 1990. Soil development on stable landforms and

    implications for landscape studies. Geomorphology 3, 391–398.

    Hannibal, H., 1912. A synopsis of the recent and Tertiary

    freshwater Mollusca of the Californian province, based upon

    an ontogenetic classification. Proc. Malacol. Soc. Lond. 10,

    112–211.

    Hayek, L.-A.C., Bernor, R.L., Solounias, N., Steigerwald, P., 1992.

    Preliminary studies of hipparionine horse diet as measured by

    tooth wear. Ann. Zool. Fenn. 28, 187–200.

    Hermanson, J.W., MacFadden, B.J., 1992. Evolutionary and func-

    tional morphology of the shoulder region and stay apparatus in

    fossil and extant horses (Equidae). J. Vertebr. Paleontol. 12,

    377–386.

    Hermanson, J.W., MacFadden, B.J., 1996. Evolutionary and func-

    tional morphology of the knee in fossil and extant horses

    (Equidae). J. Vertebr. Paleontol. 16, 349–357.

    Herrington, H.B., Taylor, D.W., 1958. Pliocene and Pleistocene

    Sphaeriidae (Pelecypoda) from the central United States. Occas.

    Pap. Mich. Univ. Dept. Zool. 596 (23 pp.).

    Hibbard, C.W., Phillis, L.F., 1945. The occurrence of Eucastor and

    Epigaulus in the lower Pliocene of Trego County, Kansas. Bull.

    Univ. Kans. 30, 549–555.

    Hoffman, J., Prothero, D.R., 2002. Magnetic stratigraphy of the

    upper Miocene (early Hemphillian) Rattlesnake Formation,

    central Oregon. Abstr. Geol. Soc. Amer. 36 (6), 135.

    Honey,


Recommended