Content uploaded by Christopher Robert Scotese
Author content
All content in this area was uploaded by Christopher Robert Scotese on May 14, 2024
Content may be subject to copyright.
1
The Cretaceous World: Plate Tectonics, Paleogeography, and Paleoclimate
by
Christopher R. Scotese, Christian Vérard, Landon Burgener, Reece P. Elling, and
Adam T. Kocsis
04/20/2024
2
Contents
Abstract 6
A. Introduction 7
B. Data and Methods 14
1. Global Plate Model: Philosophy and Model Construction 14
2. Geological and Geophysical Evidence Used to Build the Cretaceous Plate Tectonic Model 17
1) Age of the Ocean Basins. 18
2) Oceanic Fracture Zones 19
3) Hot Spots and Hot Spot Tracks 24
4) Paleopoles and Apparent Polar Wander (APW) paths). 24
5) Continental Tectonics. 27
6) The Rules of Plate Tectonics 27
7) Synthetic Seafloor Spreading Isochrons 29
8) Subduction Graveyards 30
9) Large Igneous Provinces (LIPs) 31
10) Paleobiogeography 32
11) Paleoclimate 34
12) True Polar Wander 35
3. Geological and Geophysical Evidence Used to Map Ancient Paleogeography 35
4. Eustatic Changes in Sea Level Derived from Estimates of Continental Flooding 38
1) Introduction 38
2) Estimating Past Sea Level from Changes in Continental Flooding 40
5. Modeling Cretaceous Paleoclimate: Computer Simulations and Lithologic
Indicators of Climate 45
1) Introduction 45
2) Computer Simulations of Paleoclimate 46
a. Paleogeography 47
b. Insolation 47
c. Atmospheric CO2 48
3) Lithologic Indicators of Climate and Köppen Belts 51
4) Combining Lithologic Indicators of Climate with Quantitative
Paleoclimate Proxies 55
6. Modeling Cretaceous Paleorivers 58
1) Drainage system and paleorivers 58
2) Method 58
3
3) Flow length 59
C. Chronological Review of Plate Tectonics, Paleogeography, and Paleoclimate during
the Cretaceous 61
1. Plate Tectonics during the Cretaceous 61
1) Overview 61
2) Earliest Cretaceous (Berriasian – Barremian, 145 Ma – 125 Ma) 69
a. Intra-Pangean Ocean Basins 69
b. Extra-Pangean Ocean Basins 70
3) Early Cretaceous (Aptian - Albian, 125 Ma – 100 Ma) 76
a. Northwest Sector of Pangea 76
b. Central Sector of Pangea 76
c. Southeast Sector of Pangea 77
d. Ocean Basins Exterior to Pangea 78
4) Mid-Cretaceous (Cenomanian-Turonian, 100 Ma – 90 Ma) 83
5) Late Cretaceous (Coniacian - Maastrichtian, 85 Ma – 65 Ma) 91
a. Overview 91
b. Circum-Arctic and North American Cordillera 91
c. North Atlantic and Europe 91
d. Central Atlantic Ocean and Caribbean 92
e. South Atlantic Ocean and Southwest Indian Ocean 93
f. Western Indian Ocean, Madagascar, and India 93
g. Southeast Indian Ocean, Australia, and Zealandia 94
h. Western Pacific and Northeast Asia 95
i. Deccan Flood Basalt 96
2. Paleogeography during the Cretaceous 96
1) Overview 96
2) Cretaceous Mountain Ranges 97
3) Cretaceous Sea Level 98
a. Continental Area 99
b. Mid-ocean Ridge and Trench Length 100
c. Continental Ice 102
4) Cretaceous Landmasses, Landbridges and Oceanic Gateways 103
a. Introduction 103
b. Early Cretaceous (145 Ma – 100 Ma) 114
c. Late Cretaceous (100 Ma – 65 Ma) 122
4
5) Cretaceous Rivers 123
a. Overview 123
b. Average River Length and the Number of Rivers 124
c. Effect of Climate on River Systems 126
d. Effect of Plate Tectonics on River Systems 128
e. Cretaceous Origin of Some Modern Rivers 136
6) Oceanic Circulation during the Cretaceous 137
a. Introduction 138
b. Cretaceous Oceanic Surface Currents 143
3. Paleoclimate during the Cretaceous 144
1) Introduction 144
2) Cretaceous Climate in the Context of Mesozoic Climate Change 149
3) Chronological Review of Climate during the Cretaceous 152
a. Overview 152
b. Early Cretaceous Climate 156
c. Mid to Late Cretaceous – Paleogene Hothouse (128 Ma – 39.4 Ma) 163
d. Cretaceous Oceanic Anoxic Events (OAEs) 165
e. The Problem of Cold Polar Regions during the Cretaceous 166
f. The End Cretaceous Impact Winter 67
4. Changing Patterns of Precipitation during the Cretaceous 69
1) Persistent Patterns of Regional Precipitation 169
2) Changing Patterns of Regional Precipitation 171
a. Western Hemisphere 175
b. Eastern Hemisphere 176
3) Comparison of Precipitation Patterns: Computer Simulation versus
Geological Evidence 176
D. Summary 178
1. Introduction 178
2. Summary of Plate Tectonic Data and Methods 178
3. Summary of Paleogeographic Data and Methods 179
4. Summary of Paleoclimatic Data and Methods 180
5. Summary of Plate Tectonics Events during the Cretaceous 180
6. Summary of Paleogeographic Events during the Cretaceous 183
7. Summary of Paleoclimatic Events during the Cretaceous 184
8. Summary of Cretaceous Rivers 186
5
9. Summary of Cretaceous Ocean Circulation 183
10. Summary of Cretaceous Precipitation Patterns 187
E. Discussion 188
F. Conclusions 191
1. Some Conclusions about the Cretaceous 191
2. How well do we know the Cretaceous? 192
3. Future Earth System History Research and the Use of Artificial Intelligence (AI) 194
G. Acknowledgements 196
H. References Cited 197
I. List of Figures 232
J. Appendix 236
Table 1 236
Table 2 237
Table 3 240
Table 4 247
Table 5 248
Table 6 250
Table 7 260
Table 8 263
Table 9 267
Table 10 268
Table 11 273
Table 12 275
Table 13 276
Table 14 278
6
Abstract
The tectonics, geography, and climate of the Cretaceous world was a very different from the modern world. At the
start of the Cretaceous, the supercontinent of Pangea had just begun to break apart and only a few small ocean
basins separated Laurasia, West Gondwana, and East Gondwana. Unlike the modern world, there were no
significant continent-continent collisions during the Cretaceous and the continents were low-lying and easily
flooded. The transition from a Pangea-like configuration to a more dispersed continental arrangement had
important effects on global sea level and climate. During the Early Cretaceous, as the continents rifted apart, the
new continental rifts were transformed into young ocean basins. The oceanic lithosphere in these young ocean
basins was thermally elevated, which boosted sea level. Sea level, on average, was ~70 m higher than the present-
day. Sea level was highest during the mid-Cretaceous (90 Ma – 80 Ma), with a subsidiary peak ~ 120 million years
ago (early Aptian). Overall, the Cretaceous was much warmer than the present-day (> 10˚C warmer). These very
warm times produced oceanic anoxic events (OAEs) and high temperatures in equatorial regions sometimes made
terrestrial and shallow marine ecosystems uninhabitable (temperatures > 40˚C). This is unlike anything we have
seen in the last 35 million years and may presage the eventual results of man-made global warming. This mostly
stable, hot climate regime endured for nearly 80 million years before dramatically terminating with the Chicxulub
bolide impact 66 million years ago. Temperatures plummeted to icehouse levels in the “impact winter” resulting
from sunlight-absorbing dust and aerosols. As a consequence of the collapse of the food chain, ~75% of all species
were wiped out (Sepkoski, 1996). The effect of this extinction event on global ecosystems was second only to the
great Permo-Triassic Extinction (McGhee et al., 2013).
7
A. Introduction
The Cretaceous Period was defined by D’Halloy in 1822. It was divided into two Epochs (Lower Cretaceous and
Upper Cretaceous) by Conybeare and Phillips (1822), each Epoch with 6 stages (D’Orbigny, 1840; Figure 1, Cohen
et al., 2013). The Cretaceous is the longest Period of the Phanerozoic Era (79 million years) and is 33% longer than
the other periods of long duration such as the Cambrian, Devonian, and Carboniferous (each ~60 million years in
length). In this paper, we present 17 plate tectonic and paleogeographic reconstructions and nine paleoclimatic
reconstructions for the Cretaceous (Table 1).
A lot happened during the Cretaceous. It was the acme of the age of reptiles (dinosaurs and pterosaurs), and the
great extinction event at the end of the period set the stage for the rise of mammals. Swimming reptiles
(ichthyosaurs, plesiosaurs, and mosasaurs), hard-shelled cephalopods (ammonites), and planktic microorganisms
(foraminifera and coccoliths) filled the seas. Excellent summaries of the ecology and evolution of life during the
Cretaceous can be found in Bakker (1986, 1995); Brusatte (2018), Chatterjee (1997); Dingus and Rowe (1998), Ellis
(2003); Everhart (2005), Fastovsky and Weishampel (1996);Morley (2000), Paul (2010, 2022a&b); Skelton (2003);
Weishampel et al., (1990); Wellnhofer (1991); and Willis and McElwain (2002); Winton (2013). In this essay, we
attempt to paint a comprehensive picture of what the Cretaceous world was like – the tectonic activity, the
geographic distribution of the continents and ocean basins, the relative amounts of land and sea, the climatic
zones, the river systems, and ocean currents.
8
Figure 1. Cretaceous Timescale (Cohen et al., 2013; version 2023/06)
The plate tectonic reconstructions (e.g., Figure 2A) are based on marine magnetic anomalies, the trends of oceanic
fracture zones, hot spot tracks, mantle tomography, and a new synthesis of paleomagnetic data (Elling, 2022). This
global plate model is used to predict the changing age of the ocean floor and, hence, the changing volume of the
ocean basins and the resulting eustatic sea level change and continental flooding. An additional 17 maps (e.g.,
Figures 2B) illustrate the changing paleogeography of the Cretaceous.
9
Paleogeographic maps illustrate the changing distribution of mountains, uplands, lowlands, shallow seas, and deep
ocean basins. No major continent-continent collisional mountain ranges formed during the Cretaceous. Nearly all
mountain-building occurred along the subducting margins of the Panthallasic and Tethys oceans. Andean-type
subduction and the collision of exotic terranes formed long, linear mountain belts. The interiors of the continents
were largely flat, with the exception of Central Asia where late Paleozoic mountain ranges still towered over the
landscape. High sea level (+150 m; this study) during the mid-Cretaceous (Cenomanian-Turonian-Coniacian)
flooded the continents. Sea level at the beginning and end of the Cretaceous was only slightly higher than modern
sea level (0 m to + 40 m).
Special attention is given in this paper to the changing climate of the Cretaceous. Paleoclimatic maps illustrate
ancient temperatures (Figure 3A), rainfall (Figure 3B), ocean circulation (Figure 3C) and most importantly the
changing extent of Köppen Climatic Belts (e.g., Figure 2C; Burgener et al., 2023). Köppen Climatic Belts are defined
by seasonal variations in temperature and precipitation (Köppen, 1918) and are the best way to characterize and
visualize past climates. Variations in regional climates create a mosaic of environments and ecological habitats. The
extent of the principal Köppen zones - Tropical Everwet (A), Subtropical Arid (B), Warm Temperate (C), Cool
Temperate (D), and Polar (E) - determines the distribution of the environments and habitats. There were no large
permanent ice caps during the Cretaceous, though small ice caps (~4 million km2 ) may have existed at the South
pole during the earliest and latest Cretaceous (Scotese et al., 2021). The paleo-Köppen maps also show the likely
drainage pattern and the course of Cretaceous rivers (e.g., Figure 2C).
Included in the Supplemental Materials are complete sets of all kinds of maps discussed in the body of the text.
Also included are computer animations of: plate motions, paleogeography, oceanic circulation, and Köppen belt
evolution, as well as supporting documentation in the form of tables, data files, and spreadsheets. The
Supplemental Materials are available in two on-line archives. One archive,
https://doi.org/10.5281/zenodo.10659104, deals exclusively with Cretaceous material. The other archive,
https://doi.org/10.5281/zenodo.10659112, contains material that is Phanerozoic in scope.
.
How to read this chapter. This essay is organized into two principal sections: 1) a detailed discussion of the data
and methods used to produce the plate tectonic, paleogeographic, and paleoclimatic maps of the Cretaceous and,
2) a description of the chronology of Cretaceous plate tectonic, paleogeographic, and paleoclimatic events. Though
it was written to be read through from beginning to end, each of these sections stands alone and can be read in
any order. An alternate approach would be to start with the concise outline in the Summary section and then read
the more detailed descriptions in each of the sections on plate tectonics, paleogeography, and paleoclimate. For
those who want to quickly get to the gist of this essay, the Conclusions contain a condensed overview of the
10
important ideas presented in this work. All of the tables cited in the text are located in the Appendix at the end of
this essay.
11
12
Figure 2. Example Cretaceous plate tectonics, paleogeographic, and paleoclimatic maps
A. Plate Tectonic Reconstruction for the early Aptian (120 Ma), black dots – hot spots, dark gray – hot spot tracks &
LIPs, medium gray – continental lithosphere, light gray – areas of continental convergence, red lines – mid-ocean
ridges, blue lines – subduction zones, color shading – age of ocean floor; color legend units - millions of years; B.
Paleogeographic Reconstruction, dark blue – deep oceans, light blue – flooded continental lithosphere & ocean
islands, green – lowlands, tan – uplands, brown – mountains; color legend units – meters; C. Paleoclimatic
Reconstruction, Köppen Belts: dark green - Tropical Everwet, yellow & tan -Subtropical Arid, light green – Warm
Temperate, purple – Cold Temperate, Paleorivers: blue dashed lines– rivers, red dashed lines, dry rivers.
13
14
Figure 3. Example Cretaceous Paleoclimatic Maps, A. Global Average Temperature for the early Aptian (120 Ma),
red – yellow - green = warm; pale green – cool; purple – cool; blue - ice; The faint contours are isotherms; color
legend units - ˚C; B. Annual Precipitation, blue = wet; orange = arid; color legend units – mm/day C. Oceanic
Circulation, arrows indicate current flow; dark blue = high velocity; white = low velocity; color legend units –
centimeters/second.
B. Data and Methods
1. Global Plate Model: Philosophy and Model Construction
The first step in producing a plate tectonic reconstruction is to build a global plate tectonic model that describes
the evolution of the continents and ocean basins. The Cretaceous plate tectonic and paleogeographic maps do not
stand alone but rather are part of a global plate tectonic and paleogeographic model that describes the evolution
of the continents and ocean basins during the last 1.5 billion years (Scotese and Elling, 2017; Scotese, 2018). This
dynamic Earth model has been assembled through the research efforts of the PALEOMAP Project during the last 40
years (Scotese and Sager, 1988; Scotese et al., 1988; Scotese, 1990; Scotese and McKerrow, 1990; Scotese, 2001;
Scotese and Dammrose, 2008; Scotese, 2014b; Scotese, 2016; Scotese and Elling, 2017; Scotese, 2018; Scotese,
2021.)
This description of global plate motions has two important components: 1) a global set of tectonic elements
defined by ancient plate boundaries (rifts, subduction zones, and sutures that mark zones of collision), and 2) a
description of how these tectonic elements move through time. The first component, the tectonic element map,
defines the crustal fragments that have had an history of independent motion. A description of the tectonic
elements used to produce the Cretaceous plate tectonic reconstructions is given in Table 2. Figure 4 shows the
modern location of these tectonic elements. Figure 5 illustrates the paleo-position of these tectonic elements
during the Cenomanian (100 Ma).
15
Figure 4. Tectonic Elements (static polygons), Scotese (2016, 2021), see Table 2 for legend.
Figure 5. Albian/Cenomanian (100 Ma) reconstruction of tectonic elements (Scotese, 2016, 2021), see Table 2 for
legend. The gaps between tectonic elements in south-central Asia and along the North American Cordillera
indicate continental regions of compression and convergence.
The second component of the PALEOMAP Global Plate Model is a set of hierarchical finite rotations that precisely
describes how these tectonic elements have moved through time. This plate modeling technique was first
described by Ross and Scotese (1988) and is based on the concept of a plate tectonic circuit (Cox and Hart, 1986).
As an example, Table 3, lists the finite rotations that describe the movement of the continents that comprised
16
Gondwana during the Cretaceous. Figure 6 illustrates the hierarchical relationship of these relative rotations. In
other words, “India moves relative to Madagascar”, “Madagascar moves relative to Somalia”, and “Somalia moves
relative to Arabia” and so on. A listing of the complete PALEOMAP Global Plate Model is provided in the
Supplemental Materials.
Through the years, the software used to produce plate tectonic and paleogeographic reconstructions has evolved.
The software started out as simple FORTRAN programs (Scotese and Baker, 1975; Scotese, 1976, 1983; Scotese et
al., 1980) and were subsequently translated into a succession of programming languages (e.g., Pascal, C, C++,
Visual Basic, Python, etc.; Scotese et al., 1985; Scotese and Denham, 1988; Walsh and Scotese, 1993). Plate
modeling software has evolved to the point where we can produce plate tectonic reconstructions for any instant in
time as far back as the early Proterozoic (Eglington et al., 2017; Scotese and Elling, 2017). The current version of
this plate-modelling software, GPlates (Müller et al., 2018), is widely used and can be freely download from the
EarthByte website ( https://www.gplates.org). GPlates was the principal tool used to produce the plate tectonic
reconstructions shown here. A complete set of digital files comprising the PALEOMAP Global Plate Tectonic Model
is included with the Supplemental Materials and also can be found on-line at
https://doi.org/10.5281/zenodo.5460860
Figure 6. Plate tectonic hierarchy describing the movement of the 30 tectonic elements that make up Gondwana.
Note that some plates, such as India, move with respect to different fixed plates depending on the time. These
transitional times are called cross-overs.
17
The fundamental importance of the Global Plate Tectonic Model (GPTM) cannot be overemphasized. It is
paramount. The GPTM provides the framework upon which everything else is built and evaluated. We do not make
a map for the early Cretaceous, a map for the mid-Cretaceous, and another map for the late Cretaceous, rather we
build a continuously evolving plate tectonic model that illustrates the dynamic evolution of the plate boundaries
and describes the ever-changing configuration of continents and ocean basins. There is a strong sense of
contingency. In a very real sense, the early Cretaceous plate reconstruction predicts and provides a context for the
mid-Cretaceous plate reconstruction, and the mid-Cretaceous plate reconstruction provides the context for the
late Cretaceous plate reconstruction, etc. Much like a trained neural network, the global plate tectonic model is a
complex web of linked interdependencies. No one map or time interval can stand alone.
All paleontologic, paleoclimatic, geological and geophysical data are evaluated in the context of the global plate
tectonic model. If there is a mismatch between the data and the model, either the model is adjusted to better
explain the data or the contrarian data are rejected. The goal is to eventually have a self-consistent, dynamic Earth
System model that describes how the Earth has evolved through time.
In the Supplemental Materials there is a brief essay describing how global plate tectonic models developed from
the mid 1970’s to the present-day. This chronology describes the plate modeling software and research groups
that produced them. Recently, Buffan et al. (2023) estimated the reliability and similarity of recent global plate
tectonic models. Their conclusions were that the models are in very good agreement back to 100 million years, but
diverge significantly for times greater than 300 million years.
2. Geological and Geophysical Evidence Used to Build the Cretaceous Plate Tectonic Model
Figure 7 summarizes the various kinds of information used to build the PALEOMAP Global Plate Tectonic Model.
The diagram shows how the number and the relative importance of various lines of evidence changes through
time. For example, when making Precambrian plate tectonic reconstructions, only 6 of the 12 lines of evidence can
be brought to bear. The other four lines of evidence are not available. For the Cretaceous, we are fortunate that
all 12 lines of evidence can be used to find a solution. In order of importance, these lines of evidence are: 1) Age of
the Ocean Basins, 2) Oceanic Fracture Zones, 3) Hot Spots and Hot Spot Tracks, 4) Paleopoles and Apparent Polar
Wander (APW) paths, 5) Continental Tectonics, 6) the Rules of Plate Tectonics 7) Synthetic Seafloor Spreading
Isochrons, 8) Large Igneous Provinces, 9) Subduction Zone Graveyards, 10) Paleobiogeography, 11) Paleoclimate,
and 12) True Polar Wander. In the next section, we describe the role that each line of evidence plays when making
Cretaceous plate tectonic reconstructions.
18
Many of these lines of evidence are plotted on Figure 8. The Tectonic Map of the World is an update of the map
originally published by Larson et al.(1985) and is similar to other recent efforts (Francois et al., 2021). It illustrates a
variety of continental tectonic features such as : the trends of mountain ranges, ophiolites, Large Igneous
Provinces (LIPs), and continental volcanic arcs , as well as important oceanic tectonic features such as: the age of
the ocean floor, trends of fracture zones, hot spot tracks, oceanic volcanic island arcs, and subduction-related
features such as trenches, accretionary prisms, and fore-arc basins (Figure 8). Large-format versions of the
Tectonic Map of the World are available in the Supplemental Materials; one version identifies more than 300
oceanic and continental features.
Figure 7. Lines of evidence used to produce a plate tectonic reconstruction. The width of the colored bands
indicates the relative importance of each kind of evidence. Note that the width of the bands changes back through
time and that fewer lines of evidence are available for times greater than 600 million years.
1) Age of the Ocean Basins. As described in most introductory Earth Science textbooks , the polarity of Earth’s
magnetic field frequently reverses causing the rocks that form at mid-ocean ridges to be magnetized in opposite
19
directions. These linear magnetic anomalies were first dated and mapped on a global scale by Larson et al. (1985);
Cande et al. (1989). The first global digital compilation of oceanic magnetic anomalies was made by the research
team at the Institute for Geophysics at the University of Texas in the mid-1980’s and was published in the 1990’s
(Royer et al., 1992; Müller et al., 1993, 1997) and updated by Müller et al. (2008) and Seton et al.(2020).
2) Oceanic Fracture Zones. Major oceanic fracture zones (FZ) generally form at right angles to the mid-ocean
ridges. In many cases, the zig-zag pattern of major fracture zones represents the offset patterns of the continental
rifts that formed when continents began to separate. Often oceanic fracture zones can be traced from one side of
the ocean to the other side (e.g. Romanche, St. Paul, and Falkland-Aghulas fracture zones; Granot and Dyment,
2015) and therefore provide important constraints on the pre-breakup configuration of the continents. Also shown
(Figure 8) is the “tectonic fabric” of the ocean basins ( Gahagan et al., 1988; Smith and Sandwell, 1997; Sandwell et
al., 2013) which has been generated by seafloor spreading, hot spot, and subduction-related processes.
20
Figure 8A. Tectonic Map of the World (part 1) (See Figure 9 for legend).
21
Figure 8B. Tectonic Map of the World (part 2) (See Figure 9 for legend).
22
Figure 8C. Tectonic Map of the World (part 3) (See Figure 9 for legend).
23
Figure 8D. Tectonic Map of the World (part 4) (See Figure 9 for legend).
24
Figure 9. Legend for Tectonic Map of the World.
3) Hot Spots and Hot Spot Tracks. There are ~60 important mantle plumes or hot spots (Richards et al., 1989; Ernst,
2014; Koppers et al., 2021), approximately 30 of these hotspots were active during the Cretaceous. Volcanic
islands and plateaus form where these hotspots penetrate the oceanic lithosphere. The motion of the plates over
the relatively fixed hotspots generates hotspot tracks, the most well-known being the Hawaiian-Emperor hot spot
track (Figure 8). Hot spots generally do not produce hot spot tracks on continents because the thicker continental
crust and lithosphere is more difficult to penetrate and often deflects the mantle plume.
Ten well-dated hotspot tracks in the Atlantic and Indian Ocean, were used to constrain the motion of the plates
relative to the mantle during the Cretaceous (Müller et al., 1993). Several hot spot tracks in the Pacific, including
the Hawaiian-Emperor hot spot track, were also used to constrain relative plate motions. Paleomagnetic
information, however, indicates that these hot spots may have moved about ~20˚ relative to the Atlantic and
Indian Ocean hot spots (Sager, 2007; Antretter, 2001; Koppers et al., 2021).
4) Paleopoles and Apparent Polar Wander (APW) paths. All plate tectonic reconstructions rely on paleomagnetic
measurements to estimate the latitudinal position of the continents. The inclination (I) of the remanent
25
magnetization of an oriented rock sample provides a direct estimate of the ancient latitudinal position of a
continent (McElhinny, 1973). Though paleobiogeographic, paleoclimatic, and tectonic information can provide
additional constraints (Figure 7), paleopoles are the single most important quantitative line of evidence. For an
excellent summary of paleomagnetic methods see McElhinny, 1973; van der Voo, 1993; Butler, 1998; McElhinny
and McFadden, 2000; Tauxe, 2002; Lowrie, 2007). Van Hinsbergen (2015) has constructed a website,
https://paleolatitude.org/ that estimates the paleolatitude of ancient geographic terranes based on paleomagnetic
information.
Though paleomagnetic data are extremely useful, there are large uncertainties in the location of a paleopole due
to confounding diagenetic, tectonic, and sedimentologic effects. The greatest uncertainty is the age of the
magnetization. Often the age of the magnetization of a paleomagnetic sample is reset by younger thermal and
chemical events. For these reasons, individual paleomagnetic determinations are often unreliable. It is not wise to
rely on a single paleopole to orient a paleocontinent or terrane, no matter how well-determined the pole appears
to be. Paleomagnetic analysis is a statistical endeavor (see gray dots and circle in Figure 10). When building a plate
tectonic model, it is necessary to use as many paleomagnetic measurements as possible and not reduce the
available data set by preselecting preferred paleopoles prior to testing their validity. If possible, paleopole
compilations should be based on site-level statistics rather than aggregated paleopole positions (Vaes, 2023; Vaes
et al. 2023; ).
In order to provide a consistent and comprehensive paleomagnetic framework, more than 15,000 paleopoles
covering the last 1.5 billion years were compiled by Elling (2022) from four primary sources: van der Voo (1993),
Scotese and van der Voo (2017), 1636 paleopoles; Torsvik et al. (2008, 2012), 941 paleopoles; Viekkolainen et al.
(2017), 3798 paleopoles; and Pisarevsky(2005), 9514 paleopoles. These paleopoles were combined with additional
sources: Evans (2017), 298 paleopoles; Merdith (2017), 153 paleopoles; Tetley (2018), 345 paleopoles; Pisarevsky
et al. (2022), 10023 paleopoles ; and Vaes (2023), Vaes et al. (2023), 705 paleopoles. The paleopoles from these
compilations were standardized in a format that included 18 principal fields (Table 4). A complete listing of all of
the paleopoles in the PALEOMAP Paleomagnetic Database is provided in the Supplemental Materials. The
PALEOMAP Global Plate Tectonic Model was then used to reconstruct the paleo-coordinates of these paleopoles.
Global Mean Poles (GMPs) describing the location of the Earth’s spin axis through time were determined at
intervals of 10, 20 and 50 million years, depending on the age of the plate tectonic reconstruction. Fisherian
statistics (Fisher 1953; Fisher et al., 1987) were calculated using the on-line tools at the paleomagnetism.org
website (Van Hinsbergen et al., 2015). Each paleopole was only used once when calculating the GMPs in order to
minimize the bias of poorly dated, long-lived paleopoles. Table 5 lists the Mesozoic and Cenozoic global mean
26
paleopoles in the reference frames of Africa, North America, South America, Europe, India, China, Australia, and
Antarctica. A complete list of the paleopoles used to calculate these GMPs is given in the Supplemental Materials.
Figure 10 plots the apparent polar wander path of the South Pole in African coordinates during the last 320 million
years. The locations of the paleopoles that comprise the consensus APW path (stars in Figure 10) are given in Table
5. These independently determined apparent polar wander paths are in good agreement. The southward motion
of Africa during the Jurassic was reversed in the early Cretaceous (JK Loop) and Africa moved steadily northward
during the Cretaceous and Cenozoic.
27
Figure 10 Apparent Polar Wander Path of the South Pole in African coordinates during the last 320 million years
(Carboniferous), stars = age in millions of years of mean path, the diameter of the star is proportional to the
uncertainty (A95); black path = Schettino and Scotese (2005); red path = Torsvik et al. (2008, 2012); light green
path= Scotese (2016); orange path = Scotese and Elling (2017), Elling (2022); dark green path = Vérard (2019); pink
path = Merdith et al. (2021); blue path = Vaes et al. (2023). Note that the locations of the South Pole at 200 Ma and
100 Ma are nearly coincident; the locations of the South Pole are also nearly coincident at 160 Ma and 120 Ma.
The blue path (Vaes et al., 2023) is more erratic because no smoothing factor has been applied. These zigs and zags
are a good representation of the fundamental uncertainty of the paleomagnetic data. The Cenozoic portion of the
APW path of Scotese (2016; green) diverges from the other paths. This is due to the fact that it is based on a
combined hot spot (mantle)/paleomagnetic reference frame. This divergence implies that either the Earth’s
magnetic field was not a perfect axial geocentric dipole during the Cenozoic, or that a rapid episode of True Polar
Wander has occurred during the last 10 million years. The dispersion of the paleomagnetic data during the
Cenozoic does not permit resolution of this dilemma. The gray dots and circles illustrate typical dispersion of
individual paleopoles (0 – 5 million years) and precision estimates (A95) of each paleopole.
5) Continental Tectonics. While most of the direct evidence for plate motions comes from the ocean basins, the
continents also provide important clues regarding the history of plate tectonic interactions. Five important
continental lines of evidence are: 1) the timing of continental rifting, 2) subduction-related volcanic activity along
Andean-type mountain ranges, continental island arcs and oceanic island arcs, 3) the obduction of ophiolites onto
continents due to the closure of both back-arc basins and major ocean, 4) the complex tectonic histories
associated with continent-continent, and 5) collision of exotic terranes along ocean-facing active continental. Some
of the key sources that were used to construct the history of continental tectonics are given in Table 6. These
references are listed in chronological order providing a concise history of the development of our ideas on these
topics. The lineations on the continents (Figure 8) map the trends of ancient and modern mountain ranges.
6) The Rules of Plate Tectonics. Unfortunately, one cannot produce plate tectonic reconstructions with a high level
of confidence that are based solely on available geological and geophysical evidence. This is especially true for the
distant past. Too much of the geological record is missing and the available geophysical data (e.g., paleopoles) are
often in conflict or can be interpreted in a variety of ways. To resolve this dilemma, we must rely on the insights
we have gained over the last 50 years regarding the dynamics of the plate tectonic process (e.g., Scotese and
Rowley, 1985; Scotese et al., 1988; Scotese, 1991; Scotese, 2014). The key insights are summarized in Table 7 as
the Rules of Plate Tectonics. A more complete description of the Rules of Plate Tectonics is provided with the
Supplemental Materials.
28
One of the most important insights we have learned regarding the evolution of plates and plate boundaries is the
nature of the driving forces that motivates the plates. Plates are driven primarily by two endogenic forces (i.e.
forces that arise from within the oceanic lithosphere): slab pull and ridge push. Only at a secondary level are plates
affected by the movement of the underlying asthenosphere. It is true that Hot Spots and Large Igneous Provinces
(LIP) can profoundly weaken the continental lithosphere and help to focus the forces that tear continents apart,
but hot spots do not drive plate motions. Approximately 70% - 80% of the driving force of plate tectonics comes
from the sinking of the cooler, denser oceanic lithosphere (slab pull and slab rollback).
“Ridge push”, on the other hand, is a bit of a misnomer. As a plate moves away from the mid-ocean ridge it cools
and becomes denser and heavier, creating lateral gradients in the gravitational potential energy (GPE). A plate
tends to gain weight as it approaches the trench; this added weight helps it to subduct faster. In a very real sense,
the plates drive themselves (Forsyth and Uyeda, 1975; Clennett et al., 2023).
Understanding the nature of plate tectonic driving forces allows us to properly motivate plate movements
(Clennett et al., 2020). Oceanic plates will tend to move toward subduction zones - pulled along by slab pull. The
plates attached to the greatest volume of subducting oceanic lithosphere will move the fastest (Forsyth and
Uyeda, 1975). Subduction is inexorable and subduction zones will continue to draw the plates together until the
oceanic lithosphere between them is entirely consumed. This eventually results in a continent-continent collision
(i.e., Wilson Cycle (Wilson, 1966; 1968; Burke and Dewey, 1975; Wilson et al., 2019) or Wegener Cycle (Nance et
al., 1988; Nance and Murphy, 2013), or the obduction of an island arc. Subduction is a reliable driving force, pulling
the plates ever onward and downward. Though the notion that plates are passively carried along by well-organized
mantle convection cells continues to be taught in schools and in the popular media, this explanation of plate
motions is incorrect and misleading.
Another important insight gained from our modeling of plate motions is that plate tectonics is a catastrophic
system (Rule 12; Figure 11). Most of the time plate geometries evolve in a slow and steady manner, but every once
and a while, seemingly out of nowhere, interplate stresses change rapidly and dramatically produce an entirely
different plate tectonic regime. The two most important plate tectonic events that can trigger global plate
reorganizations are: 1) continent-continent collision and 2) the complete subduction of a mid-ocean ridge.
These catastrophic plate tectonic events often trigger coincident Earth System-wide events such as rapid changes
in global sea level or abrupt shifts in climate, which in turn may cause global mass extinctions. Our modern
understanding of Earth System processes combines the uniformitarian model, which emphasizes that the present
is the key to the past (Lyell, 1830, 1832, 1833) with the catastrophic model, which characterizes Earth history as
long periods of stasis interrupted by short periods of catastrophic change (i.e., mass extinctions; Cuvier, 1831).
29
Figure 11. Plate Tectonics as a Catastrophic System (Scotese, 2014).
7) Synthetic Seafloor Spreading Isochrons. Plate tectonics is a grand recycling system. On average, it takes about
100 million years for the ocean floor created at a mid-ocean ridge to return to the mantle. Since all ocean floor is
created in a nearly symmetric fashion at a mid-ocean ridge, if half of the ocean floor is missing, we can reconstitute
the missing portion by creating a mirror-image of the remaining half. For example, the conjugate partners of the
old ocean floor in the central and western Pacific Ocean have been subducted. Synthetic isochrons representing
the missing ocean floor are shown as the unshaded isochrons in Figure 2A.
The recreation of synthetic ocean floor is at best an educated guess. It is unlikely that the missing ocean floor was
originally created in a perfectly symmetrical fashion. Also, this subducted ocean floor may have been broken into
smaller plates before it was subducted. This is what happened when the Farallon plate was subdivided into the
Cocos, Nazca (Wortel And Cloetingh, 1981; Lonsdale, 2005), and Kula plates (Engebretson et al., 1985) as it
approached the western American subduction zone. We must remember that we will never know the exact extent
of these vast, long-gone regions of subducted oceanic lithosphere.
The problem of reconstructing subducted oceanic plates is much more severe when conjugate, synthetic isochrons
are lacking (i.e., for times > 180 million years). In these cases, our reconstructions of long-subducted oceanic plates
is little more than guesswork.
30
8) Subduction Graveyards. Seismologists use shear and compressional waves to map the interior of the Earth (e.g.,
Spakman, 1988; van der Meer et al., 2018). The depths of important boundaries such as the inner core – outer
core boundary (5154 km), the core-mantle boundary (2889 km), the lower mantle – upper mantle boundary (660
km), the 410 km discontinuity (Kennett and Engdahl, 1991), the base of the continental lithosphere (up to 400 km;
Conrad and Lithgow-Bertelloni, 2006; Artemieva, 2006), as well as the base of the oceanic lithosphere ( low
velocity zone, ~150 km; Conrad and Lithgow-Bertelloni) are well-determined. In recent years, mantle tomographic
maps have been produced that identify regions of slow and fast velocities in the mantle that are thought to
represent less dense (warmer) and more dense (cooler) regions. It has been hypothesized that the cooler and
more dense material of the mantle correspond with the location of subducted oceanic lithosphere, or subduction
graveyards (van der Meer et al., 2018).
Figure 12 is a “time-lapse” image that illustrates the changing location of Cretaceous subduction zones. The
location of ocean trenches changes relatively slowly through time (~2 km/my). This is due to the inherent inertia of
the large, entrained volume of subducting oceanic lithosphere. Young subduction zones can move more quickly. A
phenomenon called slab rollback (Stern, 2002; Royden and Husson, 2006; Boutelier and Cruden, 2013) results
when oceanic lithosphere can freely move in a retrograde fashion in a direction opposite of convergent plate
motion. slab rollback produces back-arc basins (Uyeda and Kanamori, 1979; Barker and Omaston, 1981), whereas
trenches that are actively overridden by the adjacent continental plate produce Andean-style margins (Lamb,
2004).
Superimposed on the locations of Cretaceous trenches (blues line, Figure 12) are the corresponding, deeply buried
subduction zone graveyards (colored zones). These subduction graveyards encircle the Panthalassic Ocean and also
form a broad NW-SE trending belt that marks the former northern border of the Tethys Ocean. Other subduction
graveyards indicate that there may have been Cretaceous subduction beneath northern Africa and along the rim of
the western Pacific ocean basin (dashed lines, Figure 12; van der Meer, 2017).
The fact that most of these subduction graveyards closely match the predicted location of Cretaceous trenches
strongly suggests that there has been little or no True Polar Wander since the early Cretaceous, otherwise there
would be a demonstrable mismatch between our paleomagnetically derived reference frame and the mantle
reference frame. The lack of significant True Polar Wander during the last 150 million years, and possibly the last
300 million years, is discussed more fully later in the paper.
31
Figure 12. Location of Cretaceous trenches and subduction zone graveyards. The thin, dark blue lines mark the
successive location of subduction zone trenches of the Cretaceous (145 Ma – 65 Ma). Note how the location of the
trenches along western North and South America move progressively westward from 145 Ma to 65 Ma. The
trenches in the Southwest Pacific, northeast of Australia, moved northward; whereas, the location of the trenches
bordering the southern margin of Eurasia remained stationary throughout the Cretaceous. The colored bands
indicate the location of corresponding subduction zone graveyards of Cretaceous age. The labelled age indicates
when the lithosphere in that graveyard was subducted, i.e. the location of the trench at that time (van der Meer,
2017; van der Meer et al., 2018). The trends and age progression of the subduction zone graveyards, for the most
part, match the ages of the predicted trench locations.
9) Large Igneous Provinces (LIPs). Large Igneous Provinces, or LIPs, are the extrusive component of voluminous
mantle plumes or hotspots. Mahoney and Coffin (1997), Foulger and Jurdy (2007), Ernst (2014), Neal et al.(2015),
Ernst and Youbi (2017), Srivastava et al. (2022) , and Percival et al. (this volume) have documented over 200 LIPs
that occur as far back as the Archean. The statistics of more than 30 LIPs and hot spot tracks that were erupted
during the Cretaceous are listed in Table 8. Roughly half of these LIPS were erupted on the continents forming
extensive volcanic edifices (e.g. Deccan Traps); other LIPs, erupted on the ocean floor, formed large submarine
plateaus (e.g., Ontong Java Plateau; Erba et al., 2015). As one might expect, the largest LIP accumulations are
found on slow moving plates (e.g., Kerguelen Plateau; Erba et al., 2015). When plumes penetrate fast-moving
plates, they typically form large, long-lived, hot spot tracks (e.g., Ninety East Ridge). The breakup of large oceanic
plateaus, such as the Manihiki – Hikurangi (120 Ma), can provide important constraints on plate evolution (Seton
et al., 2012).
32
Large Igneous Provinces have had an important effect on paleoclimate and evolution. There is an extensive
literature describing the relationship between these massive volcanic eruptions and Global Warming (Larson,
1991; Kidder and Worsley, 2012; Godderis et al., 2012; Ernst and Youbi, 2017; McKenzie and Jiang, 2019; Scotese
et al., 2021; Percival et al. (this volume), as well as, the correlation between LIP eruptions and major extinction
events (Rampino and Stothers, 1988; Wignall, 2001, 2015; Morgan et al., 2004; Kidder and Worsley, 2010; Bond
and Wignall, 2014; Rampino and Self, 2015; Ernst and Youbi, N., 2017; Bond and Gasby, 2017; Clapham and Renne,
2019; Schobben et al., 2019; Suarez et al, 2019)
10) Paleobiogeography. Paleobiogeography describes the past geographic distribution of various fossil taxa.
Modern biogeographic patterns are complex and are controlled by a variety of factors including: temperature,
geographic barriers, ocean currents, and dispersal history (vicariance). Paleobiogeography provides important
constraints for pre-Pangean plate tectonics (Figure 13; Middlemiss et al. 1971; Hallam, 1973; Hughes, 1973;
McKerrow and Scotese, 1990; Harper and Servais, 2013;Harper et al., 2013; Torsvik and Cocks, 2017; Cocks and
Torsvik, 2020; Harper et al., 2023; Servais et al., 2023) but is less diagnostic for Mesozoic and Cenozoic plate
movements. McKenna (1973), Briggs (1987), Brown and Lomolino (1998) provide a good overview of the general
principles of biogeography and Lieberman (2000, 2003) discusses the how fossils and evolutionary theory are
applied to the subject of paleobiogeography.
Figure 13. Early Ordovician paleobiogeography with ocean currents. Faunal provinces blue = Laurentian, red =
Baltic, cyan = Siberian, green= China- Australia, orange = NW Africa-C. Europe - Avalonia, after Harper et al. 2013,
33
see also Harper et al., 2023; Servais et al., 2023). Brachiopod provinces are defined by temperature (latitude) and
ocean currents.:
Cretaceous paleobiogeography can be divided into the study of terrestrial and marine realms. The most useful
terrestrial components are dinosaurs and land plants. Dinosaurs were strictly terrestrial and could not easily cross
wide bodies of water (>200 km). They were, however, highly mobile and could migrate across vast tracks of land in
relatively short time intervals (1000’s of years). Consequently, early dinosaurs and other reptile groups
(Lystrosaurus and Cynognathus; Colbert 1973) were widely distributed across Pangea prior to the formation of the
Atlantic and Indian Oceans.
Throughout most of the Jurassic, the dinosaurs of North America-Europe, Asia, and Gondwana were recognizably
different (Kraus et al., 2019). As the oceans widened during the late Jurassic and early Cretaceous, dinosaurs
became progressively isolated and distinct regional groups were established (Upchurch et al., 2002). The dinosaurs
of East Gondwana (India-Australia-Antarctica) became distinct from the dinosaurs of West Gondwana (Africa –
South America) by the Valanginian (~130 Ma). As the South Atlantic widened the dinosaurs of Africa and South
America became endemic by the mid-Cretaceous (Albian; 110 – 100 Ma; Upchurch et al., 2002). The migration of
Cretaceous dinosaurs was somewhat been restricted by different climatic regimes (Köppen belts) as well as
oceanic barriers (Dunhill et al., 2016). Theropods and ornithischians seemed to have preferred middle and high
latitudes, whereas, sauropods show a preference for warmer subtropical latitudes (Chiarenza et al., 2022).
Though deep, widening ocean basins provided barriers to dinosaur migration, changes in sea level were probably
more effective, though more short-term, barriers (Jabri et al., 2010). For example, the Cretaceous mid-Continent
Seaway divided North America into western (Laramidia) and eastern (Appalachia) landmasses. Laramidia shared
many taxa with eastern Asia during the Cretaceous because a landbridge extended across the Bering Sea, whereas
Appalachia developed a less diverse, endemic dinosaur fauna. In a few special cases, hot spot tracks and large
volcanic edifices (e.g. Kerguelen plateau; Ali and Krause, 2011; Kraus et al., 2019) created landbridges linking
continental landmasses across widening ocean basins.
Plants were the other important component of the Cretaceous terrestrial realm. The first flowering plants
appeared in the Early Cretaceous (~140 million years ago; Willis and McElwain (2002). Their rapid diversification
during the mid and Late Cretaceous transformed the ecosphere and led to their global dominance by the end of
the Cretaceous. Because plants are not locomotive, they do not easily spread from one place to another and are
primarily adapted to local climatic conditions. The phytogeography of the Cretaceous can be described by seven
global plant biomes: Tropical everwet, Tropical summerwet, Subtropical desert, Winterwet, Mid-latitude desert,
34
Warm temperate, and Cool temperate (Horrel, 1991; Upchurch et al., 1999.) There was no tundra or subpolar
plant assemblage during the Cretaceous.
The marine faunas of the Cretaceous have long been subdivided into a warm-water “Tethyan” realm and a cool
water “Boreal” realm (Casey and Rawson, 1973). These biogeographic differences were recognized well-before the
advent of plate tectonics. Recent investigations of planktonic coccolithophorid taxa (Braarudisphaera bigelowii)
have hypothesized that distinct Cretaceous marine provinces were segregated by both temperature and evolving
ocean currents (de Lourdes-Fonseca et al., 2019; Figure 14).
Figure 14. Braarudisphaera bigelowii migration paths during the Maastrichtian. Reliability (weak to very good)
depends on the number of supporting observations (from de Lurdes Fonseca et al., 2019).
11) Paleoclimate. The advent of supercomputers has made global paleoclimate simulations more practicable, if not
routine (e.g., Lunt et al., 2016; Valdes et al., 2017, Haywood et al., 2019; Li et al., 2022; Lunt et al., 2023 ). The
increase in the number of simulations means that we now have multiple paleoclimate simulations for each stage of
the Cretaceous. These simulations now provide a baseline for modelling Cretaceous environments (Köppen Zones;
Figures 2C) as well as quantitative estimates of global temperature (Figure 3A), precipitation (Figure 3B), runoff,
ocean circulation (Figure 3C), upwelling, and more. For the most part, the results of these paleoclimatic models
compare favorably with the geologic record of coals, bauxites, evaporites, calcretes, glendonites, dropstones, and
tillites (Boucot et al., 2013; Bao et al., 2023; Burgener et al., 2023; Rogov et al., 2023). The latitudinal distribution
of climatically sensitive lithologies (Boucot et al., 2013) and the reconstruction of past Köppen belts can often
provide an important test of paleomagnetic estimates of paleolatitude (Bao et al., 2023; Burgener et al., 2023).
35
12) True Polar Wander. The continents move across the face of the globe because they are motivated by plate
tectonics. It is also possible, however, that a component of their latitudinal motion may be due to the rotation of
the entire Earth beneath the spin axis. This rotation of the entire Earth relative to the spin axis might have been
called “Whole Earth Rotation” or “Global Rotation”; however, for historical reasons it was called True Polar
Wander (Goldreich and Toomre, 1969). The term True Polar Wander (TPW) was coined to contrast it with
Apparent Polar Wander. Apparent Polar Wander refers to the paleomagnetic estimate of the motion of the spin
axis (north and south geographic poles) relative to a specific continent (Creer, 1954; see Figure 10). Each continent
has its own, unique Apparent Polar Wander (APW) path (see Table 5). If the entire Earth rotates with respect to
the spin axis, then this component of motion is shared by all the continents, hence it is “True Polar Wander”
(TPW). Paleomagnetic measurements cannot be easily used to differentiate between continental motions due to
plate tectonics and continental motions due to True Polar Wander because paleomagnetic data lack the precision
to do so.
A small amount of True Polar Wander undoubtedly occurs because the distribution of dense material within the
Earth (i.e., subduction zone graveyards) constantly changes. This redistribution of dense material affects the
Earth’s instantaneous moment of inertia which causes the Earth to shift slightly to rebalance the load (Gold, 1965).
Over 10’s or 100’s of millions of years these global readjustments may add up to 100’s of km (Steinberger and
Torsvik, 2010; Vaes, 2023).
In summary, the average rate of TPW (mm/yr) is a fraction of the average rate of plate movement (cm/yr) and is
likely to be below the resolution of the paleomagnetic data (Vaes, 2023). The large, rapid, and often recursive
episodes of TPW that have been proposed by some researchers (Kirschvink, 1997; Evans, 2003; Steinberger and
Torsvik, 2010; Muttoni et al., 2013; Eglington et al., 2017; Muttoni and Kent, 2019; Evans et al., 2022; Le Pichon et
al., 2022), we believe, are based on an insufficient number of paleopole determinations (N < 50).
3. Geological and Geophysical Evidence Used to Map Ancient Paleogeography
In its strictest sense, the term paleogeography describes not only the past positions of the continents but also the
location of ancient shorelines, shallow seas, deep oceans, land and mountains. A complete paleogeographic map
also models the paleotopography of the land surface and paleobathymetry of the ocean basins (Figure 2B). The
digital basis of the paleogeographic maps shown here is a paleo-digital elevation model or paleoDEM (Scotese and
Wright, 2018). Each paleoDEM is composed of over 6 million grid cells that capture digital elevation information at
a 10 km x 10 km horizontal resolution and 40 m vertical resolution. Paleo-DEMs allow us to visualize and analyze
the changing surface of the Earth using GIS software and other computer modeling techniques. For example, we
36
can use a paleoDEM to calculate the average elevation of the land surface, the average depth of the ocean floor,
the area of the continents flooded by the ocean, or the relative area of continental versus oceanic crust. The
paleoDEMs that are the basis of the Cretaceous paleogeographic maps are provided in the Supplemental
Materials.
To construct an ancient paleogeography, the geological lithofacies that define the ancient depositional
environments must be mapped. For example, a thick sequence of pure limestones might represent warm, shallow
water environments like the Bahamas Platform or a vast epeiric sea. Extensive sets of massive, cross-bedded
sandstones may once have been wind-blown, desert dunes. A terrane composed of andesite and granodiorite may
have been a continental arc or Andean mountain range. Table 9 (Ziegler et al., 1985) summarizes the lithofacies
and rock types that correspond to the depositional environments that have been used to interpret the ancient
topography and bathymetry.
Geologists have been collecting lithologic information and producing lithofacies and paleoenvironmental maps for
more than 200 years (William Smith, 1815). During the late 1970’s and early 1980’s, the Paleogeographic Atlas
Project, under the leadership of Prof. A. M. Ziegler, in the Department of Geophysical Sciences, University of
Chicago, compiled a database of more than 125,000 lithological and paleoenvironmental records for the Mesozoic
and Cenozoic (Ziegler, 1975; Ziegler & Scotese, 1977; Ziegler et al., 1985). This database - supplemented by the
Phanerozoic reef database of Flugel and Kiessling (2002)and Kiessling et al. (2002), a database of climatically
sensitive lithofacies (Boucot et al., 2013), and environmental information from the Paleobiology Database (Alroy et
al., 2008) - was used to produce the Cretaceous paleogeographic maps in this review. This digital paleogeographic
information was supplemented by more than 50 sources that feature Cretaceous paleogeographic maps (Table
10).
Information from these four lithologic databases was standardized, sorted by time interval, and plotted in the
simplified fashion shown in Figure 15 (Scotese, 2021). These maps show the control points for the paleogeographic
interpretations. In these simplified maps, the paleocoastlines (black lines) weave their way between geologic data
indicating marine environments (circles) or terrestrial environments (plus signs). The size of the symbol is
proportional to the duration of the stratigraphic interval; i.e., small symbols are better dated than large symbols.
37
Figure 15. Control points used to map the location of the paleocoastline during the early Jurassic (200 Ma) , blue
circles – marine deposits, red plus signs – terrestrial deposits. Smaller symbols have better age constraints (from
Scotese, 2021).
Lithologic data can only be used to map paleogeographic environments where the rock record is fairly complete.
However, there are many instances where the rock record has been eroded, destroyed by tectonic processes, or
covered by younger strata. For these areas, a second, more interpretive approach was taken to restore the
paleogeography. In these instances, the paleoenvironments and paleogeography must be inferred from the
tectonic history of a region. As discussed in the previous section, the PALEOMAP Global Plate Tectonic Model
(Scotese, 2016) provides the tectonic framework required to make these inferences and interpretations. The plate
tectonic reconstructions (Scotese, 2017, 2018) are used to “model” the expected changes in topography and
bathymetry caused by plate tectonic events, such as: sea floor spreading, continental rifting, subduction along
active margins, and continental collision, as well as other isostatic events such as glacial rebound (Peltier, 2004).
For example, to produce a paleogeographic map for the early Cretaceous, young tectonic features, such as recent
uplifts or volcanic eruptions (e.g. Mid-African Rift), must be removed or reduced in size, whereas older tectonic
features, such as ancient mountain ranges (e.g. Appalachian mountains), must be restored to their former extent.
This approach is similar to the more recent dynamic techniques described by Vérard et al. (2015), Baatsen et al.
(2015), Vérard (2019), Markwick (2019), van der Linden (2020), Poblete et al. (2020) , and Boschman et al. (2022).
In a similar manner, the paleobathymetry of the ocean floor must be restored back through time. As ocean floor
moves away from a spreading ridge, it cools and subsides. In many respects restoring the past bathymetry of the
ocean floor is much easier than estimating the elevation of ancient mountain ranges (Rowley et al., 2001; Rowley
and Currie, 2006; Rowley and Garzione, 2007). The amount that the sea floor subsides through time follows a
38
regular mathematical rule where the amount of thermal subsidence is inversely proportional to the square root of
the age of the oceanic crust (Parsons and Sclater, 1977; Sclater et al. 1980). To restore the ancient ocean floor to
its former depths, the bathymetry of the ocean floor was adjusted using the depth/age relationship originally
published by Stein & Stein (1992) and later updated by Rowley (2018).
Once the paleogeography for each time interval has been mapped and corrections to the topography and
bathymetry have been duly noted, this information was converted into the aforementioned elevation model, or
“paleoDEM”(Scotese and Wright, 2018). By substituting a modified color look-up table, the paleogeography can
also be subtly modified to represent eustatic changes in sea level (Vail, 1977a,b; Hallam, 1984; Ross & Ross, 1985;
Haq et al., 1987, 1988; Algeo and Seslavinsky, 1995; Müller et al., 2008; Haq and Schutter, 2009; Miller et al., 2005;
Snedden and Liu, 2010, 2011; Vérard et al., 2015; Haq, 2014; Simmons, 2012, 2020; Van der Meer et al., 2017,
2022; Davis and Simmons, 2023).
For a more detailed discussion of the data and methods used to produce paleogeographic maps, see Scotese
(2021). A series of tutorials that describe how to build digitial paleogeographic maps from scratch can be found in
the Supplemental Materials.
4. Eustatic Changes in Sea Level Derived from Estimates of Continental Flooding
1) Introduction
In general, sea level was much higher during the past (Simmons, 2012, 2020). The modern era is characterized by
recent continental collisions (e.g., India-Asia, Africa-S. Europe, Arabia – Turkey and Iran) that have raised the
landscape and by large continental ice sheets that have removed water from the ocean basins and dropped sea
level. During the Cretaceous, based on our estimates, sea level was relatively low (+50-70 m) during the earliest
and latest parts of the period, and highest during the mid-Cretaceous (Cenomanian – Turonian, +150m; this study).
Changes in sea level can drastically change the paleogeography of a continent. This section describes the
methodology used to estimate changes in sea level during the Cretaceous.
The surface of the continents, however, is not flat. It gently slopes upward from the coastline across the broad
continental interiors (~.4 m/km), and then rises rapidly in the foothills of the mountains (~1.2 m/km). Figure 16 is a
simplified diagram that shows how the average slope of the present-day continents varies from about ~400 meters
above sea level to ~400 meters below sea level. This brackets the likely range of potential sea level change during
the Phanerozoic.
39
Figure 16. Continental flooding curve. The solid black line represents the modern continental gradient. The dashed
black line represents the reduced continental gradient during the Devonian (middle Paleozoic). This curve is similar
to a hypsometric curve that plots the area of the Earth represented by various elevations.
The vertical axis in Figure 16 is sea level. Zero meters indicates the location of the modern coastline. Note that the
horizontal axis is not distance, but rather the percentage of the global surface area that is land. As sea level rises,
continents are progressively flooded. For the modern world, land covers ~30% of the globe. If we were to
completely melt both the Greenland and Antarctic icecaps, sea level would rise ~70 meters (Clark, 2016) and the
global land area would be reduced to ~26.5%. During the height of the last ice age (Last Glacial Maximum, 20 - 18
ky), massive continental ice sheets withdrew 60 million km3 of water from the oceans, enough to drop sea level
120 meters (Gornitz, 2007). Correspondingly, at that time, land covered ~33% of the Earth’s surface (Figure 16).
During the past 540 million years, according to our paleogeographic models, the percentage of land area has
varied from a minimum of ~14% during the Early – Middle Paleozoic (500 – 400 Ma)) to a maximum of ~33% during
the Last Glacial Maximum (20 ky).
The solid line in figure 17 plots the changing percentage of land area during the last 540 million years. The dashed
line, which is the mirror image of land area, is the percentage of the continents covered by shallow sea. There are
four other time intervals when the continents were as emergent as they are today: the Eocene – early Miocene (45
- 15 Ma), the Early Cretaceous (140 Ma), and the Middle Jurassic (180- 160 Ma), and the Early – Middle Triassic
40
(250 – 220 Ma). It is interesting to note that though sea level fell rapidly during the Permo-Carboniferous Ice Age,
global sea level was relatively high (~ +80 m) when compared to the present-day.
Figure 17. Continental Flooding. The percentage of global area that is land (solid line) and continent covered by
shallow seas (dashed line). Important eustatic events are labelled with numbers 1-16.
In this regard, the present is not the key to the past. During the past 540 million years the continents, more often
than not, have been flooded by the oceans. According to the model, the times of relatively high continental
flooding were: the early – middle Paleogene (66 – 50 Ma), the mid-Cretaceous (100 – 80 Ma), the Aptian – Albian
(130 – 110 Ma), and throughout the early and middle Paleozoic (500 – 400 Ma). It is interesting to note that an
important cross-over occurred 360 million years ago (Figure 17). At that time, the shallow seas began to drain
away from the continents and the percentage of land gradually increased. It is probably no coincidence that this
late Devonian transition also marks the evolution of land-dwelling tetrapods, the first extensive tropical
rainforests, and the rapid expansion of the terrestrial ecosystem.
2) Estimating Past Sea Level from Changes in Continental Flooding
The seas go in and the seas go out because the level of the oceans (i.e. sea level) changes with respect to the land.
Though most of these forces act regionally, their combined effect is global. Global changes in sea level is called
41
“eustasy” (Suess, 1908; Chamberlin,1898, 1909; Grabau, 1924; Wagreich et al., 2014). Eustatic changes in sea level
are globally contemporaneous and provide an important stratigraphic calibration tool. The episodic pattern of the
seas coming in and the seas going out provides a standard that allows us to correlate far flung geologic and
evolutionary events. This recurrent change in sea level is at the core of the study of stratigraphy, biostratigraphy,
sequence stratigraphy (Sloss, 1963; Vail et al., 1977a, b; Simmons, 2012,2020; Davies and Simmons, 2023); long-
term eustatic sea level change may be the most important driver of global climate change (Fischer, 1981, 1982,
1984; Berner, 1994, 2004; Berner et al., 1983; Berner and Kothavala, 2001).
One approach to estimating changing sea level would be to model all of the various factors that can cause sea level
to rise or fall. This approach, however, is fraught with difficulties, uncertainties, and complexities (Simmons 2012,
2020). We have chosen a simpler, more direct way to estimate sea level change. High sea level, whatever the
cause, floods the continents. If we can accurately measure the amount of continental flooding through time, then
we can use this measurement to retrodict past sea levels. This section of the paper describes how we used
continental flooding to produce a novel Phanerozoic sea level curve.
As illustrated in Figure 16, there are several sea level calibration points along the modern continental flooding
curve (solid black line): Last Glacial Maximum (-120 m, 33% land), present-day sea level (0 m, 30% land), and no
polar ice caps (70 m, 26% continental area). We can use measurements of land area obtained from the
paleogeographic maps to estimate past sea levels. For example, during the late Cretaceous (80 Ma), shallow seas
covered most of the continents and the land area was reduced to ~21%. Using the modern continental flooding
(i.e. hypsometry) curve, this change in land area corresponds to a sea level rise of ~220 m (Figure 16). Maximum
continental flooding occurred during the early and middle Paleozoic (500 – 400 Ma)million years ago. At that time,
land areas accounted for only 16% of the Earth’s surface. If we used the present-day flooding curve, we would
predict that sea level during the early and middle Paleozoic was nearly 400 m higher than present-day sea level!
This, however, is clearly an overestimate. We need to improve our predictions of past sea level by taking into
account how the average elevation of the continents has changed through time.
The Earth’s surface is constantly changing. Mountain-building pushes the continents upward which steepens the
continental flooding curve; rift and drift processes stretch the continents which flattens the continental flooding
curve. Today’s topography, unfortunately, is not a good analog for past topographies. The continents have been
recently uplifted by Andean-style mountain building and a series of Cenozoic continent-continent collisions. The
average elevation of the modern land surface is nearly 600 meters (Figure 19). This is as high as or higher than any
other time during the last 540 million years. In order to use continental flooding to estimate past sea level, we
need to produce continental flooding curves that better match the continental topographies of these past times.
42
The best estimate of past continental topography is the changing average elevation of continental land areas
through time. The average elevation of the land surface through time can be directly measured from the
paleogeographic maps and the corresponding paleo-digital elevation models (paleoDEMs). Figure 18 illustrates the
average elevation above sea level of the continents during the Phanerozoic.
Figure 18. Average elevation of the land surface above sea level during the Phanerozoic. The solid black line is the
10 million year moving average.
Major peaks occur during times of continental collision and most of the sharp drops take place during times of
continental extension associated with the early phases of ocean basin formation. This curve has two peaks.
Present-day topography is the highest peak (~600m). A second peak spans 100 million years, and represents the
uplift of the continents due to the continent-continent collisions that formed Pangea (380 Ma -280 Ma). Prior to
the formation of Pangea, continental topography was much more subdued. During the early and mid-Paleozoic
(540 – 400 Ma) and during most of the Mesozoic, the average height of the land surface was only ~400 meters,
43
two-thirds of the modern value. During the Cenozoic continental elevations trended steadily higher, peaking in the
last 30 million years following the collision of Africa with Europe and India with Asia.
As noted earlier, the dashed line in Figure 16 is the continental flooding curve based on the topography of the
continents during the mid-Paleozoic. The mid-Paleozoic flooding curve is similar in shape to the modern
continental flooding curve but the curve crosses zero meters (the coastline) at a level that corresponds to a land
area of 26%, rather than 30%. In the mid-Paleozoic, because the continents were so low-lying, even a modest rise
in global sea level was sufficient to inundate much of the continents. For example, a 200 m rise in sea level during
the mid-Paleozoic would have been sufficient to further reduce the land area to 15% (Figure 16).
The modern continental flooding curve and the middle Paleozoic flooding curve bracket the range of expected
paleotopographic variation of continental lowland areas during the Phanerozoic. We have used these maximum
and minimum limits to construct a Phanerozoic global sea level curve (Figure 19). By taking into account the
reduced topography of earlier time periods, the amount of sea level change required to flood the continents is
greatly reduced. For example, during the mid-Cretaceous, the average elevation of the continents was ~400 m, or
~66% of the modern average elevation. This means that the amount of sea level rise during the mid-Cretaceous
that was required to flood the same continental area was reduced from 210 m to 140 m (a 33% reduction).
44
Figure 19. Phanerozoic Global Sea Level derived from continental flooding and the changing elevation of the
continents. Dashed line = estimate of sea level change based on modern hypsometry. Solid line = estimate of sea
level that takes into account reduced continental hypsometry.
In a similar fashion, because of the reduced Paleozoic topography, the predicted mid-Paleozoic sea level maximum
of 400 m was reduced to a more plausible, but still elevated level (+260 m). Taking into account the changing
elevation of the continents (Figure 18), we have estimated changes in Phanerozoic sea level (Figure 19) that are
predicted by the changing degree of continental flooding (Figure 16).
In Figure 20, the Phanerozoic sea level curve derived from the paleogeographic maps is compared to other
published sea level curves estimates (Vail, 1977a,b; Hallam, 1984; Algeo and Seslavinsky, 1995; Müller at al.,
2008a; Miller et al., 2005; Snedden and Liu, 2010, 2011; Vérard et al., 2015, van der Meer et al., 2014, 2022). All of
the Phanerozoic sea level curves can be described as a double-humped, with peaks in the early-middle Paleozoic
and the Cretaceous-early Cenozoic). The lowest sea level occurs during the Triassic and Jurassic between 240 - 160
Ma (<100 m) and the late Cenozoic (50 – 0 Ma).
45
Figure 20. A Comparison of Phanerozoic Sea Level Curves; blue – Hallam (1984), blue dashes – Vail (1977), ), green
– van der Meer et al. (2022), green dashes (540 – 250 Ma) – Algeo and Seslavinsky (1995), red – Snedden & Liu
(2010), red dashes – Vérard et al. (2015), black dashes (0 -140 Ma) – Müller et al. (2008), black dots (0 – 100 Ma) –
Miller et al (2005), black – this study.
Our estimate of Cretaceous sea level is in good agreement with Müller et al. (2008a) and Miller et al. (2005) sea
level curves. These curves represent the most conservative estimate of sea level change (< 200 m) during the last
200 million years. Our curve is in the mid-range of the other curves for Triassic low-stand. The Paleozoic portion of
our curve is similar to the Vail (1977a,b) curve, and substantially higher than all other curves except Hallam (1984
In summary, our estimate of Cretaceous sea level change based on the degree of continental flooding suggests
that sea level was highest (+150 m) during the Cenomanian-Turonian (90 Ma – 80 Ma), with a subsidiary highstand
during the early Aptian (120 Ma). Sea level was lowest during the earliest Cretaceous (+10 m) and at the end of the
Cretaceous (+25 m).
5. Modeling Cretaceous Paleoclimate: Computer Simulations and Lithologic Indicators of Climate
1) Introduction
46
Cretaceous paleoclimate simulations have progressed a long way from the first computer models of Barron et al.,
(1981); Barron and Washington (1982a,b); Barron, (1983, 1984); and Barron and Washington (1984,1985). During
the intervening 40 years, a multitude of simulations of Cretaceous climate have been run, for example: Schneider
et al., 1985; Crowley et al., 1986; Horrel, 1991; Barron et al., 1995; Valdes et al., 1996; Herman and Spicer, 1997;
Bush and Philander, 1997; Sloan and Pollard, 1998; Price et al., 1998; Upchurch et al., 1999; Poulsen et al., 1999;
DeConto et al., 1999, Beerling and Woodward, 2001; Poulsen et al., 2001; Markwick and Valdes, 2004; Poulsen et
al., 2007; Zhou et al., 2008; Scotese et al., 2007, 2008; Flogel et al., 2011; Hasegawa et al., 2012; Hay and Flogel,
2012; Donnadieu et al., 2016; Ladant et al., 2020; and Landwehrs et al. 2021. These advances are attributable to a
better understanding of the complexities of the Earth’s climate system, a substantial increase in computer
processing power (Moore’s Law), a wealth of digital measurements of the modern climate, and a growing
community of scientific researchers using advanced global climate simulations (IPCC, 2007; 2018; 2019)
These advances in computational climate modeling are now complemented by a growing data base of lithologic
indicators of past climates such as coals, bauxites, coral reefs, evaporites, calcretes, tillites, glendonites, and
dropstones (Habicht, 1979; Hallam, 1984; Ziegler et al., 1987; Frakes et al., 1992; Sellwood et al., 1994; Chumakov,
(1995, 1997); Parrish, 1998; Markwick, 2007; Boucot et al., 2013; Burgener et al., 2023). In addition, geochemical
proxies for paleotemperature (Veizer and Hoefs, 1976; Veizer, 1995; Veizer et al., 1999; Prokoph et al., 2008;
Grossman et al.2012a&b; Veizer and Prokoph, 2015; Henkes et al., 2018; Grossman and Joachimski, 2020;
Grossman and Joachimski, 2022; Gaskell et al., 2022; Judd et al., 2023 ) and estimates of the ancient atmospheric
concentration of the greenhouse gas CO2 (Royer et al., 2004; Foster et al., 2017; Milles et al., 2019; Rae et al.,
2021; Hönisch et al., 2023) in combination with novel statistical techniques (Bayesian statistics; Tierney and
Tingley, 2014, 2015; Judd et al., 2023; Burgener et al., 2023) are providing a clearer vision of climate change in
deep time.
2) Computer Simulations of Paleoclimate
The estimates of past temperature, precipitation, and oceanic circulation presented in this study were produced by
the Hadley Center Model (HadCM3) developed by Paul Valdes at the University of Bristol (Valdes et al., 2017;
Valdes et al., 2021). HadCM3,is a coupled atmosphere-ocean-vegetation model with a horizontal resolution of
3.75˚ (longitude) x 2.5˚ (latitude). The atmospheric simulation has 19 levels and the oceanic simulation has 20
levels. A more detailed description of the workings of the HadCM3 model can be found in Pope et al. (2000),
Gordon et al. (2000), and Valdes at al. (2017). Recent computer simulations of Cretaceous climate are also
available for review (Lunt et al, 2016; Valdes et al., 2017; Haywood et al., 2019; Valdes et al., 2021; Li et al., 2022;
Willeit et al., 2022; Lunt et al., 2023).
47
In order to simulate Cretaceous climates the HadCM3 model requires four additional input parameters (Valdes et
al., 2021): 1) the topography of the land surface and the bathymetry of the ocean floor during the Cretaceous, 2)
an estimate of the amount of insolation received by the Earth during the Cretaceous, 3) an estimate of the
changing amount of atmospheric CO2 during the Cretaceous (Foster et al., 2017; Scotese et al., 2022), and 4) the
preindustrial concentration of ozone, which varies as a function of latitude (Beerling et al., 2011). One should also
note that the HadCM3 model conserves the combined volume of water in the atmosphere and in the oceans.
a. Paleogeography. The 17 Cretaceous paleogeographic maps used in the HadCM3 simulations are derived from
the PALEOMAP Paleogeographic Atlas (Scotese, 2016; Scotese and Wright, 2018; Scotese, 2021). The original high-
resolution elevation grid (.1˚ x .1˚) was reduced to a ~111 km x ~111 km (1˚ x 1˚) grid. These data were re-gridded
to 3.75o x 2.5o resolution that matched the GCM using a simple area or volume conserving algorithm. The
bathymetry was lightly smoothed using a simple binomial filter to ensure that the ocean properties were
numerically stable. The areas at high latitudes had this filter applied multiple times. This gridding procedure
sometimes produced single oceanic grid points surrounded by land grid points, particularly along complicated
coastlines. These oceanic singletons were manually removed. Similarly, important ocean gateways were
sometimes widened to ensure that the regridded coastlines permitted the free-flow of oceanwaters.
The paleogeographic reconstructions also include an estimate of land ice area (Scotese and Wright, 2018). These
were converted to GCM boundary conditions assuming a simple parabolic shape to estimate the ice sheet height
(Valdes et al., 2021). Unlike the Cretaceous paleoclimate simulations of Lunt et al. (2016) which had no polar ice
caps, the Cretaceous reconstructions in this study suggest that there may have been small areas of polar land ice
during the earliest Cretaceous (< 4 million km2, about twice the area of Greenland).
b. Insolation. The amount of energy received from the Sun during the Cretaceous was approximately 98.4% of the
modern value. This is based on the model of Gough (1981) which estimates that solar insolation has increased 30%
since the formation of the solar system at a rate of 1.5˚ C per 100 million years. Due to the lower level of insolation
the mid-Cretaceous should have been 1.6˚ C cooler than the present-day. This, of course, was not the case because
levels of atmospheric CO2, on average, were three times higher during the Cretaceous (Cretaceous CO2 = 1140
ppm, Pre-Industrial CO2 = 380 ppm; Mills et al., 2019; Scotese et al., 2022).
48
c. Atmospheric CO2
Global Climate Models, such as CESM (UCAR) and the HadCM3 (University of Bristol, U.K.), require estimates of
atmospheric CO2 to modulate the temperatures produced by these simulations. Therefore, an accurate estimate of
the ancient concentration of atmospheric CO2 is essential to successfully model past climates.
There have been only a few studies that have attempted to describe the variation in atmospheric CO2 during the
past 540 million years. Figure 21 illustrates estimates of the changing level of CO2 during the Phanerozoic (Foster
et al, 2017; Scotese et al.2022), with two additional high-resolution curves for the Cenozoic (Rae et al., 2021;
Hönisch et al. 2023). The Phanerozoic CO2 curves suffer from gaps in the CO proxy record and time intervals when
the range of CO2 proxy estimates vary widely.
49
50
Figure 21. Comparison of Phanerozoic Global Average Temperature with Phanerozoic CO2 Levels. (top) solid line -
Global Average Temperature, dashed line - 40 million year running average, B. Estimates of Atmospheric Level of
CO2 , dotted line - (Scotese et al., 2022), red line – Foster et al. (2017), thin black line (Cenozoic only) – Rae et al.
(2021), thin blue line (Cenozoic only) – Hönisch et al. (2023), black dots – CO2 proxy data binned in 1 my intervals,
open black dots - rejected proxy data. The rectangles at the bottom of the figure indicate alternating warmer
(black) and cooler (white) global temperature intervals. This vertical gray line connects CO2 peaks with warmer
intervals (Scotese et al., 2021).
These data gaps are especially large in the early and middle Paleozoic (460 Ma – 320 Ma), the late Permian and
early Triassic (270 Ma – 230 Ma), and the mid-Jurassic (180 Ma – 150 Ma). During the Cambrian and early
Ordovician (540Ma – 460 Ma), there were no land plants and few kinds of plankton. For these reasons, there are
no reliable proxy data for early Paleozoic. It should be noted that proxy values for CO2 are quite variable during the
early Permian (360 ppm – 1440 ppm), the Triassic (400 ppm – 3240 ppm), and the Paleocene (360 ppm – 1200
ppm) (Foster et al., 2017).
In order to improve the estimates of the changing levels of atmospheric CO2 , we have used our growing
understanding of the temperature history of the Phanerozoic (Scotese et al., 2021) to refine and update the Foster
CO2 curve. The sequence of black and white intervals along the time axis refers to times of relative global warming
(black) and global cooling (white). Only 50% of the CO2 peaks in the Foster curve correspond with warm intervals.
Assuming CO2 is the predominant cause of global warming, it is reasonable to make minor adjustments to the CO2
curve so that peaks in the CO2 curve correspond with times of maximum warming.
Using our updated estimate of Phanerozoic temperatures (Scotese et al., 2021), we can now fill in the gaps and
reject contradictory CO2 data. For example, during the Triassic and the late Cretaceous the CO2, estimates less
than ~1080 ppm were rejected (open circles; Figure 21). At first, this may seem like circular reasoning. No CO2
proxies, however, were used to determine the cool and warm intervals (Scotese et al., 2021). We are simply
redrawing the CO2 curve so that it conforms to what we independently know about global temperature change.
These adjustments significantly improve the fit between the revised CO2 curve and the CO2 proxy data.
51
There are still large gaps in the CO2 proxy record, consequently, the estimation of CO2 concentration is speculative
(e.g., Cambrian, early Triassic, Cenomanian-Turonian). This curve, however, represents a testable hypothesis that
awaits confirmation or rejection with additional proxy data.
In recent years, a new modeling approach has been developed with regard to the use of CO2 in computer
simulations (Li et al. 2022; Lunt et al., 2023) Rather than rely on estimates of CO2 based on proxy data (Figure 21),
the CO2 level is derived or tuned from the expected global temperature. In other words, the global average
temperature is assumed to be known and the level of CO2 required to achieve that temperature is then derived,
secondarily. This reverse engineering of CO2 levels, requires a well-established and credible Phanerozoic
Temperature Curve (Figure 21A; Scotese et al., 2021).
3) Lithologic Indicators of Climate and Köppen Belts
An alternate approach to paleoclimate modeling uses geological and paleontological data to directly reconstruct
the climate of the past (Habicht, 1979; Ziegler et al., (1979, 1981); Parrish et al., 1982; Sellwood and Price, 1994;
Parrish, 1998; Hart, 2000; Gibbs et al., 2002; Rees et al., 2002; Sellwood and Valdes, 2006; Markwick, 2007; Boucot
et al., 2013; Scotese et al., 2021; Burgener et al., 2023). The simplest approach uses lithologic indicators of climate
(e.g., coals, bauxite, coral reefs, evaporites, calcretes, kaolinites, tillites, glendonites, and dropstones; see Boucot
et al., 2013 for more information) to map ancient climatic belts.
Using modern temperature and rainfall records, we can map five major and eight minor climatic zones called
“Köppen Climate Belts” (Figure 22 ). The Köppen Climate Belts are largely defined by seasonal variations in
temperature and precipitation (Köppen, 1918, 1936). These variations give rise to regional climates and create the
mosaic of diverse environments that cover the Earth. These environments include: (A) tropical rainforests near the
Equator (dark green), (B) desert belts at subtropical latitudes (yellow-tan) that transition into (C) warm temperate
grasslands and forests (light green). In the modern world, as we move poleward, warm-temperate regions are
replaced by (D) seasonally warm/cold temperate regions (purple) and (E) finally frigid polar regions (blue). Each of
these climatic zones is characterized by a distinctive flora, fauna, land-cover, and depositional environments.
52
Figure 22. Modern Köppen Belts (Burgener et al., 2023)
The principal lithologic indicators of climate - coals, evaporites, and glacial deposits - can be used to map
paleoKöppen Climate Belts for ancient time periods (Ziegler et al., 2003; Boucot et al., 2013). Other important
lithologic indicators of climate are: soil minerals such as bauxite, an aluminum ore, which forms in warm, wet
climates; calcrete, or caliche, which forms in semi-arid regions; and kaolinite, which forms in regions with climates
that are sometimes wet and sometimes dry (warm temperate climate belt). Dropstones, like tillites, are important
indicators of frozen lakes or sea ice. A glendonite (Rogov et al., 2023) is a pseudomorph of ikaite, a low
temperature, hydrated polymorph of CaCO3 that forms at temperatures <4˚ C. The legend inset on Figure 23
summarizes the association of the various lithologic indicators of climate with warm/cool and wet/dry
environmental conditions.
The paleogeographic distribution of bauxites, in particular, helps us understand Cretaceous climates. Bauxites, as a
general rule, reflect tropical-subtropical humid, monsoonal conditions. Their modern occurrence is almost entirely
restricted to the Equatorial Wet Belt. The occurrence of bauxite deposits in northern Europe and Siberia during the
late Jurassic, Cretaceous, Paleocene, and Eocene times (Boucot et al., 2013) is one of the strongest geological
indications of warm and wet conditions at high latitudes.
53
Over the past 20 years, a global database of over 15,000 lithologic indicators of climate was assembled (Boucot et
al., 2013; see Supplemental Materials). For a thorough discussion of both lithologic and biological indicators of
climate, see Parrish, 1998; Boucot et al., 2013; and Cao et al., 2019.
54
Figure 23. A. Mid-Cretaceous (100 Ma) and B. Early Permian (280 Ma) lithologic indicators of climate and paleo-
Köppen belts (Boucot et al., 2013).
An icehouse world is simply defined as a time when the Earth is covered by permanent ice at either pole. For
permanent ice to accumulate in the polar regions (>67˚ N and S), the temperatures must remain below freezing
during the summer months. In other words, the global average temperature (GAT) must be less than 18˚C and the
average annual temperature of the polar region must be below -10 ˚C. Though the tropics remain warm (26˚C) in
an icehouse world, the polar regions are frigid (-10˚C to - 50˚C).
There also have been times in the past when there was no ice above the polar circle – even during the winter (e.g.,
Late Cretaceous, Figure 23B). During these hothouse times, the average temperature of the Earth was generally
above 20˚C (68˚F), and the polar regions were relatively warm (5˚C to 15˚C), and no ice could accumulate. It is a
well-established fact that no polar ice existed during the Paleocene-Eocene Thermal Maximum (55.6 Ma;
McInerney and Wing, 2011 ) or the Cenomanian-Turonian Thermal Maximum (93 Ma; Ziegler et al., 1985).
Despite these limitations, the Köppen approach does provide another important bit of information. This procedure
describes how the Pole-to-Equator temperature gradient has changed through time. The relative widths of the
equatorial wet belt and the subtropical arid belt do not change significantly through time because they are
55
controlled by Hadley Cell Circulation (Ziegler et al., 2003). The changing Pole-to-Equator temperature gradient is
due almost exclusively to the changing width of the Warm Temperate, Cool Temperate, and Polar Belts.
In icehouse worlds, like the present-day, Pole-to-Equator temperature gradient is very steep. The temperature falls
0.75˚ - 1˚ C per degree of latitude as we move towards the Pole (e.g., if we start at 30˚C at the Equator, we end up
with temperatures of -40C˚ to -60˚C at the pole). During hothouse worlds (e.g., Cenomanian-Turonian Thermal
Maximum, 93 Ma), the pole-to-Equator temperature gradient was much shallower, approximately 0.20˚ - 0.33˚ C
per degree of latitude. That means that if we start out at 30˚C at the Equator, the temperature at the Pole would
still be well above freezing (0˚ to 12˚C).
4) Combining Lithologic Indicators of Climate with Quantitative Paleoclimate Proxies
Recently, a more quantitative approach has been taken to estimate the actual range of paleotemperatures and
paleoprecipitation (Tierney and Tingley, 2014, 2015; Judd et al., 2023; Burgener et al., 2023). This approach assigns
a precise numerical range of temperature and precipitation to specific lithologic indicators of climate (Zhang et al.,
2016) and combines them with more traditional quantitative paleoclimate proxies such as various isotopic and
molecular systems (e.g., δ18O, clumped isotopes, and TEX86; Veizer et al., 1999; Grossman et al.2012a&b; Veizer
and Prokoph, 2015; O’Brien et al., 2017; Henkes et al., 2018; Song et al., 2019; Grossman and Joachimski,
2020,2022; Gaskell et al., 2022; Judd et al., 2023).
Using Bayesian statistics, the lithologic indicators of climate and the geochemical proxy data are integrated to
estimate the mean annual temperature (MAT), the warmest mean monthly temperature (WMMT), and the mean
annual precipitation (MAP). The MAT, MAP, and WMMT form the basis of the paleo-Köppen map shown in Figure
2C (Burgener et al., 2023).
The three methods used to produce paleoclimatic reconstructions: computer simulation (e.g., HadCM3, CESM),
lithologic indicators of climate (Boucot et al., 2013), and a Bayesian analysis of paleoclimatic proxies (Burgener et
al., 2023) give subtly different results. The best way to compare and contrast these different methods is to
compare the pole-to-equator gradient diagrams produced by these methods. Figure 24 illustrates the Pole-to-
Equator gradients for the Hauterivian-Barremian (A), the Albian (B), and the Turonian (C). The gray curve in each
diagram is the modern Pole-to-Equator gradient. In the present-day world, the average temperature near the
Equator is ~26˚C. The temperature remains nearly constant in the subtropics (0˚ - 15˚ N and S latitude) and then
begins to decrease rapidly. Freezing temperatures are reached at 60˚ latitude; falling below -35˚ C at the poles.
The modern pole-to-equator temperature curve lies well below the curves for the Cretaceous.
56
This is not a surprise. The modern global average temperature (GAT) is 14.5˚C, whereas estimates of GAT for the
Cretaceous range from 18˚C to 28˚C. The computer simulations (blue lines) are similar to each other, and the two
methods based on geological evidence are likewise similar. There are a few other general tendencies. The
computer simulations tend to produce warmer tropical temperatures. The exception is the Turonian, when all
models produced equally high tropical temperatures (~30˚C). Conversely, the mid to high latitude temperatures
observed in computer simulations are 10˚ - 20˚C cooler than the results from the proxy methods. The exception is
the Hauterivian-Barremian result. During this cooler period, the mid and high latitude temperatures obtained from
all methods are roughly equivalent.
Though we can be encouraged by the similar results obtained from all three disparate methodologies, interesting
anomalies still persist. The Bayesian method displays an unlikely flat gradient during the Albian with unsustainably
warm temperatures at moderate and high latitudes. At polar latitudes the computer simulations display a strange
tendency toward increased warming, unlike modern polar temperatures.
57
58
Figure 24. Pole-to-Equator temperature gradients during the Cretaceous (A - Hauterivian-Barremian, B - Albian,
and, C -Turonian). A comparison of pole-to-equator global surface temperature gradients from two proxy-based
models (Scotese et al., 2021 – dashed red line; Burgener et al., 2023 – solid red line) and two computer simulations
(Valdes et al., 2021 – solid blue line; Li et al., 2022 – dashed blue line). The modern average pole-to-equator
gradient (gray line) is shown for comparison (Legates and Willmott, 1990).
5. Modeling Cretaceous Paleorivers
1) Drainage system and paleorivers
Thanks to the paleo-DEMs (quantified topography; Scotese and Wright, 2018), it was possible to model drainage
systems and paleorivers during the Phanerozoic. The drainage systems were initially established without
considering precipitation, and thus provides metrics about how water would flow given sufficient rainfall. The
paleorivers that were filled with water were mapped once the amount of precipitation generated by HadCM3
climate model of Valdes et al. (2021) was applied. Other researchers have taken a similar approach using the
PALEOMAP paleoDEMS to model landscape dynamics on a global scale (Salles et al., 2023).
2) Method
We used the Hydrology tools of the ArcGIS® software to both produce the drainage maps and predict the location
of paleo-rivers for each stage of the Cretaceous. Emerged land areas were mapped using the Cretaceous Paleo-
DEMs. Land surfaces were then prepared using the Fill function of ArcGIS®, which removed spurious sinks or peaks,
and flow directions were computed and forced to flow down the topographic gradient. Note that all basins below
sea-level but not in contact with ocean basins were considered to be lakes and were filled up with the Fill function
of ArcGIS®. ArcGIS®. , The Basin function of ArcGIS® was used to delineate drainage basins by identifying the
drainage divide between basins. The corresponding raster files were then converted into polygons and the sizes of
the basins (geodetic surfaces) were computed. The Flow Length tool of ArcGIS® calculated the linear distance from
any given pixel along the drainage system to the river mouth. In other words, the flow length map shows how far a
drop of rainwater would need to travel to reach the shoreline, presumably a river delta.
The Cretaceous paleorivers (Figure 25) were mapped by calculating the flow accumulation (Flow Accumulation tool
of ArcGIS®) weighted from the precipitation maps of Valdes et al. (2021). Using HadCM3BL-M2.1DA, a variant of the
Hadley Centre HadCM3 model (http://www.ipcc-data.org/sim/gcm_clim/SRES_TAR/hadcm3_info.html),
simulations were run for every Cretaceous time slice (Table 1). Note that runoff was then merely taken into
account by bringing a quantity of water (precipitation) to the nearest ocean (as the crow flies) and not by following
the drainage system (the paleo-rivers) as defined here. For the sake of clarity, a conditional rule was set to only
select and display the paleorivers with the largest flux.
59
Figure 25. Early Campanian (80 Ma) drainage pattern and paleorivers. The thickness of the lines representing the
paleorivers is proportional to the river discharge.
3) Flow length
In each time slice, there are many small basins and only few large basins. The distribution is exponential. The same
is true regarding flow length, which represents the average river length. The mean values of the river length (and
associated standard deviation) were calculated on the basis of their logarithmic (log10) values, which
approximates a Gaussian distribution.
In general, the average river length during the Phanerozoic responds inversely to sea level change (Figure 26).
During the Cretaceous, the average river length decreases from a Cretaceous high at 145 Ma to an all-time low at
80 million years. Average river length then increases towards the end of the Cretaceous. As the dispersal or
aggregation of continents is not the major factor over this time interval, it is clear the length of rivers changed due
to rising and falling sea level.
60
Figure 26. Drainage Basin Size (gray line) and Mean Flow Length (red lines) during the Phanerozoic
61
C. Chronological Review of Plate Tectonics, Paleogeography, and Paleoclimate during the Cretaceous
1. Plate Tectonics during the Cretaceous
1) Overview
This section reviews plate tectonic events for four intervals during the Cretaceous: the earliest Cretaceous
(Berriasian – Barremian), the Early Cetaceous (Aptian – Albian), the mid- Cretaceous (Cenomanian-Turonian), and
the late Cretaceous (Coniacian – Maastrichtian). To better understand the complex series of Cretaceous plate
tectonic events, the reader is referred to figures 28 – 44 , as well as additional plate tectonic reconstructions in the
Supplemental Materials. While reading the text, the reader should simultaneously review the animation of
Cretaceous plate motions (“Plate Tectonics: 200 million years - Today” provided in the Supplemental Materials.
Scrolling backwards and forwards through the animation will give the reader a much better understanding of the
dynamic nature of Cretaceous plate tectonics.
Though the initial breakup of Pangea began in the early-middle Jurassic (200 Ma -170 Ma; Scotese and Schettino,
2017), most of the modern ocean basins formed by successive rifting events during the Cretaceous. North America
continued to separate from NW Africa during the Cretaceous widening the Central Atlantic. Simultaneously, East
Gondwana (Madagascar, India, Australia, and Antarctic) rifted away from West Gondwana (Africa and South
America) widening the Western Indian Ocean. In the South Atlantic, Africa and South America were fully separated
by the Albian (110 – 105 Ma). Shortly thereafter (~95 Ma, Cenomanian), India separated from Madagascar and
Australia slowly rifted away from East Antarctica. North America remained connected to Northern Europe during
the Cretaceous, though the Labrador Sea had partially opened, and multiple zones of extension had appeared
between East Greenland and Rockall Bank (northwestern Europe).
Other significant Cretaceous tectonic events include: the collisions of Stikinia and Wrangellia along the western
margin of North America (143 Ma and 100 Ma, respectively), the opening of the Circum-Arctic Ocean (145-125Ma),
the insertion of the Caribbean plate between North and South America (95-45 Ma), the rifting of Zealandia from
eastern Australia opening the Tasman Sea (85 Ma – 55 Ma), as well as the complex development of the Pacific,
Farallon, Izanagi, and Phoenix oceanic plates. Two major subduction zone systems dominated the Cretaceous
world: the Circum-Panthalassic subduction zone, and the northward dipping Tethyan subduction zone. This
complex series of Cretaceous rifting events is illustrated on the tectonic timeline. (Figure 27). Each major
Cretaceous rifting event is indicated by a split in the tectonic tree.
62
Figure 27. Plate Tectonic Tree Diagram for the Mesozoic and Cenozoic. Branches represent rifting events that form
new ocean basins (blue lettering). Coalescing root-like structure indicate continent-continent collisions (e.g. India
/Asia). Continents and continental terranes labelled with black italic letters, The abbreviations in capital letters
across the top of the diagram are the modern plates: ANT- Antarctica, AUS- Australia, SC – Scotia Sea, SAM – South
63
America, SOM – Somalia, AFR – Africa, ARB – Arabia, EUR – Europe, BAJ – Baja California, NAM- North America,
and CAR – Caribbean. Other abbreviations are: Nq – Neuquen Basin, DGMB – Dangerous Grounds and Macclesfield
Bank, and CoSa – Corsica and Sardinia.
64
Figure 28. Jurassic/Cretaceous Boundary (145 Ma) , see Figure 1A for legend.
65
Figure 29. Early Cretaceous (latest Berriasian, 140 Ma), see Figure 1A for legend.
66
Figure 30. Early Cretaceous (Valanginian, 135 Ma), see Figure 1A for legend.
67
Figure 31. Early Cretaceous (early Hauterivian, 130 Ma), see Figure 1A for legend.
68
Figure 32. Early Cretaceous (earliest Barremian, 125 Ma), see Figure 1A for legend.
69
2) Earliest Cretaceous (Berriasian – Barremian, 145 Ma – 125 Ma), Figures 28 – 32.
a. Intra-Pangean Ocean Basins
Seven of the eleven intra-Pangean ocean basins had opened by the start of the Cretaceous (145 Ma): the Central
Atlantic, the Gulf of Mexico, the Proto-Caribbean, the South Atlantic, the Western Indian Ocean, the Northeast
Indian Ocean (Argo Sea), and the Canada Basin. An intra-Pangean ocean basin is defined as an ocean basin formed
by the breakup of the encircling Pangean supercontinent.
The Central Atlantic, which had opened in the early-middle Jurassic (~175 Ma; Scotese and Schettino (2017), was
approximately 1500 km wide by the early Aptian (125 Ma). The Bahamas Platform hot spot track divided the
Central Atlantic from the Proto-Caribbean and Gulf of Mexico. Rifting in the Gulf of Mexico, which had opened
synchronously with the Central Atlantic, was finished by the earliest Berriasian (144 Ma; Ross and Scotese, 1988;
Pindell et al., 1988; Pindell and Barrett, 1990; Pindell et al., 2005; Pindell and Kennan, 2009 ). Most of the Gulf of
Mexico is underlain by hyper-attenuated continental crust, however a small crescent of oceanic crust lies at the
center of the basin (Buffler and Sawyer, 1983).
The Proto-Caribbean ocean basin was divided into two parts by the Yucatan peninsula. The triangular-shaped
eastern Proto-Caribbean ocean basin lay between the Bahamas Platform, the northern coast of Venezuela and the
eastern shores of Yucatan. The western half of the Proto-Caribbean ocean basin was located between western
Colombia and eastern Honduras. Both halves of the Proto-Caribbean ocean basin opened as North America, pulled
away from South America, together with Africa, in the late Jurassic and early Cretaceous.
Portions of the volcanic islands of Cuba and Hispaniola (Greater Antilles) formed the southwestern margin of the
Proto-Caribbean ocean basin. The Greater Antilles were originally a continental island arc that was continuous with
the Andean margins of Mexico and South America.
By the Berriasian (140 Ma), the South Atlantic had just begun to open. The eruption of the Párana-Entendeka flood
basalts (130 – 138 Ma) occurred during the early rift phase. Oceanic crust (early drift phase) had just begun to form
in the portion of the South Atlantic south of the Rio Grande Rise - Walvis Ridge hot spot track. The central portion
of the South Atlantic from the São Paolo Plateau to the Benue Trough, was characterized by rift lakes and crustal
extension similar to that in the present-day East African Rift System. The northern segment of the South Atlantic
margin, from the Benue Trough to Guinea Plateau/Demerara Rise had yet to experience significant extension. As
Africa separated from South America, some of the deformation was taken up by right-lateral transtensional
deformation in southern South America that created rift basins in the vicinity of the Rio de le Plata and south of
the Rio Negro.
70
The Somali Basin, Mozambique Basin, and Weddell Sea were actively opening during the earliest Cretaceous. East
Gondwana (Madagascar, India, Antarctic, and Australia) drifted southward, away from West Gondwana (Africa and
South America). Madagascar slid past northeastern Mozambique along the Davie Fracture Zone, while Queen
Maud Land slid past the eastern tip of the Falkland Plateau along the Mozambique Escarpment. This strike-slip
separation of East and West Gondwana created three narrow ocean basins: the Somali Basin in the north, the
Mozambique Basin in the center, and the Weddell Sea in the south.
Though most of East Gondwana (Madagascar, India-Greater India, Australia, and Antarctica) remained intact
during the earliest Cretaceous, by 125 Ma (latest Barremian), a small ocean basin had opened between India and
East Antarctica (Queen Maud Land) and where the eastern face of Greater India abutted against western Australia.
Plate stresses in this region appear to have been erratic during this time period and the location of the mid-ocean
ridges in the Perth Basin and Gascoyne Plain jumped to the northwest on several occasions (Powell et al., 1988).
b. Extra-Pangean Ocean Basins
The splintering remnants of Pangea were surrounded by five distinct ocean basins: the Pacific Ocean, the Tethys
Sea, the Amurian Seaway, the Angayucham Ocean, and the Wrangellian back-arc basin, also called the Cache Creek
Sea. The largest of these ocean basins was the Pacific Ocean, which was a direct descendent of the Paleozoic and
early Mesozoic, Panthalassic Ocean. Four oceanic plates: 1) Pacific Plate, 2) Farallon Plate, 3) Phoenix plate, and 4)
the Izanagi Plate continued to grow during the earliest Cretaceous (Wright et al., 2016). These four plates had
emerged from the expansion of a triple-triple junction that formed in the southwest Pacific during the early part of
the Jurassic (~190 Ma), most likely due to a massive mantle plume that also generated the thick basaltic crust
underlying the Caribbean plate.
The recumbent, v-shaped Tethys Ocean separated the northern continents of Laurasia (North America and Eurasia)
from the southern hemisphere continents that were located along the Afro-Indian-Australian margin of
Gondwana. Two mid-ocean ridges (North Tethys Ridge and South Tethys Ridge) divided the Tethys Sea into three
oceanic plates: a small Paleotethys plate to the north, a large Mesotethys oceanic plate in the middle, and the
nascent Neotethys plate to the south.
The North Tethyan Ridge formed along the northern margin of Indo-Australian Gondwana during the Early Jurassic.
Beginning in the Late Jurassic and continuing into the earliest Cretaceous, a mysterious continental fragment
(Argoland; Audley-Charles, 1988; Powell et al., 1988) rifted away from the northwestern shelf of Australia (Argo
Sea). We do not know precisely where this continental fragment ended up. Some authors believe that it is now
part of Southeast Asia (Metcalfe, 1992, 1993, 1999; Zahirovic et al., 2014; Vérard et al., 2017 Advokaat and
71
Hinsbergen, 2023). By the Berriasian, the North Tethyan Ridge had migrated northward to a position off the
southern coast of central Asia. During the Early Cretaceous the North Tethyan Ridge was progressively subducted
beneath Eurasia and the last remnants of the Paleotethys plate was subducted by the early Aptian (125 Ma).
The South Tethyan Ridge was initially a slow spreading ridge that formed along the northern margin of the Indo-
Australian margin of Australia and directly connected to the mid-ocean ridges in the Somali and Mozambique
basins. The South Tethyan Ridge separated northwest India and Greater India from a thin sliver of continental crust
that would later become the Lut block of eastern Iran. During Early Cretaceous, the South Tethyan Ridge
coincidentally aligned with the Pacific-Izanagi mid-ocean ridge, though the precise nature of plate boundaries in
the easternmost Tethys is speculative.
Active subduction zones bordered the expanding frontiers of Pangea during the earliest Cretaceous. In several
areas, back-arc basins opened along margins that were previously Andean-type subduction zones (e.g., the Qiang
Tang back-arc basin along the northern margin of Tethys). In other places, back-arc basins that had opened in the
Triassic and Jurassic collapsed and were converted into compressional, foredeep environments (Neuquen Basin,
western Argentina; Uliana and Leggarreta, 1993;Cobbold and Rossello, 2003). Along the North american margins of
the Northeast Pacific Ocean were two notable back-arc basins: the Angayucham Sea and the Cache Creek Sea. The
Angayucham Sea (Nokleberg et al., 2001) began to close in the earliest Cretaceous, as the North Slope terrane
rotated counter-clockwise away from the Arctic Islands of Canada, opening the Canada Basin. The Wrangellian
back-arc basin, which had opened in the mid-Triassic (~230 Ma), closed in the mid-Cetaceous (~110 Ma; Nokleberg
et al., 2001).
The Cretaceous was nearly devoid of continent-continent collisions. The one exception was the closure of the
Amurian Seaway in the late Jurassic and earliest Cretaceous. The Amurian Seaway diachronously closed (west-to-
east), as a collage of recently assembled continental terranes (Amuria, North China, Tarim, South China, Indochina,
and Sibumasu) collided with Mongolia and southeastern Siberia along the Mongol-Okhotsk suture (Zonenshain et
al., 1990). The Amurian Seaway was completely closed by the Barremian (125 Ma).
72
Figure 33. Early Cretaceous (early Aptian, 120 Ma), see Figure 1A for legend.
73
Figure 34. Early Cretaceous (late Aptian, 115 Ma), see Figure 1A for legend.
74
Figure 35. Early Cretaceous (early Albian, 110 Ma), see Figure 1A for legend.
75
Figure 36. Early Cretaceous (middle Albian, 105 Ma), see Figure 1A for legend.
76
3) Early Cretaceous (Aptian - Albian, 125 Ma – 100 Ma), Figures 32 – 37.
a) Northwest Sector of Pangea
By mid-Early Cretaceous (125 Ma), Pangea had divided into three major continental groupings: 1) Laurasia, (North
America, Northern Europe, and Asia) to the northwest, 2) West Gondwana, (Africa, Iberia, Southern Europe,
Arabia, and South America) in the center, and 3) East Gondwana (Madagascar, India, Greater India, Antarctica,
Australia, and Zealandia) to the southeast. During the Aptian-Albian, each of these continental groupings began to
subdivide.
In the northwest sector, Laurasia slowly began to break apart. Iberia, which was attached to northwest Africa
(Morocco) rifted away from western France slowly opening the Bay of Biscay (125 Ma - 90 Ma). Rifting propagated
northward into the North Atlantic, stretching the continental crust between Greenland, Rockall Bank, and Western
Europe.
The Proto-Caribbean ocean basin continued to open as South America, together with Africa, pulled away from
North America. The Proto-Caribbean ocean basin would continue to widen until the mid-Albian (105 Ma). By that
time, South America was completely separate from northwest Africa and had begun to move in concert with North
America. The Proto-Caribbean Ocean stopped widening because relative motion between the North American and
South American plates had essentially ceased.
Cuba and Hispaniola (Greater Antilles) formed the southwestern margin of the Proto-Caribbean ocean basin during
the Aptian-Albian. The Greater Antilles were originally a continental island arc that was continuous with the
Andean margins of Mexico and northern South America.
To the far north, during the latest Jurassic and Early Cretaceous (155 Ma – 125 Ma), the Canada Basin opened as
the North Slope of Alaska rifted away from the Arctic margin of Canada (Lawver and Scotese, 1990; Nokleberg et
al., 2001; Scotese, 2008). The opening of the Canada Basin stopped when the North Slope/Chukotka block collided
with the continental backstop made up of central Alaska and northeast Siberia (Lawver and Scotese, 1990). This
collision of the North Slope block along the Brooks Range closed the Angayucham Ocean (Barremian, 125 Ma).
b. Central Sector of Pangea
In the central sector, Africa and South America did not break apart as two rigid continental blocks. Rather, there
was a period of diachronous separation (150 Ma – 110 Ma) as the rift zone between Africa and South America
propagated northward, facilitated by crustal weaknesses caused by the Tristan da Cunha (Entendeka and Párana
77
LIP, 130-135 Ma)and the Bahamas hot spots. The oldest datable magnetic anomalies in the far South Atlantic are
M11 (133 Ma), whereas the oldest magnetic anomaly in the central South Atlantic is M3, ~125Ma.
This diachronous pattern of rifting resulted in considerable trans-tensional deformation in southeastern South
America (150 Ma – 125 Ma) and later in Central Africa (120 Ma – 110 Ma). The extension in southeastern South
America took place primarily in the Salado and Rio Negro basins. Hypothetical right-lateral shear zones separated
the more rigid part of the South American plate (Amazonia) from the more deformable Paranaiba, Rio de la Plata,
and Patagonian blocks (Figure 4). During the Aptian and Albian (120 Ma – 110 Ma), right-lateral trans-tension along
the Central African shear zone produced continental rifting in Niger and southern Sudan (Fairhead, 1988;
Unternehr et al., 1988).
The likely explanation for this unzipping of the South Atlantic can be found along the northern boundary of the
African plate. Starting in the latest Jurassic (~150 Ma), the southern boundary of the Paleotethyan plate, the North
Tethyan Ridge, began to be subducted beneath southern Eurasia. The subduction of the North Tethyan Ridge was
diachronous, starting in the west beneath Iran, and progressing eastwards towards Sibumasu. As a consequence,
the African plate, for the first time, began to be subducted northward beneath Eurasia. Gradually, as more and
more of the African plate was subducted, increasing slab-pull forces pulled Africa, together with India away from
South America.
The widening South Atlantic separated the Santos and Campos basins of eastern Brazil from the Gabon and
Kwanza basins of West Africa. What were once rift lakes became narrow, deep marine basins floored by
attenuated continental crust (São Paolo plateau) and filled with salty brines that would become thick, Aptian salt
deposits. These salt basins, which formed in the arid subtropics, were isolated from the southernmost South
Atlantic by a massive submarine, volcanic escarpment formed by the Rio Grande Rise – Walvis Ridge hot spot track.
c. Southeast Sector of Pangea
During the Early Cretaceous, in the southeast sector East Gondwana rifted apart as the western portion, India
together with Madagascar, separated from the combined Antarctica/Australia continent. Starting about 140 Ma,
the northeast coast of India (Eastern Ghats) diachronously (east to west) rifted away from the Enderby Land
margin of East Antarctica. At about 112 Ma, final separation took place as Sri Lanka tore away from the socket
formed by the Gunnerus Ridge. The Cauvery Basin (Straits of Mannar), which now separates Sri Lanka from
mainland India, was produced during this final phase of rifting. The eruption of the Rajmahal Flood Basalts in
Assam (~120 Ma) is associated with the rifting of India from Antarctica. The Rajmahal flood basalts mark the
initiation of the Kerguelen hotspot, which would generate the Kerguelen plateau and Ninety East Ridge during the
Late Cretaceous and Cenozoic.
78
During the mid-Cretaceous (~110 Ma), sea floor spreading ceased in the Somali Basin. As a consequence,
Madagascar became fixed to the African plate (Segoufin and Patriat, 1981; Coffin and Rabinowitz, 1988). Seafloor
spreading in the Mozambique Basin, however, continued and the Southwest Indian Ocean and the Weddell Sea
widened. The east-west trending mid-ocean ridge in the Mozambique Basin slid past the eastern edge of the
Mozambique Escarpment. When the ridge cleared the southern end of the escarpment , it jumped northwards
forming the series of stair-step offsets characteristic of the modern Southwest Indian Ridge.
Though Australia and Antarctica remained together throughout the Early Cretaceous, the Wharton Basin between
western Australia and Greater India continued to widen. During the Early Cretaceous, the Southcentral Indian
Ocean became wider; however, the Kerguelen hot spot landbridge may have linked India with East Antarctica.
At ~110 Ma (early Albian), sea floor spreading in the Somali Basin and along the South Tethyan mid-ocean ridge
simultaneously ceased. This resulted in the fusing of the Neotethys and Mesotethys plates. As a consequence, the
Tethyan subduction zone continued to draw Africa northwards and for the first time the Tethyan subduction zone
began to pull India northwards towards Asia.
d. Ocean Basins Exterior to Pangea
During the Aptian-Albian, the ocean basins that were exterior to Pangea (Neotethys and the Pacific Ocean)
continued to contract as the interior ocean basins expanded (Atlantic and Indian Oceans). The Pacific plate
continued to expand at the expense of the Farallon, Izanagi, and Phoenix plates. During the Aptian (~120 Ma), a
“superplume” erupted in the Pacific Ocean Basin, covering the once relatively smooth seafloor with > 100 intra-
plate volcanic islands and seamounts. Most notable was the eruption of several large, submarine volcanic plateaus
(Ontong Java Plateau, Manihiki Plateau, 123 Ma). The Ontong Java Plateau consists of more than 50 million km3 of
basaltic magma forming a 30 km thick plateau equal in area to one-third of the United States (Tarduno et al.,
1991). It has been proposed that tremendous outgassing of CO2 associated with this volcanic event led to global
warming and resulted in the early Aptian oceanic anoxic event (OAE1a) (Larson, 1991; Larson and Erba,1999).
In the northwest Pacific, exotic terranes (Omolon and Kolyma) collided along the northeastern margin of Siberia
resulting in folding and thrusting in the Verkhoyansk mountain range (~120 Ma; Parfenov et al., 2005, 2011). These
collisions were followed by the initiation (~100 Ma) of a new, west-dipping subduction zone (Okhotsk-Chukotka
volcanic arc) (Nokleberg et al., 2001).
Eastward along the Pacific margin of North America, the Wrangellian Seaway had closed by the late Albian (~100
Ma). This collision is marked by the Gravina-Nutzotin orogenic collage (Friedman, 1983; Monger, 2008).
79
Paleomagnetic data indicates that during the Cretaceous, Wrangellia moved northward approximately 2000 km
outboard of the western margin of North America. It collided with North America (British Columbia/Yukon) 850 km
south of its present-day location. During the Paleocene and Eocene, it was displaced northward along continuous,
right-lateral strike slip faults (Tintina-Rocky Mountain Trench faults).
To the south, off the coast of Central America, the easternmost portion of the Farallon plate, closely approached
the eastward-dipping Greater Antilles volcanic arc. The collision of this thickened oceanic crust (~100 Ma), flipped
the polarity of the Greater Antilles subduction zone, resulting in the insertion of the Caribbean plate between
North and South America.
80
Figure 37. Early Cretaceous (latest Albian, 100 Ma), see Figure 1A for legend.
81
Figure 38. Mid-Cretaceous (Cenomanian, 95 Ma), see Figure 1A for legend.
82
Figure 39. Mid-Cretaceous (latest Turonian , 90 Ma), see Figure 1A for legend.
83
4) Mid-Cretaceous (Cenomanian-Turonian, 100 Ma – 90 Ma), Figures 37 -39
Though only ten million years in duration (100 Ma – 90 Ma), the Cenomanian-Turonian was a tectonically active
time. Five of the eight intra-Pangean ocean basins had opened by the beginning of the Cenomanian-Turonian:
Central Atlantic, Proto-Caribbean, South Atlantic, Western Indian Ocean, and Southcentral Indian Ocean. The
Central Atlantic, the oldest of these ocean basins, was now more than 4500 km wide. The Proto-Caribbean ocean
basin, located to the southwest of the Central Atlantic, was at its maximum extent during the early Cenomanian.
However, sometime during the Cenomanian and Turonian, a new subduction zone, dipping to the west beneath
the Greater Antilles island arc (Cuba, Hispaniola, and Puerto Rico) began to consume and override Proto-Caribbean
ocean-floor. This reversal of subduction direction was triggered by the arrival of the thick, basaltic, oceanic
lithosphere that was to become the Caribbean plate.
The South Atlantic, which had progressively opened from north to south during the early Cretaceous, had
completely broken through to the Central Atlantic by the Cenomanian-Turonian, allowing a free-flowing deep-
water connection between the South Atlantic and Central Atlantic. Similarly, relatively wide oceanic gateways
linked the South Atlantic, Weddell Sea, Southwestern Indian Ocean, Western Indian Ocean and Southcentral Indian
Ocean.
The three remaining intra-Pangean ocean basins began to rift open during the Cenomanian-Turonian: Madagascar
Basin, Southeast Indian Ocean, and North Atlantic (Labrador Sea; Srivastava and Roest, 1989). India, with
Madagascar attached, remained motionless with respect to Africa until the late Cenomanian (~100 Ma; Coffin and
Rabinowitz, 1988), when volcanic activity along the eastern coast of Madagascar signaled the start of India’s
northward journey towards Asia. It is notable how straight the east margin of Madagascar and the west margin of
Madagascar are. This 1200 km long boundary was undoubtedly the site of a major N-S strike-slip fault prior to the
rifting of India from Madagascar. The timing of the strike-slip movement, the magnitude of the fault, and the sense
of offset (left-lateral vs. right-lateral), are not well resolved. Our model shows a minimum amount of strike-slip
movement between India and Madagascar.
During the Cenomanian-Turonian (95 Ma) the first phases of rifting began between Australia and Antarctica (Cande
and Mutter, 1992). Though no oceanic crust was generated, extensive crustal stretching resulted in the formation
of a series of rift lakes. Further to the west, this slow, inexorable extension produced vast volcanic plateaus
(Broken Ridge and Naturaliste Plateau) associated with the Kerguelen Plateau and hot spot. This phase of rifting
was synchronous with the initiation of the separation of Zealandia (Luyendyk, 1995; Matthews et al., 2015) from
eastern Australia and the opening of the Tasman Sea.
84
The North Atlantic Ocean basin has had a long history of pre-rift extension and stretching. In the Cenomanian-
Turonian two rift systems on either side of Greenland were in operation. Sea floor spreading was just beginning in
the Labrador Sea to the west of Greenland in the late Cenomanian (~95 Ma). The West Greenland rift system
passed up through the Labrador Sea and into Baffin Bay and was bordered on either side by high rift shoulders that
are still evident today. The Iceland hotspot passed down the length of the West Greenland rift system during the
Late Cretaceous and, at 90 million years, was located along the west coast of Baffin Island near Disko Island (70°
N). On the other side of Greenland, the East Greenland rift system may have been active as early as the late
Permian; seafloor spreading, however, did not begin in earnest until the late Paleocene. During the Cenomanian-
Turonian a broad zone of stretched continental and transitional lithosphere extended from the western margin of
Ireland, across Rockall Plateau and Hatton Bank, to the shores of southeast Greenland.
The Pacific Ocean, Tethys Sea, and Arctic Ocean surrounded the perimeter of the Cenomanian-Turonian
continents. Along the eastern margin of the Pacific Ocean, the leading edge of the Farallon plate continued to
override the older oceanic lithosphere of the proto-Caribbean plate. By the end of the Turonian (~90 Ma), a new
eastward dipping subduction zone had formed in the equatorial, eastern Pacific that clipped the Caribbean plate
from the Farallon plate. This proto-Mid America Trench linked the eastward dipping subduction zones of Mexico,
to the north, and Colombia - Peru, to the south.
During the Cenomanian-Turonian, the Pacific plate continued to grow while the Izanagi, Farallon, and Phoenix
plates continued to be subducted beneath an expanding Cretaceous Ring of Fire. In the far reaches of the South
Pacific, during the late Albian (100 Ma), the Pacific-Phoenix mid-ocean ridge jumped to a new location closer to
Antarctica. This was followed by a second ridge jump ten million years later caused by the oceanward retreat of
the subduction zone dipping westward beneath eastern Australia (van de Lagemaat et al., 2023). This rollback of
the subduction zone resulted in the opening of the Tasman Sea and the isolation of the mostly submarine
subcontinent, Zealandia, from the rest of Gondwana (Gaina et al., 1998 a,b).
Sometime during the late Albian or early Cenomanian (~100 Ma), a pair of back-arc basins opened along the
northwest rim of the Pacific Ocean: the Okhotsk- Bering back-arc basin (Parfenov et al., 2010, 2011; Vaes et al.,
2019) and the Philippine back-arc basin. The Okhotsk - Bering back-arc basin was bounded on the east by the
Kamchatka/Olyutorsky – Peninsular Alaska volcanic arc and on the west by Northeast Siberia (Kolyma, Omolon,
and Chukotka). We propose that the Philippine volcanic arc was initially located along the eastern margin of Asia
between southern Japan and Taiwan. Both of these back-arc basins opened due to slab rollback, possibly triggered
by the subduction of the oldest and densest portions of the Izanagi plate.
85
Though active subduction zones consumed oceanic lithosphere along the western margins of the Americas,
beneath the southern margin of Eurasia, beneath western Antarctica and New Zealand, and beneath the eastern
margin of Asia, there were no continent-continent collisions during the Cenomanian-Turonian. There were,
however, three episodes when island arcs collided and ophiolites were obducted onto continental margins (Oman,
Troodos, Chortis, and northern California; Metzler, 2006).
86
Figure 40. Late Cretaceous (Santonian - Coniacian, 85 Ma), see Figure 1A for legend.
87
Figure 41. Late Cretaceous (Early Campanian, 80 Ma), see Figure 1A for legend.
88
Figure 42. Late Cretaceous (Late Campanian, 75 Ma) , see Figure 1A for legend.
89
Figure 43. Late Cretaceous (Maastrichtian, 70 Ma), see Figure 1A for legend.
90
Figure 44. Cretaceous/Paleogene Boundary (65 Ma), see Figure 1A for legend.
91
5) Late Cretaceous (Coniacian - Maastrichtian, 85 Ma – 65 Ma), Figures 40 – 44.
a. Overview
By the Late Cretaceous, all of the major ocean basins had fully opened, with the exception of the North Atlantic
which was still in its initial stage of rifting. Like the modern world, the dominant subduction zone was the Circum-
Pacific Ring of Fire which produced Andean-style mountains along the eastern rim of the Pacific (North and South
American Cordillera) and back-arc basins along the northwestern (Sea of Okhotsk, Bering Sea, and Philippine Sea)
and southwestern margin of the Pacific (Tasman Sea). Subduction of the Neotethys plate continued beneath the
southern margin of Eurasia as India was drawn rapidly northward by slab-pull forces. Throughout most of the Late
Cretaceous there were no continent-continent collisions and there were no new continental rifts. The single
exception is the Alpine collision between the Adria, a promontory of northern Africa, and south-central Europe
(AustroAlpine, Carnic – South Karawanken, and Transdanubian regions), which began at the very end of the
Cretaceous.
b. Circum-Arctic and North American Cordillera
During the Late Cretaceous, the Arctic region was quiescent (Green et al., 1986). The Amerasian Basin (Canada
Basin and Makarov Basin) had opened by the mid-Cretaceous (~110 Ma; Lawver and Scotese, 1990; Nokleberg et
al., 2001) and the Eurasian Basin had yet to open (60 Ma; Srivastava, 1985; Rowley and Lottes, 1988). All the Late
Cretaceous tectonic activity in the Circum-Arctic region took place in the Canadian Cordillera and in South-Central
Alaska (Monger and Nokleberg, 1996). The Wrangellia terrane, which had collided in the mid-Cretaceous (~110
Ma) with the Stikine terrane, late in the Cretaceous (<90 Ma) began to translate northward with respect to
cratonic North America along large (~700 km) right-lateral, strike-slip faults (Rocky Mountain Trench and Tintina
Fault (Monger, 2008; Monger and Nokleberg, 1996; Engebretson et al., 1985). This northward movement would
continue until the middle Eocene.
c. North Atlantic and Europe
In the North Atlantic, the final phases of continental rifting were taking place between Greenland and
Northern Europe (Ziegler, 1988; Skogseid et al., 2000; Srivastava and Tapscott, 1986; Rowley and Lottes, 1988).
Late Cretaceous extension in the North Atlantic produced the Rockall and Hatton Bank microcontinents and the
wide, stretched margins of East Greenland and Central Norway (Vøring Plateau, Skogseid, et al., 2000). The
Labrador Sea and the southernmost part of the North Atlantic (the Rockall Trough and the Rockall-Hatton basin;
Srivastava and Roest, 1989) were the only sites of seafloor spreading in the North Atlantic during the latest
Cretaceous.
92
It is interesting to note that in the latest Cretaceous, rifting was taking place simultaneously on either side of
Greenland. This is the only known case of two dueling rifts in Phanerozoic history. The location of the western rift,
between Baffin Island and western Greenland, was probably due to the weakening of the continental lithosphere
above the southward moving Iceland hotspot (Scotese, 2008). During the latest Cretaceous and Paleocene, the
Iceland hot spot was located at the eastern edge of the Labrador Sea (Disko Island). Greenland’s westward drift
caused the not spot to move below southern Greenland emerging into the North Atlantic during the early Eocene
(Lawver and Müller, 1994). The arrival of the Iceland mantle plume, combined with highly attenuated continental
lithosphere, would trigger massive volcanic eruptions in the North Atlantic region in the Paleocene.
In the Late Cretaceous, in Southern Europe, the Adriatic Promontory (the northernmost extension of the African
continent) had partially closed the Penninic Sea, and the Alps had begun to rise from the sea (Ziegler, 1982, 1990;
Dercourt et al., 1985, 1993, 2000; Yilmaz et al., 1996; Ziegler and Horvath, 1996; Ziegler et al., 2001; Stampfli et al.,
2001; Stampfli et al., 2004; Stampfli and Kozur, 2006; Schettino and Turco, 2014; Scotese and Schettino, 2017; van
Hinsbergen et al., 2020). Along the southern margin of Eurasia, due to the subduction of the Neotethyan plate, an
on-echelon series of back-arc basins (Black Sea, southern Caspian Sea, Ladakh-Gangesi back-arc basin) opened
sequentially from west to east during the Late Cretaceous. The Ladahk-Gangesi back-arc, which produced
ophiolites of late Cretaceous age (Metzler, 2006), was the last back-arc basin to open along northern margin of
Tethys.
d. Central Atlantic Ocean and Caribbean
During the latest Cretaceous (70 million years ago), the Central Atlantic was approximately 60% as wide as
it is today. The Cruiser, Great Meteor, and Corner seamounts had just erupted at a hotpot located along the Mid-
Atlantic Ridge. The Great Meteor hot spot had produced the New England seamount chain and White Mountain
volcanics earlier in the Cretaceous.
The Late Cretaceous was a time of considerable plate tectonic activity in the Caribbean (Ross and Scotese, 1988;
Pindell and Barrett, 1990; Pindell et al., 2005). As noted earlier, late in the Cretaceous, an extension of the Farallon
plate wedged itself between North and South America. The Greater Antilles island arc moved to the northeast and
continued to override the proto-Caribbean plate. Approximately 90 million years ago, this crustal extension was
severed from the Farallon plate, giving rise to an autonomous Caribbean plate (Ross and Scotese, 1988). The
western edge of Cuba slid past the eastern margin of Yucatan and the newly-formed Panama volcanic arc rose
from the ocean.
Several tectonic events are associated with the latest Cretaceous/early Paleogene collision of Cuba with the
Bahamas Platform: 1) the opening of the Yucatan Basin between southern Cuba and the Cayman Ridge; 2) the start
93
of rapid motion along the Motagua and Polochic faults as the Chortis block slid eastward along a series of sinistral
strike-slip faults; and 3) the creation of the Aves Ridge, a proto-Lesser Antilles island arc that developed on the
leading edge of the Caribbean plate.
e. South Atlantic Ocean and Southwest Indian Ocean
At the end of the Late Cretaceous (66 Ma), the South Atlantic was approximately half its modern width (Cande et
al., 1988; Fairhead, 1988). In the central South Atlantic, the southwestern portions of the Walvis Ridge were being
generated by the Tristan da Cunha hot spot (Coffin and Eldholm, 1994).
Between 72 Ma (C32) and 64 Ma (C28), the motion of the African plate shifted dramatically (Royer et al., 1988).
Prior to Chron 32, the Southwest Indian Ridge was spreading in a NE-SW direction. Starting at 72 Ma, the spreading
rate slowed and the spreading direction became more N-S (Royer et al., 1988). In the earliest Cenozoic (64 Ma),
the spreading orientation reverted to a more NE-SW direction. These changes in spreading direction corresponded
with a major ridge-jump in the South Atlantic. In the latest Cretaceous - earliest Paleocene, the Bouvet Triple
Junction, which had been located ~1000 km to the southwest of Africa in the Agulhas Basin, jumped to a new
location adjacent to the Falkland Plateau. Volcanic leakage from this ridge-jump generated the Meteor Rise
(African plate) and the Islas Orcadas Rise (South American plate).
It is notable that this wobble in the movement of the African plate can be seen along many of its margins. The
equatorial fracture zones that linked Africa and South America underwent a compressive distortion that resulted in
the uplift and inversion of sedimentary basins adjacent to the transform faults (Jubilee oil field). Oceanic crust was
obducted along the Arabian margin of northeast Africa (Oman) as back-arc basins began to close. One could
speculate that this adjustment in Africa's plate motion may have been caused by the resistance Africa encountered
as a result of the continent-continent collision between northern Africa and southern Europe during the latest
Cretaceous (Alpine Orogeny).
f. Western Indian Ocean, Madagascar, and India
As noted earlier, about 95 million years ago rifting began between India and Madagascar as India began its
northward movement towards Asia. India pulled away from Madagascar, opening the Madagascar Basin (Segoufin
and Patriat, 1981; Royer et al., 1992; Royer and Sandwell, 1989). The Seychelles, Laxmi Ridge, Mascarene Bank,
and Nazareth Bank represent small blocks of continental crust that were torn away from India during the initial
rifting event. These continental blocks remained attached to the Indian plate as the Madagascar Basin opened.
They became separated from India in the early Paleocene as a result of a ridge jump from the Madagascar Basin to
the Arabian Sea (~61 Ma).
94
During the Late Cretaceous, India moved northwards and NeoTethys narrowed (Patriat and Segoufin, 1988; Royer
and Sandwell, 1989; Royer et al., 1992). Chatterjee and Scotese (1999) suggest that in the latest Cretaceous, India
began to collide with island arcs that fringed the southern margin of Asia. The classical Himalayan ophiolites
(Metzler, 2006) represent the ocean floor that had formed in back-arc basins behind these island arcs. An alternate
hypothesis (Burg, 2011; Vérard et al., 2017) supposes that in the latest Cretaceous, India collided with an
archipelago that was the eastward extension of an island arc that collided with Arabia (Oman ophiolite) in the Late
Cretaceous (80 - 70 Ma, Campanian). The ophiolitic remnants found on the island of Masirah, off the coast of
Oman, may be a obducted portion of this island arc. In the Maastrichtian, the northward-moving Indian
subcontinent collided with eastward extension of this island arc. This collision created a temporary landbridge
between Africa and India. This fleeting connection may explain the anomalous appearance of African dinosaurs in
India during the Maastrichtian (Chatterjee and Scotese, 1999).
An alternate hypothesis proposes that in the late Maastrichtian the leading edge of India collided with the Lut
block (eastern Iran). In the reconstruction of Gondwana presented here, the Lut block lies nestled between
southernmost Arabia and northwest India. When India and Madagascar broke away from East Africa in the
Jurassic, the Lut block remained attached to Arabia. Later in the mid-Cretaceous, when a realignment of plate
forces caused India to move northward towards Asia, the Lut block similarly broke free.
It has long been noted that India set plate tectonic speed records during the Late Cretaceous and earliest Cenozoic
(Molnar and Tapponier, 1975; Kumar et al., 2007). This rapid northward movement (>20 cm/yr) was due to the fact
that India was attached to a large area of subducting oceanic lithosphere that produced powerful slab-pull forces
The eruption of the Deccan Large Igneous Province (Ernst, 2014) at the Cretaceous/Paleogene boundary may have
contributed to India’s rapid movement by warming the base of the Indian plate and thereby reducing mantle drag
(Cande and Stegman, 2011).
To the south of Madagascar, a broad volcanic, oceanic plateau (Madagascar Plateau) was generated by the Marion
Dufresne hot spot (Duncan and Storey, 1992; Coffin and Eldholm, 1994). Just prior to the end of the Cretaceous,
this hot spot crossed the Southwest Indian Ridge, and began to generate the Conrad Rise. A little further to the
east, starting at 90 Ma, the Kerguelen hotspot was creating an island chain that would become the Ninety East
Ridge.
g. Southeast Indian Ocean, Australia, and Zealandia
The Late Cretaceous was a time of significant tectonic activity in the Southeast Indian Ocean and in Australia.
During the mid-Cretaceous, Australia was separated from the rapidly, moving Indian plate by the long, north-
south, dextral transform fault system (Larson, 1975; Larson et al., 1979; Powell et al., 1988; Royer and Sandwell,
95
1989) that ran parallel to the Ninety East Ridge (Duncan and Storey, 1992). In the Late Cretaceous, Australia had
separated from Antarctica (Cande and Mutter, 1982; Wilkes Land), and very slow seafloorspreading (<2cm/yr)
between Australia and Antarctica continued through the Maastrichtian and into the Paleogene. Australia and India
joined together to form the Indo-Australian plate (42 Ma) after the last major segment of the Neotethyan Ridge
was subducted beneath Asia. The ocean floor attached to northern Australia was, for the first time, subducted into
the Tethyan trench pulling Australia northward.
During the Late Cretaceous the Tasman Sea back-arc basin continued to open progressively from north to south
along the eastern margin of Australia (Gaina et al., 1998a,b). In the latest Cretaceous the northern segment of this
rift system between the Queensland plateau and the New Caledonia began to open. The Tasman Sea stopped
opening in the Paleocene (~55 Ma).
It should be noted that initial seafloor spreading in the Tasman Sea (~90 Ma) was the westernmost extension of
the Southwest Pacific mid-ocean ridge. When Australia began to move rapidly northward in tandem with India,
spreading in the Tasman Sea ceased and the South West Pacific ridge (southeast of Australia) and South East
Indian Ridge (south of Australia) joined together (~50 Ma). This is when the Alpine Fault became active and split
proto-New Zealand in two. The northwestern part of Zealandia, comprised North Island and the western half of
South Island, remained fixed to Australia, while the southeastern part of Zealandia, made up of much of the South
Island of New Zealand, Lord Howe Rise, and the Campbell Plateau, began to move with the Pacific plate.
h. Western Pacific and Northeast Asia
Only a fragmentary record exists of the plate tectonic history of the Western Pacific and Northeast Asia (Sharman
and Risch, 1988, Parfenov et al., 2010, 2011; Vaes et al., 2019). The Northeast margin of Asia was an active margin
with subduction beneath Northeast Siberia, the Sea of Okhotsk, Japan, the Yellow Sea, the South China Sea,
eastern Borneo, and the Philippines. A speculative model is proposed here regarding the tectonic evolution of the
Philippine and the Olyutorsky volcanic arcs. Both arcs began as continental volcanic arcs along the northeastern
margin of Eurasia. The Olyutorsky arc was originally located between the Koryak peninsula and Sakhalin Island. The
Philippine island arc was originally located between southern Japan and the modern location of Taiwan. It is
proposed that during the Albian (~110 ma) both volcanic arcs rifted away from northeast Asia due to slab rollback.
The Olyutorsky back-arc basin continued to widen during the Late Cretaceous and achieved its maximum width in
the early Campanian (~80 Ma). Subduction resumed beneath northeast Asia, drawing the arcs back towards Asia
and collapsing the back-arc basin. The Olyutorsky arc collided with northeast Siberia in the Paleogene. After the
Philippines rifted away from eastern Asia, the Philippine back-arc basin continued to widen forming the Philippine
Sea during the latest Cretaceous – late Eocene (Chron 16).
96
i. Deccan Flood Basalts
Though strictly not a plate tectonic event, the eruption of the Deccan Flood Basalts in India (Duncan and Storey,
1992; Coffin and Eldholm, 1994; Courtillot, et al., 1988; Courtillot et al., 1990; Courtillot, 1999; Courtillot and
Renne, 2003) ranks as the most significant tectonic/volcanic event of the Late Cretaceous. The Deccan Flood
Basalts erupted from fissures centered in west-central India 66 million years ago. Lava flows were magnetized in
both normal and reverse directions, corresponding to chrons C30, C29R and C29. Approximately 25% as large as
the massive West Siberian flood basalts, the Deccan flood basalts covered more than 500,000 km2 and buried India
to a maximum depth of 2.5 kilometers in the Western Ghats.
According to the best estimates, the eruption of the Deccan flood basalts predates the end of the Cretaceous by
about 0.5 million years (Miller et al., 2010; Renne et al., 2015; Schoene et al., 2019). Due to the release of massive
amounts of CO2 into the atmosphere, there was a brief period of moderate global warming prior to the start of the
Cenozoic (Olsson et al., 2002; Ravizza and Peucker-Ehrenbrink, 2003).
One of the most interesting coincidences of geological history is the fact that the eruption of the Deccan Flood
Basalts and the impact of the bolide at Chicxulub were nearly simultaneous. This has led some researchers to
speculate that it was the Deccan Flood Basalts, not the Chicxulub impact event, that caused the end-Cretaceous
mass extinction (Officer and Drake, 1985; Courtillot et al., 1988; Keller et al., 2008). Others have suggested that the
Deccan flood basalts were also triggered by a bolide impact in the vicinity of Mumbai (Bombay High; Chatterjee et
al., 2006). Still others suggest that there may be a causal relation between the Chicxulub impact event and the
eruption of the Deccan Flood basalt. Eugene Shoemaker (personal communication, ~1990) noted that the location
of the Chicxulub impact site at 65.5 Ma (25˚N, 70˚W) was nearly antipodal to the location of the Deccan flood
basalts (25S, 55E). It is proposed that, like a bullet piecing an apple, the shock waves from the impact site passed
through the Earth's mantle and core, and were refocused on the other side of the Earth, enhancing eruptions that
were already underway (Richards et al., 2015; Renne et al., 2015).
2. Paleogeography during the Cretaceous
1) Overview
This section describes the mountains, landmasses, shallow seas, rivers, landbridges, and oceanic gateways that
characterize the paleogeography of the Cretaceous. These paleogeographic features are intimately connected with
the changing climate and plate tectonic activity of the Cretaceous.
There were no major continent-continent collisions during the Cretaceous. The lack of major mountain-building
episodes, plus relatively fast episodes of seafloor spreading during the Early Cretaceous resulted in high sea levels
97
and extensive continental flooding by the mid-Cretaceous. Unlike the modern world, the continents of the
Cretaceous world were flooded by warm, shallow (epeiric) seas. During much of the Cretaceous, sea level was ~50
- 150 meters higher than modern sea level. During the Cretaceous, 11% of the continents were flooded, compared
with the modern figure of 7%. Small ephemeral ice caps may have covered the South Pole during parts of the Early
and Late Cretaceous, contributing to numerous rapid episodes global sea level change. Similarly, a precipitous,
though short-lived, fall in sea level may have occurred at the end of the Cretaceous due to the KT impact-winter
and the growth of a very short-lived ice cap on Antarctica.
2) Cretaceous Mountain Ranges
Four distinct kinds of uplift characterize Cretaceous mountainous areas: 1) old collisional mountain belts, which
were remnants of either late Paleozoic or the early Mesozoic collision and accretion; 2) Andean-style uplift due to
subduction beneath a Circum-Pangean Ring of Fire or along the northern margin of Tethys; 3) young rifted margins
with elevated rift shoulders, or 4) LIPs and hot spot activity.
The most prominent old collision belts were: the Caledonides - Northern Appalachians, the remnants of the
equatorial Central Pangean Mountain Range, the Urals, the Tasman-Trans-Antarctic-Cape-Sierra de la Ventana
ranges, and the multiple generations of Andean-type margins and collapsed back-arc basins in Central Asia (i.e.,
the Timanides, Tien Shan, Qinling-Qilian Shan, Songpan Ganze, Tien Shan, and trans-Mongolian accretionary zone),
Sengor et al., 1988, 1993, 2014a,b; Sengor and Natalin, 1996; Ziegler et al., 1996; Windley et al., 2007).
The youngest collisional events that produced significant relief were the Cimmerides (Middle Triassic – Early
Jurassic), the accretion of the Stikine (latest Jurassic) and Wrangellia arc (late Albian, 100 Ma; Monger and
Nokleberg, 1996; Monger, 2008) and the subsequent growth of the Early Cretaceous Canadian plutonic complex,
the collapse of the Neuquen Basin in western Argentina (latest Jurassic – earliest Cretaceous), and the diachronous
(west to east) closure of the Amurian Seaway (Triassic – Early Cretaceous; Zonenshain et al., 1990). Andean-type
subduction throughout the Cretaceous generated long, linear mountain chains (South America, Canadian
Cordillera, eastern China, and the northern margin of Tethys) or continental island arcs (e.g., Greater Antilles,
Zealandia, Olyutorska – Peninsular Alaska arc), and oceanic volcanic arcs (Philippines).
98
Figure 45. Cretaceous sea level derived from continental flooding. The black dots are the control points. See Davies
and Simmons (2023) for a discussion of alternative Cretaceous sea level curves.
3) Cretaceous Sea Level
We derived the changes in Cretaceous sea level illustrated in Figure 45 by measuring the amount of continental
flooding on each of the paleogeographic maps. This approach limits our temporal resolution. We cannot identify
sea level changes that take place over time intervals less than five million years. On average, Cretaceous sea level
was 70 m higher than modern sea level. Maximum sea level occurred during the mid-Cretaceous highstand (140 -
220 m) with a subsidiary peak (50 – 70 m) during the early Aptian. Sea level was relatively low (0 - 20 m) during the
earliest part of the Cretaceous. There was a precipitous drop in sea level during the late Campanian and early
Maastrichtian (~75 Ma). A table with the exact sea level values, as well as the sea level estimates for several other
eustatic curves, is given in the Supplemental Materials.
One of the outstanding puzzles of the Cretaceous is the cause and magnitude of eustatic sea level change. We can
deconstruct Cretaceous sea level change into eight factors: 1) changes in the relative area of continental
lithosphere and oceanic lithosphere, 2) changes in the volume of the ocean basin as a result of changes in the
average age of the ocean floor, 3) the volume of oceanic crust produced by excess marine volcanism (e.g.
seamounts and LIPS), 4) the relative length of mid-ocean ridges versus deep ocean trenches, 5) the number and
size of back-arc basins, 6) the amount of sediment fill along passive continental margins, 7) the amount of water
sequestered in continental aquifers and importantly, 8) the volume of continental ice caps. Wright et al. (2020)
provide an excellent review of several of these factors (2, 3, 5, and 6). They estimated that sea floor spreading in
99
young back-arc basins may contribute 120 – 150 m to sea level rise; the addition of passive-margin sedimentary
prisms and LIPs to the ocean basins may raise sea level by 75 – 165 m, and 45 m, respectively. A recent review of
Phanerozoic eustasy by van der Meer et al. (2022), estimated the contribution of continental ice caps to sea level
change at ~90 m.
Figure 46. The changing area of the continents during the late Neoproterozoic and Phanerozoic (0 – 750 Ma). The
effects that rifting and terrane accretion (increase in area) and continental collisions (decrease in area) had on
continental area are labelled. The long-term trend is indicated by the dotted line.
a. Continental Area
Figure 46 illustrates the changing area of continental lithosphere during the Phanerozoic. Changes in continental
area are primarily controlled by plate tectonic processes that either stretch or compress the continental
lithosphere. The area of the continents increases during the pre-drift rifting phase and decreases as a consequence
of crustal shortening due to convergence along Andean-type margins and zones of collision. The changing area of
the continents is a primary control on global sea level. If the area of the continents increases then conversely the
area and volume of the ocean basins decreases and displaces seawater onto the continents. Conversely, If the area
of the continents shrinks then water flows back into the enlarged ocean basins and sea level falls. Continental area,
on average, steadily increased during the Paleozoic and into the Triassic (Figure 46). Maximum continental area
100
occurred during the Cretaceous. This certainly contributed to higher than average sea level during the Cretaceous
by displacing water out of the ocean basins and onto the continents.
An up and down sawtooth pattern of alternating continental extension and contraction characterizes the
Cretaceous. These changes correlate reasonably well with alternating intervals of high and low sea level. On
average, however, there was no long-term change in continental area during the Cretaceous. In contrast,
continental area and sea level fell sharply during the Cenozoic due to a series of continent-continent collisions (e.g.
India – Asia, Africa – Europe).
Figure 47. Length of mid-ocean ridges (black line) and subduction zones (dashed line) during the Cretaceous. Y-axis
= kilometers, X-axis = millions of years
b. Mid-ocean Ridge and Trench Length
Cretaceous sea level changes can be viewed in the context of the changing length of mid-ocean ridges and the
changing length of subduction zone trenches. Lithosphere production at mid-ocean ridges must be balanced by the
amount of lithosphere removed by subduction (otherwise the Earth would either expand or contract). This implicit
balance allows us to qualitatively estimate sea level changes during the Cretaceous by comparing the relative
length of mid-ocean ridges and subduction zones.
101
The basic assumption is that the length of mid-ocean ridges should be in balance with the length of subduction
zones. The modern ratio of mid-ocean ridges to subduction zones is ~2:1. The total length of mid-ocean ridges,
including transform faults, is ~100,000 km. The total length of subduction zones is ~50,000 km. The rate of
subduction is determined by the age of the subducting oceanic lithosphere. Old oceanic lithosphere is denser and
heavier than young oceanic lithosphere and, therefore, old oceanic lithosphere subducts faster (12 cm/yr) than
young oceanic lithosphere (<10 cm/yr).
Therefore, all things being equal, an increase in the length of subduction zones would necessarily cause mid-ocean
ridges to spread faster which in turn, would cause global sea level to rise. Conversely, a decrease in the length of
subduction zones would force mid-ocean ridges to slow down producing a sea level fall. In summary, the plate
tectonic factor that most controls sea level is the total length of oceanic trenches and rate of subduction.
During the Early Cretaceous, Pangea was just beginning to rift apart and there were few intra-Pangean oceans. This
required that Early Cretaceous ridges in Panthallasa would have been spreading faster to balance the amount of
subduction, which was nearly constant during the Cretaceous. Consequently, additional young ocean floor would
have been produced in Panthallasa during the early Cretaceous. Because there is a lag time on the order of tens of
millions of years from the time when rapid seafloor spreading starts and the time when the average age of the
ocean basins becomes substantially younger, the effects of this rapid sea floor spreading would not be seen until
the mid-Cretaceous.
During the Cretaceous rapid sea-floor spreading in Panthalassa was gradually augmented by the appearance of
numerous intra-Pangean mid-ocean ridges. The appearance of these new rifts and mid-ocean ridges is the classic
explanation that associates the breakup of supercontinents with global sea level rise (Pitman, 1978;
supercontinent cycle, Fischer, 1981, 1982, 1984; Nance and Murphy, 2013; Nance et al., 1988, 2017)
Though it is not statistically significant, it is interesting to note that the lengths of mid-ocean ridges and subduction
zones appear to change in tandem between 125 Ma and 90 Ma. An increase in trench length is immediately
correlated with a corresponding increase in mid-ocean ridge length. This may suggest that mid-ocean ridges
directly respond to subduction and reinforces the notion that slab-pull forces dominate the plate tectonic process.
Finally, it should be noted that trench length was gradually reduced during the Late Cretaceous (90 Ma – 65 Ma)
suggesting that a global slowing of plate motion was at least partially responsible for the fall in sea level during the
Cenozoic.
102
Figure 48. Latitudinal limits of snow (dotted line) and ice (black line) during the Phanerozoic. The latitudinal limits
of snow and ice were estimated from the global temperature model of Scotese et al. (2021). The 0˚ isotherm was
used to map the winter snowline and the -10˚C isotherm was used to map the limit of permanent ice.
The red line indicates the latitudinal limit of glacial deposits (tillites and dropstones; Boucot et al., 2013)
c. Continental Ice
The Cretaceous has long been considered to be a hothouse world (Global Average Temperature, 22˚C; Scotese et
al., 2021). The average temperature in the polar regions (>67˚ N and S) was well above freezing. Throughout most
of the Cretaceous, permanent ice only existed at latitudes > 85˚ N and S. During the Cenomanian-Turonian Thermal
Maximum, there was no snow and ice at the poles, even during the winter months. However, cool polar
temperatures may have prevailed during the Early Cretaceous (Berriasian – Hauterivian; Aptian) with a small polar
icecap in Antarctic extending down to 80˚ S with an area of ~ 4 million km2 (about the size of India). The waxing
and waning of an ice cap this size could have caused sea level fluctuations in the range of 15 – 25 m.
103
4) Cretaceous Landmasses, Landbridges and Oceanic Gateways
a. Introduction
The breakup of Pangea resulted in both the destruction of land connections (landbridges) between the continents
and the formation of new marine connections (oceanic gateways) between the world’s oceans. These plate
tectonic effects, combined with changing sea levels, produced an evolving mosaic of landmasses, shallow seas, and
deep ocean basins. In this section, we chronologically describe the creation and destruction of landbridges and
oceanic gateways during the Cretaceous. It is recommended that the reader review the animation, ”Plate
Tectonics, Paleogeography, and Ice Ages” that is provided with the Supplemental Materials while reading this
section.
104
Figure 49. Jurassic/Cretaceous Boundary (145 Ma), see Figure 1B for legend.
105
Figure 50. Early Cretaceous (latest Berriasian, 140 Ma), see Figure 1B for legend.
106
Figure 51. Early Cretaceous (Valanginian, 135 Ma), see Figure 1B for legend.
107
Figure 52. Early Cretaceous (early Hauterivian, 130 Ma), see Figure 1B for legend.
108
Figure 53. Early Cretaceous (earliest Barremian, 125 Ma), see Figure 1B for legend.
109
Figure 54. Early Cretaceous (early Aptian, 120 Ma), see Figure 1B for legend.
110
Figure 55. Early Cretaceous (late Aptian, 115 Ma), see Figure 1B for legend.
111
Figure 56. Early Cretaceous (early Albian, 110 Ma), see Figure 1B for legend.
112
Figure 57. Early Cretaceous (middle Albian, 105 Ma), see Figure 1B for legend.
113
Figure 58. Early Cretaceous (latest Albian, 100 Ma), see Figure 1B for legend.
114
b. Early Cretaceous (145 Ma – 100 Ma, Figures 49 – 58)
Sea level was low during the Early Cretaceous, as a consequence; the land areas of the Earth were mostly
connected (Figure 49). However, a notable exception was the isolation of the landmasses of the northern
hemisphere (Laurasia) from the southern hemisphere supercontinent (Gondwana). Europe was isolated from
North Africa by the western Tethys Ocean, and North America was separated from Western Gondwana (South
America and Africa) by the Proto-Caribbean Sea. The closest approach between these two supercontinents was the
narrow gap (115 km @ 145 Ma) between southern Yucatan and northernmost South America. Some inter-
continental migration may have occurred by island-hopping along the Greater Antilles volcanic arc, Bahamas hot
spot track, Azores hot spot track, and between southern Iberia and Morocco (Figure 49).
Though the supercontinent of Laurasia remained intact throughout the Early Cretaceous, rising sea level flooded
the continent and produced short-lived, isolated landmasses. For example, the land areas of Europe and Asia
though connected during the Berriasian (145 Ma), were separated by the Straits of Turgai (also known as the Obik
Sea, Ural Sea, and West Siberian Sea) during most of the Early Cretaceous (140 Ma – ~110 Ma; (Figures 50 – 56).
During the entire Cretaceous, North America was connected to Asia across the Bering Sea landbridge. This
connection was intermittent during the earliest Cretaceous (145 Ma – 125 Ma) and was finally solidified after the
collision of the North Slope of Alaska – Chukotka block during the early Albian (~110 Ma; Figure 56). Also of note
was the fact that Greater Britain (Britain plus Ireland) was an isolated island during most of the Cretaceous, but
was connected to North America via eastern Greenland during the earliest Cretaceous (145 Ma – 130 Ma; Figures
49 - 52).
East Gondwana and West Gondwana were technically separate landmasses at the beginning of the Cretaceous
(145 Ma; Figure 49); however, they were only separated by a narrow (~70 km), shallow water gap across the Davie
Strait ( < -500 m deep, Roche and Ringebach, 2022). During the Valanginian (~140 Ma), this shallow water
continental shelf was replaced by a deep oceanic separation (> -2000 m). The ocean between Madagascar and
Africa widened during the remainder of the Early Cretaceous and achieved maximum width (500 km) when
Madagascar stopped moving relative to Africa (105 Ma; Figure 57).
The East Gondwana landmass composed of India, Madagascar, Australia, Antarctica, and Zealandia remained intact
throughout most of the Early Cretaceous (145 Ma – 125 Ma; Figures 49 – 51). During the last half of the Early
Cretaceous (Aptian – Albian, 120 Ma – 100 Ma; Figures 54 – 56), a tenuous connection between Antarctica and
India may have been maintained by island-hopping along the Rajmahal – Kerguelen hotspot track.
115
Figure 59. Mid-Cretaceous (Cenomanian, 95 Ma), see Figure 1B for legend.
116
Figure 60. Mid-Cretaceous (latest Turonian , 90 Ma), see Figure 1B for legend.
117
Figure 61. Late Cretaceous (Santonian - Coniacian, 85 Ma), see Figure 1B for legend.
118
Figure 62. Late Cretaceous (early Campanian, 80 Ma), see Figure 1B for legend.
119
Figure 63. Late Cretaceous (late Campanian, 75 Ma), see Figure 1B for legend.
120
Figure 64. Late Cretaceous (Maastrichtian, 70 Ma), see Figure 1B for legend.
121
Figure 65. Cretaceous/Paleogene Boundary (65 Ma), see Figure 1B for legend.
122
c. Late Cretaceous (100 Ma – 65 Ma; Figures 58 – 65)
The early part of the Late Cretaceous (~100 Ma – 85 Ma) was a tectonically active time. Australia and Antarctica
began to break apart during the Cenomanian (95 Ma). During the Cenomanian-Turonian (95 - 90 Ma), a deep
ocean trough separated the conjugate margins of Australia and Antarctica, but the two continents were still
connected by a landbridge in the vicinity of Tasmania (Figures 59 and 60). Though a shallow water connection was
maintained across the Tasman Straits throughout the remainder of the Cretaceous, the Antarctic and Australian
landmasses first became separated due to rising sea levels during the Coniacian – Santonian (~85 Ma; Figure 61).
Zealandia (modern New Zealand) began to separate from eastern Australia during the late Cretaceous (~100 Ma).
Extensive continental stretching produced a wide shallow sea with only a small area of emergent volcanic land. The
first oceanic lithosphere formed in the Tasman Sea ~ 85 million years ago (Figure 61).
The partnership between India and Madagascar ended during the Cenomanian (`~95 Ma). India together with
several small continental blocks (Seychelles, Laxmi Ridge, Mascarene Plateau, and Nazareth Bank) rifted away and
moved rapidly to the north-northeast (Figures 59 - 65).
Though South America and Africa had begun to rift apart during the earliest Cretaceous, they did not finally go
their separate ways until the late Albian (105 Ma; Figure 57). The last dinosaurs crossed the Amazon-Cape Palmas
landbridge approximately 110 million years ago. The oldest ocean floor formed in this region of the South Atlantic,
near the Romanche Fracture Zone, approximately 105 Ma (Cretaceous Quiet Zone). It is interesting to note that
the South American continent remained a single, contiguous landmass throughout the entire Cretaceous. Even
during the highstands of sea level from the Cenomanian through to the Santonian, it was possible for terrestrial
fauna to migrate from the tip of Patagonia to highlands of Venezuela. In contrast, for most of the Late Cretaceous,
the flooded African continent was divided into two landmasses: 1) Northwest Africa, and 2) Afro-Arabia. Beginning
in the late Albian (~105 Ma; Figure 57), the Trans-African Seaway crossed from Nigeria, northward through Niger-
Chad, and across Libya into the Mediterranean Sea. This shallow-water seaway finally closed during the Campanian
(~75 Ma; Figure 63).
During the Late Cretaceous, rising sea level divided North America into three separate landmasses: Laramidia,
Appalachia, and Greater Greenland. As mentioned previously, throughout the Cretaceous North America was
connected to Asia via the Bering Sea landbridge. By the earliest Cenomanian (100 Ma; Figure 58), Laramidia was
separated from Appalachia by the Mid-Continent Seaway. During times of maximum sea level (90 Ma – 80 Ma;
Figures 60 – 62), the Trans-Hudson/Labrador Seaway isolated the northern half of Appalachia from Greater
Greenland which was made up of Greenland, the Arctic Islands and four-corners region of north-central Canada. A
123
single North American landmass emerged when these seaways retreated from the continent during the
Campanian (75 Ma – 70 Ma; Figure 63 – 64).
5) Cretaceous Rivers
a. Overview
Rivers, especially at elevations > 200 m, are the great agents of erosion that level mountains and create
continental peneplains. Though a vast amount of sediment is preserved in river deltas, evidence of the rivers that
produced these accumulations are rarely preserved in the geological record. River lowlands, which are flooded and
reworked during highstands of sea level, are the first to disappear. Ongoing continental flooding erases all but the
most upland river systems. For these reasons, the Cretaceous river systems shown in Figures 67 – 82 are very
speculative. They are derived entirely from the paleo-digital elevation models (paleoDEMs) that are the basis of
the paleogeographic maps. No attempt has been made to ground-truth the predictions made by the simulation
with evidence from the geological record.
As described previously, these digital elevation models can be used to recreate ancient drainage networks. These
drainage networks represent potential river systems. To determine whether a drainage system will be filled with
water, we must add the predictions of rainfall and runoff made by global climate models. The number and length
of river systems are ultimately controlled by a combination of changes in topography and sea level.
The river systems presented here are unnatural, in the sense that they do not evolve through time. The geometry
of actual rivers and drainage systems grows and changes in almost an organic fashion. Young upland rivers are
straight and rapid; old rivers meander slowly across the plains. One river system may capture another river system
creating a mega-river network. None of these evolving patterns of river growth and maturity are shown on these
maps. The rivers are each a unique representation of the specific topography. In this regard, the location and
pattern of rivers may change drastically from one stage to another (e.g., compare the rivers of the Berriasian and
the Campanian). Nevertheless, there are some generalizations that can be made from the long-lived aspects of
Cretaceous paleogeography that have produced enduring river geometries. In the following sections we discuss
average river length and the number of rivers during the Cretaceous, the effect of climate on river systems, the
effect of plate tectonics on river basin geometry, and the Cretaceous origin of some of the modern rivers.
124
Figure 66. Average river length (black line) and the number of rivers (red line) during the Cretaceous.
b. Average River Length and the Number of Rivers.
Figure 26 illustrates that during the Phanerozoic, when sea level is high the average river length is reduced, and
conversely, when sea level is low the average river length increases. The pattern holds true for the Cretaceous
(Figure 66). Figure 66 plots the average length of major river systems (black line) during the Cretaceous. As
expected, rivers are shorter during times of high sea level due to reduce landmass area. Minimum river length
(~2500 km) occurred during the mid-Cretaceous (Cenomanian to Santonian, 95 Ma – 80 Ma). This interval
corresponds with the peak in Cretaceous sea level (Turonian, ~150 m; this study). Maximum river length occurred
during the earliest Cretaceous (~5000 km; Berriasian – Valanginian, 145 Ma – 140 Ma; Figure 67 and 68), a period
of low sea level. Figure 66 also plots the number of major river systems during the Cretaceous (red line). It is
somewhat counterintuitive that the number of river systems remained relatively constant during the Cretaceous.
During each stage of the Cretaceous, approximately 50 rivers longer than 250 km traversed the continents. The
number of rivers is apparently independent of landmass area; rivers simply become shorter as sea level rises and
longer as sea level falls. This relationship can be seen by comparing Berriasian and latest Turonian maps.
125
Figure 67. Early Cretaceous Drainage System and Rivers, Jurassic-Cretaceous Boundary (145 Ma), for Legend see
Table 11.
Figure 68. Early Cretaceous (latest Berriasian, 140 Ma), for Legend see Table 11.
126
Figure 69. Early Cretaceous (Valanginian, 135 Ma), for Legend see Table 11.
c. Effect of Climate on River Systems
As stated previously, rivers exist where the prescribed drainage system is filled with water. The presence of water
depends both on the amount of regional precipitation and the flow length of the river system. Flow length is
directly proportional to the size of the drainage basin. The largest rivers, like the Amazon River, will be found in
large drainage basins that lie within the everwet climate zone (Köppen Zone A). Major rivers can also be found in
the Warm Temperate and Cool Temperate rainy belts.
As discussed in the section on methodology, the HadCM3 climate model was used to predict the amount of
precipitation during the Cretaceous. These estimates of precipitation are generally in good agreement with the
Köppen climatic belts derived from lithologic indicators of climate and paleoclimate proxies (Burgener et al., 2023).
However, the agreement between these two independent predictors of precipitation is not perfect. Upon close
inspection of paleo-river maps, one may note several river systems that originate in regions that are predicted by
the Köppen maps to be arid and therefore should be dry river beds. A good example of such a dry river system can
be found in the arid zones surrounding the narrow South Atlantic and the northern arid belt of eastern China
during the Aptian (120 Ma).
127
Figure 70. Early Cretaceous (early Hauterivian, 130 Ma), for Legend see Table 11.
Figure 71. Early Cretaceous (earliest Barremian, 125 Ma), for Legend see Table 11.
128
Figure 72. Early Cretaceous (early Aptian, 120 Ma), for Legend see Table 11.
d. Effect of Plate Tectonics on River Systems
Rivers flow from high areas to low areas, therefore the quickest way to change the geometry of a river basin is to
alter the gradient of the topography. During the Cretaceous, several events profoundly affected paleotopography.
These events can be characterized as: the formation of Andean margins adjacent to subduction zones, continental
collision, uplift associated with hot spot activity prior to continental rifting, and the formation of a medial rift valley
and rift shoulders during the early phases of continental breakup.
During the Cretaceous, Andean margins were common along the western margins of the Americas and in eastern
Asia. Rivers adjacent to these Andean margins tended to flow parallel to the mountain range in the depression
(foredeep) caused by the thrusting of the mountain range over the adjacent craton. Examples of these mountain-
parallel river systems are the MacKenzie, Orinoco, Patagonian, and Amur rivers (see Table 11).
Though there were no major continental collisions during the Cretaceous, there were three minor collisions that
affected the geometry of river basins. These tectonic events coincidentally took place in the Aptian (121 Ma – 113
Ma). The Mongol-Okhotsk ocean closed by the Aptian, diverting the Amur River eastwards (Figures 73 and 74).
Previously the Amur River flowed south-to-north along the eastern margin of Asia (Figure 71). In northeast Siberia,
the collision of Kolyma along the Verkhoyansk mountains created an elongate NW-SE foredeep through which the
129
Lena River flowed into the Arctic Ocean (Figures 73 – 75). Finally, the North Slope block collided with the central
Alaska during the late Aptian (115 Ma) creating the Yukon river system (Figure 73).
During the Early Cretaceous (145 Ma – 130 Ma; Figures 67 and 70), West Gondwana began to rift apart. Uplift in
eastern most Brazil and central Africa, created no less than eight major river systems (Figure 69). These rivers
radiated away from the core of West Gondwana (proto-Amazon, Senegal, Nile, trans-Saharan, Arabian, Lamu,
Zambezi, and Patagonian rivers). A similar phase of Early Cretaceous uplift between Australia and Antarctica
created the proto-Darling-Murray river system (Figure 73) and a vast, putative river system across eastern
Antarctica (Figure 74).
By the Barremian (125 Ma), a long, N-S river (Australis river) flowed down the rift valley that separated South
America and Africa (Figure 71). Within 10 million years, emergent rift shoulders along the margins of the South
Atlantic forced a new set of rivers to flow westward across South America (proto-Amazon and Parana rivers) and
eastward across Africa (proto-Congo and Namib rivers) (Figure 73). It should be noted that these rivers flowed in
directions that were opposite of their modern counterparts (Table 11).
Figure 73. Early Cretaceous (late Aptian, 115 Ma), for Legend see Table 11.
130
Figure 74. Early Cretaceous (middle Albian, 105 Ma), for Legend see Table 11.
Figure 75. Early Cretaceous (latest Albian, 100 Ma).
131
Figure 76. Mid-Cretaceous (Cenomanian, 95 Ma), for Legend see Table 11.
Figure 77. Mid-Cretaceous (latest Turonian , 90 Ma), for Legend see Table 11.
132
Figure 78. Late Cretaceous (Santonian - Coniacian, 85 Ma), for Legend see Table 11.
Figure 79. Late Cretaceous (Early Campanian, 80 Ma), for Legend see Table 11.
133
Figure 80. Late Cretaceous (Late Campanian, 75 Ma), for Legend see Table 11.
Figure 81. Late Cretaceous (Maastrichtian, 70 Ma), for Legend see Table 11.
134
Figure 82. Cretaceous/Paleogene Boundary (65 Ma), for Legend see Table 11.
135
Figure 83. Major Modern River Systems, for legend see Table 11. Actual modern rivers – pink lines (top), Computer
simulation – light blue lines (bottom).
136
e. Cretaceous Origin of Some Modern Rivers
Some of the rivers alluded to in the previous section are unique to the Cretaceous (e.g., Australis, trans-Saharan
Arabian, Lamu, Namib, and Patagonian). This is not surprising. Over the vastness of geological time, the
topography of the continents has changed substantially, creating new drainage basins and rivers. It is a little
surprising, however, that many of these Cretaceous rivers have modern counterparts (e.g., proto-Amazon, Amur,
proto-Congo, Lena, and MacKenzie).
Figure 83 illustrates some of the major modern river systems with lengths greater than 2500 km. Table 11 lists
both the newly described Cretaceous rivers as well as Cretaceous rivers that have modern counterparts. A check
mark, “√”, indicates that the named river can be found on the corresponding the map. A simple naming convention
has been used to describe important aspects of these rivers. The term “proto-River name” implies that the
mapped river may be the ancestor of its modern counterpart, but that the geographic location, extent, and flow
direction may be somewhat different. The suffix “-r” indicates that the geographic location of the paleoriver is
similar to its modern counterpart, but the flow direction was reversed in the Cretaceous (e.g., Amazon-r). The
suffix “-s” indicates that the geographic location and the flow direction are the same as its modern counterpart,
but the extent of the river system is much smaller (e.g., Mekong-s). The suffix “-f” indicates that the river system
was flooded at that time. In a few cases the letters “N, E, S, and W” are used to indicate the direction that the
paleoriver flows.
A quick review of Table 11 finds that more than half of the 35 longest modern river systems had counterparts
during much of the Cretaceous. The most notable are the Nile, Amazon, Yangtze, Mississippi, Huang-He, Congo-r,
Zambezi, Amur, Lena, MacKenzie, Mekong, Niger, and Murray-Darling.
The modern rivers that do not have Cretaceous counterparts fall into 3 categories: 1) rivers that occupied portions
of the continents that were mostly flooded during the Cretaceous (e.g., Yenisey, Ob-Irtysh, Volga, and Rio Grande),
2) rivers that were created in response to plate tectonic events during the Cenozoic (e.g., Danube, Rhone, Rhine,
Loire,Tigres-Euphrates, Indus, Syr, Amu, Ganges, Brahmaputra, Columbia, and Colorado), and 3) modern rivers in
eastern South America that for most of the Cretaceous were located within the southern arid belt (Sao Francisco,
Paraná, Paraguay-Uruguay) and therefore, were not filled with water.
Some of the most intriguing and putative Cretaceous river systems that do not have any modern counterparts can
be found in the Early Cretaceous (145 Ma – 100 Ma), the Patagonian, trans-Asian, trans-Saharan, trans-Antarctic ,
and trans-Indian rivers; and Late Cretaceous (100 Ma – 65 Ma), the Namib, Greenland and trans-Europe rivers.
137
Four major river systems, which have no modern equivalents, span the entire Cretaceous: the Arabian, Lamu,
trans-Hudson, and trans-Siberian rivers.
6) Oceanic Circulation during the Cretaceous
a. Introduction
In this final section on paleogeography, we move from the continents to the ocean basins and discuss oceanic
circulation during the Cretaceous. The circulation of the surface of the ocean is primarily driven by the prevailing
winds (Figure 84). The tropical Easterlies and the temperate Westerlies push the ocean waters ahead of them. At
the Equator, strong ocean currents generally move from east to west (Figure 85). At mid-latitudes (35˚ - 40˚ N and
S), the ocean currents move predominantly from west to east. When an ocean current runs into a continent, it is
deflected to the north or south depending on the hemisphere and the latitude. This deflection is due to the Coriolis
force which tends to drive the currents poleward.
138
Figure 84. Modern Surface Winds
139
Figure 85. Modern Surface Oceanic Circulation (source: https://en.wikipedia.org/wiki/Ocean_current)
At mid-latitudes in the northern hemisphere (~30˚ N), the gyres off the western coasts of continents spin clock-
wise; at high latitudes (55˚ - 65˚ N) the gyres spin counter-clockwise; and polar gyres spin clockwise. (Figure 85).
The opposite pattern is observed in the southern hemisphere: off the western coasts of southern continents, a
mid-latitude ocean current is deflected to the north forming a counter-clockwise gyre (~30 S). High latitude gyres
in the southern hemisphere rotate clockwise (Figure 85). These fundamental patterns of ocean circulation also
apply to the Cretaceous. All of the Cretaceous ocean circulation models presented here (Figures 86 – 88) follow
this general pattern and were produced by the HadCM3 global climate model. What made the Cretaceous
different from the present-day was the changing size and inter-connectedness of the ocean basins.
140
Figure 86. Oceanic Circulation, 145 Ma, 130 Ma, 120 Ma, Valdes et al. (2021). See Figure 3C for legend.
141
Figure 87. Oceanic Circulation 105 Ma, 95 Ma, 90 Ma, Valdes et al. (2021). See Figure 3C for legend.
142
Figure 88. Oceanic Circulation 85 Ma, 80 Ma, 70 Ma, Valdes et al. (2021). See Figure 3C for legend.
143
b. Cretaceous Oceanic Surface Currents
At the beginning of the Cretaceous, the continents were in a Pangea-like configuration. The combined Panthallasic-
Tethys ocean basin was huge (spanning >260˚ of longitude); there were only a few, small, nearly landlocked, intra-
Pangean oceans (e.g., Central Atlantic, Somali basin, Mozambique basin, and Weddell Sea). As Laurasia and
Gondwana began to break apart, these ocean basins widened and became interconnected.
During the Early Cretaceous, the westward flowing Equatorial current collided with East Africa (Figure 86 – 88).
Approximately half of the water was diverted northward, where it flowed northwest along the coast of Arabia,
then along the southern Mediterranean, and eventually recycled eastward after hitting a dead end in the western
Mediterranean. This current was active throughout most of the Cretaceous, but was closed down in the late
Cretaceous (Campanian, 75 Ma; Figure 86) due to the narrowing of western end of Tethys.
It should be noted that there was never a strong, long-lasting connection between western Tethys and the Central
Atlantic. Two factors prohibited a strong, through-going connection: 1) the passage south of Iberia was always
narrow and shallow, and 2) during the Cretaceous, the northern coast of Africa was at latitudes of ~30˚ N and
connections between the equatorial currents on either side of were inhibited by opposing winds. Rather than a
single through-going ocean current connecting Tethys and the Atlantic Realms , two stable, counter-clockwise
gyres occupied the Central Atlantic and western Tethys from the Aptian (115 Ma) until the Campanian (75 Ma).
In the southern hemisphere, a strong current flowed south along the east coast of Africa and then flowed west to
east along the Indo-Australian margin of Gondwana (Figures 86 and 87). This current intensified during Barremian
(125 Ma; Figure 88) due to the opening of the Somali basin. A strong, warm, eastward-flowing current remained a
dominant feature along the Indo-Australian margin of Gondwana until the Cenomanian-Turonian (95 Ma - 90 Ma,
when cooler waters began to flow into southern Tethys from high southerly latitudes (Weddell Sea) as a result of
the growing separation of India and Australia.
There were several other notable ocean currents and circulation changes during the Cretaceous. Throughout most
of the Cretaceous (125 Ma – 75 Ma), there was a well-organized, clockwise rotating gyre in the enclosed Arctic
Ocean. During the Coniacian (90 Ma -85 Ma), the South Atlantic had grown large enough to sustain a stable,
counterclockwise gyre. The widening Proto-Caribbean Sea permitted a through-going connection between the
Central Atlantic and the eastern Pacific Ocean from the late Aptian (115 Ma) through to the latest Cretaceous. And
finally, by the late Maastrichtian, India had moved far enough north to begin to block the westward flowing
Equatorial current. Oceanic circulation is a very dynamic feature. It is recommended that the reader view the
144
animation of ocean circulation provided in the Supplemental Materials. The Supplemental Materials also contain
simulations of oceanic circulation for every stage of the Cretaceous.
3. Paleoclimate during the Cretaceous
1) Introduction
Cretaceous climate has been a popular research topic during the past 50 years. Some of these studies have
focused on changes in global temperature and associated oceanic anoxic events (OAE) events. Other research has
modeled the Cretaceous transition to a more angiosperm-dominated flora and its effect on terrestrial ecosystems.
Additional Cretaceous climate research has been the abundance of floral evidence for warm temperatures in polar
regions (Spicer at al., 1994; Herman and Spicer, 1996; Bice et al., 2003). In addition, there have been numerous
computer simulations of Cretaceous paleoclimate. Excellent reviews of these topics can be found in Parrish (1998),
Huber et al. (2000), Bice et al. (2003), Summerhayes (2015), and Skelton(2000). In this section, we consider
Cretaceous climate in the broader context of Mesozoic climate change, chronologically review climatic events
during the Cretaceous, and review the changing patterns of regional precipitation.
For a overview of climate change during the Cretaceous, the reader is referred to the set of Cretaceous paleo-
Köppen maps (Figures 89 - 97) from Burgener et al. (2023). These maps are the first comprehensive description of
the changing extent of the warm wet tropical belt, the dry arid belt, the warm temperate belt, and cool temperate
belt during the Cretaceous that are based entirely on geological, geochemical, and paleontological proxy data. All
of the paleoclimate nomenclature used in this review is summarized in Table 12. In addition, the reader is referred
to the paleoclimatic reconstructions, digital data sets, and animations illustrating the temperature, precipitation,
and ocean circulation for each stage of the Cretaceous in the Supplemental Materials.
145
Figure 89. Paleo-Köppen Map for earliest Cretaceous (145 Ma; from Burgener et al., 2023) See Figure 1C for
legend.
Figure 90. Paleo-Köppen Map for Early Cretaceous (Hauterivian, 130 Ma; from Burgener et al., 2023). See Figure 1C
for legend.
146
Figure 91. Paleo-Köppen Map for Early Cretaceous (Aptian, 120 Ma; from Burgener et al., 2023). See Figure 1C for
legend.
Figure 92. Paleo-Köppen Map for Early Cretaceous (Albian, 105 Ma; from Burgener et al., 2023). See Figure 1C for
legend.
147
Figure 93. Paleo-Köppen Map for mid-Cretaceous (Cenomanian, 95 Ma; from Burgener et al., 2023). See Figure 1C
for legend.
Figure 94. Paleo-Köppen Map for mid-Cretaceous (Turonian , 90 Ma; from Burgener et al., 2023). See Figure 1C for
legend.
148
Figure 95. Paleo-Köppen Map for Late Cretaceous (Santonian - Coniacian, 85 Ma; from Burgener et al., 2023). See
Figure 1C for legend.
Figure 96. Paleo-Köppen Map for Late Cretaceous (Campanian, 80 Ma; from Burgener et al., 2023). See Figure 1C
for legend.
149
Figure 97. Paleo-Köppen Map for Late Cretaceous (Maastrichtian, 70 Ma; from Burgener et al., 2023). See Figure
1C for legend.
2) Cretaceous Climate in the Context of Mesozoic Climate Change
Two extreme climatic events bracket the Mesozoic Era. The Permo-Triassic Thermal Maximum (PTTM; global
average temperature ~33˚C) and the resulting Permo-Triassic Extinction Event (PTEE; 251 Ma; Wignall, 2001, 2015;
Song and Scotese, 2023) define the start of the Mesozoic Era. The End Cretaceous Extinction (ECT, 66.1 Ma; global
average temperature ~10˚C) punctuates the end of the Era. Literally times of “Fire and Ice”, these two extreme
paleoclimatic events could not have been more different. The Permo-Triassic Thermal Maximum was caused by
massive volcanic eruptions that released an unprecedented amount of CO2 into the atmosphere producing torrid,
global temperatures (> 40˚C in the Tropics), whereas the End Cretaceous Extinction was the result of a bolide
impact (Chicxulub) that plunged the world into a frigid Impact Winter that collapsed of the marine and terrestrial
ecosystems, killing-off ammonites, large marine reptiles, and the dinosaurs; setting the stage for the rise of the
mammals. Table 13 lists some of the key sources that describe in detail these eventful times.
Figure 98 illustrates how global average temperature (GAT) changed during the Mesozoic. The Mesozoic lasted 186
million years and can be divided into eight pairs of relatively cooler (C) or warmer (W) intervals (chronotemps Mz01
– Mz08, Scotese et al., 2021). The early part of the Triassic was the warmest part of the Mesozoic with a global
150
average temperature of 29˚C (Table 12; chronotemp MzW01). The Jurassic was more equable. It started out
relatively cool (GAT = 20˚C, chronotemp MzC04), warmed up a bit during the Early Toarcian Thermal Maximum
(GAT = 24˚C, chronotemp MzW05), and then cooled down at the end of the Jurassic and through the first half of the
Early Cretaceous (Late Jurassic – Early Cretaceous Cool Interval, chronotemp MzC06, 150 – 125 Ma; global average
temperature ~20˚C).
Figure 98. Global Average Temperature (GAT) during the Mesozoic (Scotese et al., 2021). Warmer Intervals (W1-8) =
white, cooler intervals (C1-8) = black, dashed black line = temperature cutoff for large, permanent polar icecaps,
Time scale from International Chronostratigraphic Chart v2020/01. Refer to Table 12 for more information about
each chronotemp.
It should be noted that, with the exception of the End Cretaceous Impact Winter, Mesozoic global average
temperatures never dipped below 18˚C, though this limit was closely approached during the earliest Cretaceous.
Permanent polar ice caps can only form if the global average temperature is less than 18˚C. Above that
temperature, during the summer months the average annual temperature of the polar region (>67 ˚ N and S) is
151
warm enough to melt any accumulation of winter snow and ice. Only when the average global temperature is
below 18˚C can snow and ice remain year-round promoting the growth of large, polar ice caps. A transition zone
exists when the global average temperature ranges between 18˚C and 22˚C. Snow and ice will be present during
the winter months and patches of permanent ice may develop close to the Pole or at relatively high elevations
(>500 m). When global average global temperatures are greater than 24˚C, even the winter months will be devoid
of snow and ice.
Figures 99 illustrates both the changing global temperature during the last 540 million years and the associated
change in the Pole-to-Equator temperature gradient. With a quick scan of this figure, one can see how global
temperature warmed and cooled during the Cretaceous and how temperature varied between the Pole and the
Equator. The Impact Winter at the end of the Cretaceous especially stands out.
Figure 99. Phanerozoic Heat Map. Colors indicate pole-to-Equator temperature gradient; dotted line is the 40
million year running average. The blue shading indicates polar icecaps (Scotese et al., 2021).
The Pole-to-Equator temperature gradient provides an excellent snapshot of global climate and provides key
insights into regional climatic patterns. Figures 101 – 109 plot pole-to-equator temperature diagrams for nine
time-intervals during the Cretaceous. These time intervals correspond to the Köppen maps in FIgures 89 - 97
(Burgener et al., 2023). Each pole-to-equator diagram is accompanied by a map that illustrates paleotemperatures
(Figures 101 - 109; Scotese et al., 2021). From the pole-to-equator diagrams one can derive the global average
temperature, the tropical temperature, the polar temperature, the latitudinal temperature gradient, the
152
equatorward limit of permanent ice, and the temperature of the deep-sea (Scotese et al., 2021). Pole-to-Equator
diagrams for each stage of the Cretaceous are provided in the Supplemental Materials.
3) Chronological Review of Climate during the Cretaceous
Figure 100. Tropical, Global Average, Deep Ocean, and Polar Temperatures during the Cretaceous (after Scotese et
al., 2021 and Grossman and Joachimsky, 2022).
a. Overview
Figure 100 illustrates the changing tropical temperatures (red line, Scotese et al., 2021; orange dashed line,
Grossman and Joachimski, 2022), global average temperatures (black line), deep ocean temperatures (blue line),
and polar temperatures (light blue line) from the late Jurassic to the late Eocene (Scotese et al., 2021). A cool
global climate persisted from the late Jurassic into the early Cretaceous (Scotese et al., 2021). Both polar regions
had cold winters, with snow and ice, and only a small (India-sized), ephemeral ice cap at the South Pole (Figures 49
-52). The climate warmed during the Albian (~110 Ma) and reached a high point for the Cretaceous during the
153
Cenomanian-Turonian Thermal Maximum (94 Ma). At that time, the global average temperature (~28˚C) was
nearly double the modern temperature; tropical temperatures exceeded 40˚C along the Equator during the
summer months. During the Cenomanian-Turonian, it would have been difficult for large terrestrial animals, like
dinosaurs, or shallow marine fauna to have survived these extremely warm (i.e. lethal > 40˚ C) temperatures. You
could have fried an egg on a Cenomanian-Turonian rocky pavement. The global climate cooled significantly during
the Late Cretaceous (Mansour and Wagreich, 2021), receiving a “coup de grace” with the arrival of the Chicxulub
bolide (66 Ma). The subsequent impact-winter caused global temperatures to fall briefly to ice age levels (10˚-
13˚C). For more information regarding the climatic events described in the following text, the reader is referred to
Table 12.
154
Figure 101. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (Jurassic – Cretaceous
Boundary 145 Ma). See Figure 2A for legend.
155
Figure 102. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures. (Hauterivian, 130 Ma).
See Figure 2A for legend.
156
Figure 103. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (early Aptian, 120 Ma).
See Figure 2A for legend.
b. Early Cretaceous Climate
The Kimmeridgian Warm Interval (MzW06, 160-150 Ma) was followed by a long interval characterized by cool, but
not frigid, temperatures (Tithonian-early Barremian Cool Interval, MzC06 150 – 126 Ma). The Tithonian-early
157
Barremian Cool Interval was punctuated by isolated warm events (Weissert (136 Ma) and Faraoni (131 Ma)
thermal excursions.
Global climate during the earliest Cretaceous (Berriasian to Barremian) can be characterized as something in-
between a hothouse and an icehouse (Frakes et al., 1992). The average global temperature was about 17˚ C. This
was five degrees warmer than the Late Cenozoic Icehouse (12˚ C), but seven to nine degrees cooler than the mid-
Cretaceous – Paleogene Hothouse (MzW08 and CzW01). Evidence for more temperate climatic conditions is based
on the occurrence of dropstones, glendonites, and a few tillites (pebbly mudstones) (Boucot et al., 2013, Rogov et
al., 2023) at polar latitudes that co-occur with evidence of temperate forests (coal, plant fossils) and dinosaurs.
Dropstones of Early Cretaceous age (Berriasian/Barremian) are widespread in South Australia, Queensland, New
South Wales, and the Northern Territory of Australia (Boucot et al., 2013). Glendonites occur in South Australia
and New South Wales. In the northern hemisphere, there are dropstones in Siberia and Svalbard, as well as
glendonites in northern Siberia, Svalbard and the Arctic Islands (Grasby et al., 2017; Brassell, 2009; Frakes and
Francis, 1988; Frakes and Francis, 1990,1993; Frakes et al., 1995; De Lurio and Frakes, 1999; Vickers et al., 2019).
The best interpretation for this mixture of cool and warm climatic indicators is that it was cold enough in the
winters for lakes and rivers to freeze over. Snow covered the ground and there were glaciers at higher elevations
and possibly a small ice cap (< 4 million km2; India-sized) at the South Pole. In the summer months it was warm
enough at high latitudes to support the growth of lush vegetation and an influx of dinosaurs migrating in from
warmer regions.
It is interesting to note that the Kimmeridgian Anoxic Event (Jenkyns et al., 2012) is the only potential oceanic
anoxic event for the time interval spanning the Middle Jurassic (Aalenian, 174 Ma) to the early Cretaceous
(Barremian, 128 Ma). It does not appear to be as widespread as the Cretaceous OAE events. The lack of anoxic
basins during the earliest Cretaceous seems quite unusual in light of the fact that there were many restricted
marine basins that would have been ideal habitats for anoxia to develop. The lack of OAEs may have been due to
the fact that the oceanic bottom waters during the Middle and Late Jurassic through to the earliest Cretaceous
were relatively well-oxygenated. The occurrence of glendonites at high latitudes during much of the Early
Cretaceous indicates that cool, oxygen-rich bottom waters were being generated at polar latitudes preventing the
bottom waters in lower latitudes from becoming anoxic.
158
Figure 104. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (middle Albian, 105 Ma).
See Figure 2A for legend.
159
Figure 105. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (late Cenomanian, 95
Ma). See Figure 2A for legend.
160
Figure 106. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (latest Turonian, 90 Ma).
See Figure 2A for legend.
161
Figure 107. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (Coniacian – Santonian,
85 Ma). See Figure 2A for legend.
162
Figure 108. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (early Campanian, 80
Ma). See Figure 2A for legend.
163
Figure 109. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (early Maastrichtian, 70
Ma). See Figure 2A for legend.
c. Mid to Late Cretaceous – Paleogene Hothouse (128 Ma – 39.4 Ma)
If one imagines where the current phase of anthropogenic global warming is heading, one immediately thinks of
the hothouse worlds of the Late Cretaceous and Eocene (Huber, 1998; Huber et al., 2000). During the Mid-
Cretaceous – Paleogene Hothouse, global temperatures were indeed much warmer than the present-day (GAT =
164
28˚C during the Late Cretaceous versus GAT = 15˚C for the present-day). It remains to be seen whether we will
succeed in warming the Earth to that degree, but at least we now know what a warmer world would look like.
The Mid-Cretaceous – Paleogene Hothouse is one of the best documented paleotemperature intervals. Table 12
lists some of the key references for this time interval and summarizes their primary conclusions regarding regional
and global temperatures. The best single source for information about the Cretaceous portion of this hothouse
interval is the O’Brien et al. (2017) summary of sea surface temperatures (SSTs) based on oxygen isotope and TEX86
temperature estimates. The TEX86 technique uses the lipid chemistry of the cell membrane of a common group of
pelagic protokaryotes (Thaumarchaeota) to estimate temperatures (Schouten et al., 2002, 2003, 2007). The
paleotemperatures are derived by measuring the ratio of key lipids (crenarchaeols). It has been noted that TEX86
temperature estimates tend to be ~50% higher than ∂18O temperature estimates (O’Brien et al., 2017; Figure 8). It
seems likely that at mid to high latitudes the unusually high temperatures recorded by TEX86 do not represent
annual average sea surface temperatures, but rather may be capturing the warmest monthly temperatures (Bijl et
al., 2023; Burgener et al., 2023).
It is notable that approximately 90% of the available TEX86 paleotemperature estimates for the Cretaceous have
been obtained from samples that are Aptian or younger in age. There are very few ∂18O temperature estimates for
times older than the mid-Albian (O’Brien et al., 2017). Fortunately, as noted earlier, geological evidence
(glendonites, dropstones, and rare tillites) from the Early Cretaceous helps to fill in these data gaps.
The mid-Cretaceous – Paleogene Hothouse began in the latest Barremian – earliest Aptian (~128 Ma) with two
thermal events, the Haupblatterton Thermal Event (Mutterlose et al., 2009) and the oldest Cretaceous oceanic
anoxic event, OAE1a, the Selli/Goguel Thermal Maximum (W7.4; Erba et al., 2015; Herrle et al., 2015; O’Brien et
al., 2017). This warm interval (MzW07; 128 -118 Ma) was followed by a cooler period during the late Aptian-early
Albian (MzC07, 116 – 108 Ma, GAT = 19˚C), which preceded the rapid ramping up to a thermal maximum during
the latest Cenomanian- earliest Turonian (MzW08, 94 – 93 Ma, GAT = 28˚C). According to O’Brien et al. (2017),
temperatures cooled gradually during the remainder of the Late Cretaceous, reaching a minimum of ~21˚C in the
late Maastrichtian, just prior to the KT impact event.
Average global temperatures during the mid-Cretaceous – Paleogene Hothouse was a ~23˚ C. Surface waters in the
Cool Temperate regions (SST = 21-23˚C) were only slightly cooler than the superheated tropical seas (29˚ C, O’Brien
et al., 2017). Oceanic bottom waters were also much warmer than the present-day (9˚ - 17˚ C; Valdes et al., 2021).
165
d. Cretaceous Oceanic Anoxic Events (OAEs)
The Selli/Goguel Thermal Maximum (OAE1a) is one of the nearly a dozen potential thermal spikes that characterize
this time interval (see Table 12). The nature and origin of these OAEs has been much debated (Schlanger and
Jenkins, 1976; Arthur and Sageman, 1994; Meyer and Kump, 2008). Previous notions that OAEs were simply the
result of rapid rises in sea level (Arthur and Sageman, 1994) or due to the stagnation of the ocean basins caused by
thermohaline density stratification (Brass et al., 1982) have fallen out of favor. There are two current schools of
thought. The first argues that the OAE’s were unusual, synchronous, global events. According to this argument,
catastrophic tectonic events triggered unusual atmospheric, biologic, geochemical, and oceanographic conditions
that promoted extensive deep-water anoxia resulting in the formation of carbon-rich black shales (Total Organic
Carbon often > 30%). Proponents of this school of thought argue that the following scenario may explain the
widespread occurrence of the carbon-rich black shales associated with the early Aptian Selli/Goguel Thermal
Maximum (OAE1a):
• The eruption of the mid-Cretaceous superplume (Larson, 1991a,b; Larson and Erba, 2015) radically
changed atmospheric and oceanic chemistry.
• Greenhouse gases from the erupting lavas, (i.e. CO2 ), warmed the Earth.
• Increased warmth accelerated chemical weathering on land; consequently, a greater flux of nutrients was
carried to the oceans.
• Land-derived nutrients, together with a higher concentration of bio-limiting metals made available by
increased hydrothermal activity associated with the extensive submarine eruptions (Duncan and Huard,
1997; Jones and Jenkyns, 2001) promoted greater marine productivity resulting in more carbon
deposition.
• The increased productivity depleted the available supply of oxygen in the water column, which led to
basin-wide anoxic or dysoxic conditions.
• Water-column anoxia, in turn, favored the preservation of carbon by inhibiting bacterial decay and carbon
recycling.
• The results were widespread carbon-rich black shales (Demaison and Moore, 1980)
A second school of thought argues that the OAEs do not represent unusual or catastrophic, global events, but
rather represent business as usual. In other words, a certain constellation of biologic, geochemical, tectonic,
atmospheric, and oceanographic conditions (e.g. Milankovitch cycles) favored the development of local, basin-
wide anoxia. The Cretaceous was unusual only in the sense that this constellation of favorable conditions was
more likely to occur than might have been expected. In essence, the paleogeography of the Aptian-Albian (and
Cenomanian-Turonian) was especially favorable for plankton blooms and high stratified sluggish oceans that
resulted in anoxic basins and promoted nutrient trapping (Meyer and Kump, 2008).
166
Each hypothesis may hold part of the answer. Each hypothesis may explain different aspects of the Earth System
processes that produce oceanic anoxic events. The catastrophic hypothesis may be the best explanation for the
rare, but truly global, mega-OAE events (i.e. Selli/Goguel Thermal Maximum (OAE1a) and Cenomanian-Turonian
Thermal Maximum (OAE2), whereas the uniformitarian hypothesis may be a better explanation for the more
frequent, regional, and less intense OAE events ( OAE1b, OAE1c, OAE1d, OAE3).
e. The Problem of Cold Polar Regions during the Cretaceous
The Aptian-Albian Cold Snap (MzC07, 116-109 Ma) separates the Selli/Goguel Thermal Maximum (OAE1a) from the
remaining Late Cretaceous OAE’s. For a brief interval in the late Aptian and Early Albian, the global climate cooled
off sufficiently for winter snow and ice to return to the northern and southern polar regions (Pucéat et al., 2003;
Jenkyns et al., 2012; Erba et al., 2015; Herrle et al., 2015; O’Brien et al., 2017). Glendonites are reported from
Ellesmere Island, Axel Heiberg Island, Svalbard, northern Greenland, and east-central Australia (Eromanga Sea)
indicating that cool bottom waters once again had chilled the deep ocean basins (Frakes and Francis, 1988; Grasby
et al., 2017; VIckers et al., 2019; Rogov et al., 2023).
The warmest Cretaceous temperatures occurred during the Cenomanian-Turonian Thermal Maximum (MzW08, 94
– 93 Ma). Second only to the Permo-Triassic Thermal Maximum, the global average temperature reached 28˚C and
the pole-to equator temperature gradient was flattened with a temperature differential of only ~20˚C degrees
between the polar region (13˚C) and the tropics (34˚C). No cold bottom-water formed during this interval. Instead,
warm, salty water from the broad, tropical shelves (Brass et al., 1982) warmed the deep oceans and resulted in
stratified oceans (Friedrich et al., 2011, 2012).
Not even a hint of ice existed at the poles during the Cenomanian-Turonian Thermal Maximum (Ziegler et al.,
1985). The presence of tropical plants and dinosaurs on Antarctica (Dettmann, 1989; Cantrill and Poole, 2012) and
above the Arctic Circle indicates that temperatures rarely fell below freezing even during the winter months (Wolfe
and Upchurch, 1987; Parrish and Spicer, 1988; Spicer and Herman, 2010; Spicer et al., 2998, 2009, 2021). Recent
descriptions of angiosperm leaf floras from Antarctica indicate that similar warm and wet conditions existed near
the South Pole during the Late Cretaceous (Hayes et al., 2006). In general, during times of hothouse conditions, the
equatorial and subtropical belts expanded slightly poleward; the Polar and Cool Temperate belt were replaced by
an expanded Warm Temperate belt that brought tropical conditions to latitudes above 50° north and south
(Paratropical Belt of Boucot et al., 2013; the megathermal rainforests of Morley, 2011).
167
Despite the overwhelming geological and paleontological evidence for warm polar regions during the Cenomanian
– Turonian Thermal Maximum, early climate simulations of the mid-Cretaceous tended to run cold and had a
difficult time modeling these warmer polar temperatures (Barron and Washington, 1982, 1985). Various attempts
have been made to modify the input parameters to the climate models to produce simulations more consistent
with the geological data. Initial attempts to fix this problem used extremely elevated levels of greenhouse gases to
warm the poles (15x modern CO2 ; Bice and Norris, 2002). There is, however, no geological support for CO2
concentrations in the Cenomanian-Turonian much above 6 times the modern value (Figure 21; van der Meer et al.,
2014; Scotese et al., 2022). The extreme high levels of CO2 needed to keep the polar regions ice-free would
necessarily make terrestrial and shallow marine habitats at low latitudes uninhabitable (Jacobs et al., 2005).
Another way to make the polar regions warmer is to modestly increase the concentration of greenhouse gases and
also modify the land cover in polar regions to a darker, denser vegetation (Otto-Bliesner and Upchurch, 1997;
Upchurch et al., 1999). The darker vegetation has a lower albedo and consequently more solar energy is absorbed
at the surface. In this model, positive feedbacks between high-latitude forests, the atmosphere, and the ocean all
contribute to significantly warmer temperatures at high latitudes during the Late Cretaceous (Upchurch et al.,
1999).
A third explanation invokes a Late Cretaceous “Super-Gulf Stream” that vigorously carried warmth from the
Equator to the Poles (Barron et al., 1993; Brady et al., 1998). Though intuitively appealing, an analysis of the
dynamics indicates that it is not possible to carry enough heat poleward using ocean currents alone. The
atmosphere must also play an important role. In addition, much like today, the paleogeography of the Late
Cretaceous presents a nearly landlocked polar region that would have been isolated from Gulf Stream-like ocean
currents.
One of the more promising approaches has been to change the high-altitude cloud parameterization that is used in
climate models like the Community Climate System Model version 3 (CCSM3, Kiehl and Shields, 2013). The high
albedo of low-altitude cumulus clouds reflects incoming sunlight back to space, which cools the Earth. Wispy, high-
altitude clouds, on the other hand, reflect thermal energy back to the surface of the Earth resulting in net global
warming (Kump and Pollard, 2008). Fewer “warm clouds” form in the modern world because anthropogenic
atmospheric pollution reduces the amount of warm cloud condensation nuclei. When cloud parameters
characteristic of pristine regions are introduced into the climate model, significant additional warming occurs,
especially in polar regions (Upchurch et al., 2015). Combined with a modest elevation in the concentration of
atmospheric CO2 (2x – 6x modern levels), the modelled temperature of polar regions remains above freezing
throughout most of the year.
168
The most radical hypothesis that has been proposed to explain the warm polar climates of the Late Cretaceous
involves a fundamental rethinking of the way the atmosphere circulates. One of the basic features of the modern
atmosphere is Hadley Cell circulation. In the Hadley Cell, warm air rises at the Equator, moves poleward, cools and
descends over the subtropical desert belt (~35° N and S; see Figure 84). In Hay’s model (Hay, 2008; Hay and Flögel,
2012, Hay et al., 2016; Hay et al., 2019), this simple, well-organized convective flow is replaced by a chaotic system
of super-cyclonic eddies, which are like mega-hurricanes. Hundreds of these mega-hurricanes would have annually
transferred vast amounts of heat from the Equator to the Poles during the Late Cretaceous. Though an intriguing
and out-of-the-box proposition, no climate model can currently simulate this complex alternative to Hadley Cell
circulation.
After reaching peak Cretaceous temperatures during the Cenomanian-Turonian Thermal Maximum, temperatures
gradually fell during the remainder of the Cretaceous. Maximum sea surface temperatures did not drop below
30˚C until late in the Santonian (84 Ma) or early in the Campanian (O’Brien et al., 2017). This gradual cooling may
have been punctuated by several ephemeral cooling events at ~85Ma, ~76 Ma, and ~71 Ma (Miller et al., 1999,
2004, 2005a,b) as evidenced by ∂18O temperature estimates from planktonic foraminifera. Also, an enigmatic
dropstone deposit of Campanian – Maastrichtian age (75 – 70 Ma) has been reported from the region of the
Anadyr River in Chukotka (Ahlberg et al., 2002).
f. The End Cretaceous Impact Winter
The fall in temperatures during the Late Cretaceous catastrophically culminated in the arrival of the bolide that
produced the 150 km diameter impact crater near the town of Chicxulub (Devil’s Tail) in northern Yucatan (Alvarez
and Alvarez, 1980; Schulte et al., 2010; Hildebrand et al., 1991). The Chicxulub impact is the largest known bolide
impact of the Phanerozoic (Spray, 2020).
The most likely scenario is that the impact event vaporized 3000 megatons of crustal material and injected this fine
particulate matter high into the atmosphere. This material fell back to Earth forming a global “clay layer”. The
Cretaceous/Paleogene boundary clay layer (Hart et al., 2012, 2013, 2014; Hart and Koutsoukos, 2015) contains
several unusual stratigraphic markers: 1) an iridium anomaly (Alvarez and Alvarez, 1980; Smit, 1999; Miller et al.,
2010), 2) microtektites (Yancey and Guillemette,2008), 3) shocked quartz (Bohor et al., 1987; Smit, 1999), and 4)
soot (from forest fires) that connect it directly to the Chicxulub impact event.
While suspended in the atmosphere, this delicate shroud of material blocked the sun and turned day into night - a
wintery night that lasted for months or years. Without sunlight, plants on land and plankton in the oceans died.
Small and large herbivores gradually starved. Without herbivores to prey on, predators then starved - all the while,
snow continued to fall (probably for several decades). As a consequence of the collapse of the food chain, ~75% of
169
all species were wiped out (Sepkoski, 1996). The effect of this extinction event on global ecosystems was second
only to the great Permo-Triassic Extinction (McGhee et al., 2013).
The ensuing Impact Winter scenario plunged the Earth into a frigid deep-freeze comparable to the coldest glacial
stages of any Phanerozoic ice age. The drastic cooling, however, was short-lived (Vellekoop et al., 2014, 2016) and
was followed by an equally short-lived period of global warming, triggered by the final massive eruption of the
Deccan LIP (Ernst, 2014; Keller et al.,2017).
The first eruptions of the Deccan LIP predate the Chicxulub impact by 1-2 million years (Chenet et al., 2008; Keller
et al., 2011, 2014; Schoene et al., 2019). It has been proposed that an earlier impact event (Shiva impact) triggered
the Deccan eruptions (Chatterjee et al., 2006); however, this hypothesis has not received much support. It seems
likely, however, that the Chicxulub impact did influence or enhance Deccan volcanism. It has been noted by several
authors that the impact site in Yucatan is nearly antipodal to the eruption site of the Deccan LIP in India. Though
the antipodal paleolatitudes are identical (26˚ N vs 26˚ S), the antipodal paleolongitudes are offset by several
thousand kilometers. Nevertheless, it seems plausible that shockwaves from the impact passed through the earth
and were reconcentrated beneath the Deccan hotspot, stimulating more voluminous eruptions (Richards et al.,
2015; Renne et al., 2015). In any event, the excess atmospheric CO2 from the Deccan eruptions caused a 4-8˚ spike
in global temperatures (Petersen et al., 2016; Bond and Grasby, 2017) that ushered in the Paleogene Warm
interval.
4. Changing Patterns of Precipitation during the Cretaceous
This section describes: 1) the persistent patterns of regional precipitation during the Cretaceous, 2) important
changes in regional precipitation during the Cretaceous, and 3) a comparison of computer simulations of paleo-
rainfall with Köppen climatic belts based on lithologic indicators of climate and paleo-precipitation proxies.
Paleoclimatic reconstructions and animations illustrating paleo-precipitation patterns for each stage of the
Cretaceous are provided in the Supplemental Materials.
1) Persistent Patterns of Regional Precipitation
As illustrated in Fig 110, the global pattern of modern precipitation is similar in many respects to the pattern of
precipitation during the mid-Cretaceous. Both the modern world and the Cretaceous world were dominated by
three regional patterns of precipitation: the tropical everwet belt; areas of high rainfall along the western margins
of continents at mid-latitudes (45˚- 50˚ N and S); and the warm, subtropical humid belt along the eastern margins
of the continents (30˚ - 40 ˚ N and S). Though the regional patterns of precipitation were similar, it is estimated
that the annual amount of precipitation was ten times higher during the Cretaceous (Bao and Hu, 2024).
170
Figure 110. Comparison of Modern (top) and mid-Cretaceous Patterns of Precipitation (bottom). See Figure 3B for
legend.
High rainfall occurs along the Equator due to the intense heating, lifting, and subsequent cooling of warm, humid
tropical air. The Equator is where the Hadley atmospheric circulation cell originates. Similar to the present-day
world, during the Cretaceous the everwet belt (Köppen zone A) was broken up by continental landmasses that
straddled the Equator.
171
The Cool Temperate Belt (Köppen zone D) is characterized by strong westerly airflow that removes moisture from
the oceans and delivers it to the western margin of continents at moderately high latitudes (45˚ - 50˚ N and S). In
the modern world this includes rainy locations such as the Pacific Northwest - British Columbia, Patagonia, and
New Zealand. During the Cretaceous, the Cool Temperate Rainy Belt also impinged upon the Pacific Northwest –
British Columbia, Patagonia, South Africa, southern Madagascar and India, as well as most of Australia.
The Warm Subtropical Belt (Köppen zone C) steals much of its moisture from either the Equatorial Rainy Belt or the
Temperate Rainy Belt. The dynamics are strongly seasonal (i.e., monsoonal). During the summer months, warming
landmasses pull air masses towards the continents. If these air masses travel over a warm, moist ocean then
torrential rains soon follow. The best example of this pattern of airflow is the modern Indian monsoons. Late in the
Cretaceous, there was an Indian-like monsoon system; however, in this case the winds were deflected towards the
southern hemisphere because India had not yet crossed the Equator.
2) Changing Patterns of Regional Precipitation
Figures 111 – 113 illustrate the variable pattern of regional precipitation during the Cretaceous (Valdes et al.,
2021). Rainfall patterns changed during the Cretaceous due to two paleogeographic effects: continental motion
across climatic belts and the opening of new ocean basins. Changing sea level, to a lesser extent, also increased
atmospheric humidity producing more frequent and intense rainfall.
172
Figure 111. Cretaceous precipitation during the early Berriasian (145 Ma), Hauterivian (130 Ma), and early Aptian
(120 Ma), Valdes et al., 2021. See Figure 3B for legend.
173
Figure 112. Cretaceous precipitation during the late Albian (105 Ma), Cenomanian (95 Ma), Turonian-Coniacian (90
Ma), Valdes et al., 2021. See Figure 3B for legend.
174
Figure 113. Cretaceous precipitation during the Santonian and Coniacian (85 Ma), Campanian (80 Ma), and
Maastrichtian (70 Ma), Valdes et al., 2021. See Figure 3B for legend.
175
a. Western Hemisphere
We begin our tour of the humid Cretaceous world (145 million years ago) in the region that would become the
South Atlanti. All of North Africa from Morocco to Arabia, together with northern South America and most of
Brazil, was occupied by a warm, wet tropical rainforest (Figure 111). Abundant rainfall across the region produced
numerous river systems flowing away from the South Atlantic into the Central Atlantic, southern margin of Tethys,
and the widening Somali Basin (see Figure 67). North America, in contrast, was much drier during the earliest
Cretaceous.
From the Valanginian to the late Aptian, (140Ma - 115 Ma), the region of high rainfall in northern South America
and northwestern Africa persisted and its southern boundary became more precisely delineated. During the
Hauterivian (130 Ma), a sharp, Sahel-like boundary separated the northern humid regions from the southern arid
regions (Figure 111). This Cretaceous Sahel was more than twice as long as its modern counterpart. It trended
diagonally across northern Gondwana from southern Colombia, bisecting Brazil, south of Gabon, across the north
Congo, and into the western Indian Ocean in the vicinity of central Somalia. Thick Aptian salt beds were deposited
in the nascent South Atlantic beneath this intense arid belt. At the same time, during the mid-Cretaceous, northern
Africa became progressively more arid (Figure 111C).
Rainfall patterns began to change dramatically as the South Atlantic opened. During the late Aptian (~115 Ma),
northern Brazil and the Ivory Coast became noticeably wetter (Figure 112). South Africa, which had been arid
throughout the Early Cretaceous, started to receive more moisture from the widening South Atlantic during the
Albian (110 Ma). West-central Africa and southernmost Africa became increasing wetter throughout the remainder
of the Cretaceous (Figures 112 and 113).
By the Cenomanian (95 Ma), the continents bordering the northern South Atlantic became much wetter (Figures
112 and 113). The high stand of sea level during the mid-Cretaceous added enough moisture to the atmosphere
that the Equatorial Wet Belt became a through-going feature stretching from the southeastern tip of Arabia to
western Ecuador (Figure 113). This equatorial everwet belt persisted through to the end of the Cretaceous.
One of the most striking changes in precipitation patterns during the Cretaceous occurred along the western
margin of the Central Atlantic in eastern North America. During the Early Cretaceous, from 145 Ma to 100 Ma,
eastern North America was characterized by an arid climate (Figure 111). Rainfall was seasonal and minimal.
Starting in the late Albian – earliest Cenomanian (105Ma -100 Ma; Figures 112), the eastern seaboard of the U.S.,
began to experience higher rates of precipitation. This trend increased steadily through the remainder of the
Cretaceous. This transition was due in part to higher sea level and the widening Central Atlantic, but primarily to
176
North America’s northward movement out of the dry subtropics and into the more humid Warm Temperate Belt
(Köppen Zone C).
Throughout the Cretaceous, Northern Europe received sufficient rainfall (~15 cm/mo). With an exception of a dry
spell in the late Aptian and Albian in Iberia (Figures 111 and 112), there was a tendency towards much wetter
climates in Europe during the mid and Late Cretaceous.
b. Eastern Hemisphere
We begin our tour of rainfall patterns in the Eastern Hemisphere with India. The changing precipitation pattern
observed in India is primarily due to India’s southward, then northward movement through three different climatic
belts. At the start of the Cretaceous (145 Ma), India was located between the arid and warm temperate zones of
the southern hemisphere (~40˚ S). Its climate was subarid (Figures 111). During the Early Cretaceous it moved
further south, arriving at 55˚S during the early Barremian (125 Ma) and enjoyed more precipitation (Figure 111). It
remained in the wet Cool Temperate belt for ~25 million years, then began to move northward, returning to the
Warm Temperate Belt. From 100 Ma to 80 Ma, India’s climate was subtropical with heavy seasonal precipitation
(Figure 112). During the Campanian and early Maastrichtian (75 Ma – 70 Ma), India moved into the southern Arid
Belt (20˚ S) and began to dry out (Figures 113). At the end of the Cretaceous, India began to accelerate towards
Eurasia and a southern hemisphere version of the Indian monsoon kicked in. By the end of the Cretaceous, India
was wetter than at any previous time period.
Like India, the record of rainfall on Australia was driven by its movement across different climatic belts. During the
Early Cretaceous, an arid Australia occupied high southerly latitudes (75˚S) in the eastern rain shadow of Antarctica
(Figure 111). By the Aptian (120 Ma), the counter-clockwise rotation of Gondwana had carried the northern
portion of Australia into warmer, wetter latitudes (60˚ S; Figures 111and 112). During the next 30 million years,
precipitation swept across Australia from north to south and by the Turonian (90 Ma), Australia was firmly
ensconced in the wet temperate belt (~55˚ S; Figure 112C), where it would remain for the rest of the period.
Eurasia, which did not move much during the Cretaceous, was divided into two precipitation regimes. South of
Mongolia, in central China and neighboring countries it was ari), except along the eastern coast. North of
Mongolia, precipitation was higher, except towards the center of Siberia, which remained arid.
d. Comparison of Precipitation Patterns: Computer Simulation versus Geological Evidence
The above description of regional precipitation during the Cretaceous is based on the paleoclimatic simulations of
the HadCM3 model (Valdes et al., 2021). To test the validity of these model predictions, we compared the
177
computer output to the patterns of precipitation revealed by the lithologic indicators and paleo-precipitation
proxies compiled by Burgener et al. (2023). 14 key localities were selected world-wide and the relative amount of
precipitation at each locality was classified as very wet, wet, subarid, or arid. The Hadley simulation results were
then compared with the precipitation estimates from the Köppen zone maps of Burgener et al. (2023). Figure 114
summarizes the results of this comparison.
Figure 114. Comparison of rainfall prediction, Burgener et al. (2023) (top cell) versus the HadCM3 model (bottom
cell). Dark green cells indicate very wet conditions, light green cells indicate wet conditions, light yellow cells
indicate semi-arid conditions, and dark yellow cells indicate arid conditions. The cells outlined in blue are an exact
match (n=42). The cells outlined in red indicate a mismatch between the computer simulation (Valdes et al., 2021)
and the Köppen map predictions (n = 4); see Figure 115.
A total of 126 localities were compared. For 42 localities (33%) there was an exact match, for 80 localities (64%)
there was a fair match (e.g., arid vs semi-arid or wet vs very wet), and for four localities (3%) there was a decided
mismatch. All of the mismatched localities were Early Cretaceous in age and in every case, the computer
simulation predicted more arid conditions.
Two of the mismatched localities are in Northern Africa during the Aptian and Albian (120 Ma – 105 Ma). As
illustrated in Figure 115, the computer simulation predicts arid to subarid conditions for North Africa, while the
Köppen reconstruction indicates much wetter conditions. In the earliest Cretaceous (145 Ma, Berriasian –
Valanginan), there are two mismatched localities. The HadCM3 model predicts that much of eastern North
America was arid, whereas the Köppen data predict a more humid environment, except along the Gulf Coast and
Mexico (Figure 115). Similarly, the HadCM3 model predicts that all of South Africa was arid during the earliest
178
Cretaceous, whereas the Köppen data predict that though the northern half of South Africa was dry, the southern
half was very wet. It is also notable that the precipitation patterns in northern South America are reversed. The
Hadley model predicts that the western part of northern South Americas was dry and the eastern part was wet.
Whereas, the Köppen model indicates that the west was wet and east was dry (Figure 115).
Figure 115. Precipitation results with the largest mismatch. (top) Aptian-Albian (120 Ma), (bottom) Berriasian -
Valanginian (145 Ma).
D. Summary
1. Introduction
This essay is organized into six principal sections: 1) plate tectonic data and methods, 2) paleogeographic data and
methods, 3) paleoclimatic data and methods, 4) plate tectonic events during the Cretaceous, 5) paleogeographic
events during the Cretaceous, and 6) paleoclimatic events during the Cretaceous. Three additional sections discuss
Cretaceous rivers systems, oceanic circulation, and precipitation patterns. The following summaries of these
sections were written by Claude (v2.1), an AI assistant created by Anthropic, with minor edits by the senior author.
2. Summary of Plate Tectonic Data and Methods
The paper describes the methodology used to construct global plate tectonic models, with a focus on applying
these techniques to build Cretaceous reconstructions. The first component of the model is mapping ancient plate
boundaries and defining tectonic elements based on the geological and geophysical record. The second
component involves modeling the hierarchical motions of these plates and terranes through time.
179
Integrating diverse lines of evidence helps constrain the plate tectonic model, including seafloor magnetic
anomalies, oceanic fracture zones, hot spot tracks, paleomagnetic polar wander paths, continental rift and collision
histories, subduction zone locations, plate motion geometries, synthetic seafloor isochrons, subducted slab
graveyard signatures in the mantle, large igneous provinces (LIPs), paleobiogeographic patterns, paleoclimate
simulations, and true polar wander.
The goal is to produce a self-consistent model of plate motions and geometries that matches all available
constraints during the Cretaceous within estimated uncertainties. As new geological or geophysical evidence
emerges, the model can be updated to build an increasingly accurate four-dimensional reconstruction of plate
configurations through time.
Adhering to a set of quantitative plate tectonic principles guides model development, helping to resolve inherent
ambiguities in the incomplete geologic record. Creating comprehensive reconstructions requires meticulous
integration of these global datasets. Continued model refinement will allow construction of accurate plate tectonic
models that capture the complex dynamics of Earth system processes through deep time.
3. Summary of Paleogeographic Data and Methods
Constructing ancient paleogeographic maps requires compiling databases of lithological and sedimentary records
that indicate different environmental depositional settings across continents over time. For example, mapping
extensive limestone distributions helps delimit past shallow marine environments. These paleoenvironmental
control points are combined with tectonic constraints from plate reconstructions for areas where the geological
record is incomplete.
The paleogeographic interpretations are converted into digital elevation models (paleoDEMs) that quantify ancient
continental topography and ocean bathymetry. The effects of gradual seafloor subsidence as ocean crust ages
away from spreading ridges can be removed to restore ancient ocean depths. Continental elevation models are
corrected for tectonic activity like mountain building or erosion as well as isostatic adjustments.
Changes in sea level over time can be estimated by calculating the percentage of flooded continental area from
paleogeographic maps across different intervals. Matching the amount of marine inundation against continental
hypsometry allows estimates of how much sea level rise would be required to produce the observed flooding.
However, the gradients of continental slopes have changed over time as well, necessitating corrections to this
basic flooding approach. Comparisons can be made between the flooding proxy sea level estimates and curves
derived from sedimentary records.
180
Integrating constraints from lithological evidence, plate tectonics models, and geophysical measurements results
in high fidelity paleoenvironmental reconstructions encoded in paleoDEMs that capture changing surface
conditions through deep time, including fluctuations in continental flooding, mountain -building, and global sea
level.
4. Summary of Paleoclimatic Data and Methods
Supercomputer simulations can be used to model past climates by specifying paleogeography, solar insolation
levels, and atmospheric CO2 concentrations as critical boundary conditions. Model outputs like temperature and
rainfall can be compared against real-world geological climate proxies like coal deposits and evaporites to evaluate
the accuracy of these simulations. Constraining the CO2 inputs is particularly important for tuning model
temperature outputs, but proxy CO2 records contain substantial gaps, a combination of proxy measurements of
CO2 and independently established Phanerozoic temperature histories can be used to fill in estimates for missing
intervals.
In addition to verifying climate model outputs, the distribution of climate-sensitive lithologies can be mapped
directly to delineate past Köppen climate belts and latitudinal temperature gradients. Statistical techniques like
Bayesian integration allow combining these lithologic indicators with isotope proxy measurements to
quantitatively estimate paleotemperatures and precipitation regimes. Comparing proxy-based and simulation-
based pole-to-equator temperature curves provides another avenue for model validation.
The increased sophistication of paleoenvironmental reconstructions also enables analysis of specific features like
ancient river systems. By applying simulated regional rainfall levels to digital paleotopography models, the likely
location and discharge of major paleorivers can be predicted through time. The length of river drainage basins
responds primarily to base level changes driven by fluctuations in sea level.
5. Summary of Plate Tectonics Events during the Cretaceous
When the Cretaceous began 145 million years ago, the supercontinent Pangea had begun to rift into several
pieces. By the earliest Cretaceous, multiple ocean basins had begun opening up between the fragments of Pangea:
Central Atlantic Ocean: Formed 175 million years ago during the breakup of Pangea. By the Early Cretaceous, the
ocean basin separating NW Africa and North America was 1500 km wide.
181
Gulf of Mexico: Opened as South America and Africa rifted from North America beginning in the Late Jurassic.
Underlain mostly by extremely stretched continental crust.
Proto-Caribbean Ocean: Formed between North and South America as North America moved northwest during the
Late Jurassic and Early Cretaceous, opening a seaway between Venezuela and Yucatan. The Proto-Caribbean Ocean
was bordered on the southwest by an active volcanic arc (Greater Antilles island arc).
South Atlantic Ocean: Rifting began in the Late Jurassic followed by the eruption of the Paraná-Etendeka large
igneous province at 135 Ma. Seafloor spreading began in the southern portion while the central South Atlantic
experienced continental rifting. The South Atlantic opened progressively from south to north from 150-110 million
years ago.
Western Indian Ocean: East Gondwana continents (Madagascar, India, Antarctica, Australia) rifted away from
Africa and South America beginning in the Early Jurassic, with strike-slip motion facilitating the opening of the
Somali and Mozambique Basins during the Cretaceous. Australia was still connected to India and East Antarctica.
Northeast Indian Ocean: A continental fragment "Argoland" rifted away from NW Australia in the latest Jurassic
(160 Ma) and crossed Tethys during the Cretaceous. The present-day location of Argoland is uncertain.
Canada Basin: Rifting between the North Slope of Alaska and the Canadian Arctic islands began 155 million years
ago.
Later in the Early Cretaceous, the remnants of Pangea fragmented into: Laurasia (North America, Europe, Asia) in
the northwest; West Gondwana (Africa, South America) in the center; and East Gondwana (India, Madagascar,
Antarctica, Australia) in the southeast. Surrounding Pangea were the Pacific Ocean, Tethys Ocean, Angayucham
Ocean and Wrangellian back-arc basin.
From 125-105 million years ago in the Aptian/Albian:
- The Central Atlantic reached 4500 km wide. The Proto-Caribbean, bordered by the volcanic Greater Antillean arc,
was at its widest 110 million years ago before a subduction flip would completely consume this old ocean basin.
- South America fully rifted from Africa by 110 million years ago, allowing the South Atlantic to link with the Central
Atlantic in an open marine gateway.
- East Gondwana continued rifting as India began separating from Antarctica around 120 million years ago. By 110
million years ago, seafloor spreading halted between India together with Madagascar, and Africa.
182
- In the Angayucham Ocean, North Slope Alaska collided with central Alaska and Northeast Siberia, closing this
ocean. In the Wrangellian back-arc basin, the Wrangellia terrane collided with western North America 105 million
years ago.
During the Cenomanian-Turonian (100-90 million years ago):
- Iberia rifted away from North America, opening the Bay of Biscay. Extension progressed into the North Atlantic
with rifting between Greenland and Europe.
- South America separated fully from Africa, allowing deep circulation between the South Atlantic and Central
Atlantic oceans.
- Greater India began its rapid northward drift towards Eurasia about 95 million years ago.
- The mid-ocean ridge between India and Antarctica jumped northward, facilitating Australia rifting from Antarctica
around 95 million years ago and opening the Tasman Sea.
- Back-arc basins related to slab rollback opened in the Okhotsk-Bering seas and south of Japan (Philippine back-arc
basin).
In the Late Cretaceous (100-66 million years ago):
- The North Atlantic rift system developed extensively, but the Labrador Sea was the only significant area that
experienced seafloor spreading.
- India moved rapidly northward, colliding with the Lut Block of Iran, which was located off the east coast of Arabia
(~70 million years ago), beginning initial contact between India and continental Asia.
- Rifting progressed between Australia and Antarctica, which were connected until the Late Cretaceous by large
volcanic plateaus, including the Naturaliste Plateau and Broken Ridge escarpment, which were generated by the
Kerguelen hot spot region.
- The Caribbean plate became a separate plate by 90 million years ago and continued overriding Proto-Caribbean
ocean crust.
- Back-arc extension related to slab rollback opened basins in the Black Sea and Caspian Sea and along the
southern margin of Asia prior to the collision of India.
Two important geological events occurred at the end of the period from 70-65 million years ago:
- The Deccan Traps volcanism began with enormous outpourings of basalt lava across western India, lasting several
million years. These emissions are linked to global climate changes.
- The Chicxulub bolide impacted the Yucatan peninsula, inciting chaos throughout the global environment and
biosphere, leading to the Cretaceous-Paleogene mass extinction event.
183
6. Summary of Paleogeographic Events during the Cretaceous
The Cretaceous saw the breakup of Pangaea and the destruction of landbridges between continents due to the
opening of new ocean basins. Average sea level was ~70m higher than today, causing extensive flooding and
epeiric seaways across continents.
Mountain building was limited because there were no major continent-continent collisions during the Cretaceous.
Existing ranges like the Urals, Transantarctic ranges, Caledonides, and Northern and Southern Appalachians
persisted. Most uplift occurred along active Andean margins and newly rifted continental margins.
Sea level highstands resulted from increased mid-ocean ridge length as a consequence of the breakup of Pangaea.
There were no large ice caps. Small, ephemeral ice caps on Antarctica contributed to modest sea level changes
(~30 m).
In the earliest Cretaceous (145 Ma), Laurasia and Gondwana were still connected via island arcs like the Greater
Antilles. The Turgai Strait separated the land areas of Europe and Asia until 110 Ma. North America was linked with
Asia throughout the Cretaceous via the Beringia landbridge. During the Early Cretaceous, Greater India,
Madagascar, Australia, and Antarctica all comprised a single landmass, East Gondwana.
As Pangaea rifted apart, Africa completely separated from South America in the late Albian at 105 Ma. During the
Cenomanian (~95 Ma), Madagascar became a separate island as India rifted away. Antarctica and Zealandia began
to rift away from Australia at that time.
The Tasman Gateway fully separated Australia and Antarctica by 85 Ma. The rifting of Zealandia created a
submerged continental mass to the east of the Tasman Sea. India rapidly headed north. High sea level during the
mid-Cretaceous flooded all of the continents, dividing North America, Eurasia, South America, and Africa into
smaller landmasses.
Appalachia, Laramidia and Greater Greenland formed when North America flooded. Low Campanian sea levels
around 75-70 Ma reconnected these landmasses briefly before the end-Cretaceous regression. Landbridges like
Beringia continued to link the northern continents.
184
7. Summary of Paleoclimatic Events during the Cretaceous
Early Cretaceous climate represented moderate “in-between” conditions continuing the Late Jurassic cool phase.
Global average temperatures were around 17°C based on a mix of climate indicators:
- Dropstones, glendonites and rare tillites at high latitudes showing cold winter conditions
- But also coal deposits and temperate forest biomes at higher latitudes
- Dinosaurs and other temperate species migrations
This suggests cold winters with seasonal sea, ice but summers warm enough to support forests and mild
temperatures. Oxygenated bottom waters prevented widespread ocean anoxia. The lack of black shales is striking
compared to later Cretaceous ocean anoxic events.
A stepwise warming trend commenced in the late Barremian, culminating in the hot Cenomanian-Turonian interval
95 - 80 million years ago. This stable, prolonged thermal maximum constitutes the Mid-Cretaceous-Paleogene
Hothouse (MCPH):
1) Mid-Cretaceous Paleogene Hothouse Stages:
a. Initial Early Cretaceous Warming (128-118 Ma)
- Haupblatterton thermal event
- Oceanic Anoxic Event 1a (Selli/Goguel)
b. Aptian-Albian Cold Snapshot (116-109 Ma)
- Brief return of polar ice and oxic deep oceans
c. Mid-Late Cretaceous Sustained Warmth (109-66 Ma)
- High but fluctuating temperatures, multiple OAEs
The Aptian-Albian Cold Snap was marked by glendonites on Axel Heiberg Island and in the Eromanga Sea, signaling
the return of cold polar bottom waters and ephemeral ice caps. Lasting for less than 10 million years, this cool spell
preceded sustained Cretaceous warmth.
2) Peak Cretaceous temperatures occurred during the Cenomanian-Turonian Thermal Maximum (CTTM; 94-93
Ma), considered to be the second warmest time in the Phanerozoic after the Permo-Triassic Thermal Maximum
(250 Ma). Global average temperatures reached ~28°C compared to ~15°C today. Polar temperatures averaged
185
13°C and tropical temperatures exceeded 34°C. As a consequence, pole-to-equator thermal gradients were
remarkably flat.
Cenomanian-Turonian Thermal Maximum Highlights:
- Tropical Ocean Temps > 34°C
- Polar Temps ~13°C, Ice-Free
- Low Pole-to-Equator Temperature Gradient
- Global Average Temperature ~28°C
Warm Cretaceous temperatures were at least partially driven by the mid-Cretaceous superplume, which released
immense volumes of CO2 into the atmosphere. Intense weathering would have also delivered more nutrients to
spur ocean productivity. Widespread anoxic conditions led to black shale deposition and the preservation of
organic carbon (multiple OAE events).
The exact mechanisms behind the low Pole-to-Equator gradients and warm polar regions remain uncertain but
likely involve enhanced atmospheric and oceanic heat transport to high latitudes. While early computer
simulations of climate had difficulty simulating such low pole-to-equator thermal gradients. Improvements in cloud
and vegetation parameterizations have helped reduce the discrepancy between simulation results and the
geological record.
3) The Late Cretaceous witnessed a very gradual cooling trend, interrupted by brief warmth spikes. Temperatures
remained relatively high, with no permanent polar ice caps. There is some evidence of ephemeral cooling events
and even dropstones in the late Campanian and Maastrichtian, but predominantly warm conditions continued.
Multiple oceanic anoxic events are associated with high ocean productivity and circulation changes.
This mostly stable, hot climate regime endured for nearly 80 million years before dramatically terminating with the
Chicxulub bolide impact 66 million years ago. Temperatures plummeted to icehouse levels in the “impact winter”
resulting from sunlight-absorbing dust and aerosols. After a short interval (1000’s of years?) warming resumed and
continued through the Paleocene and Eocene. About 35 million years ago the Cenozoic Ice House began, leading to
our modern interglacial icehouse world.
In summary, Cretaceous climate was typified by extreme warmth relative to modern conditions, with the
Cenomanian-Turonian Thermal Maximum representing peak Cretaceous heat. While the initiation, duration, and
ending of such hot climates remains puzzling, the geologic record confirms meridional gradients much flatter than
186
anything seen in the Cenozoic. Understanding the mix of continental positions, ocean gateways, atmospheric heat
transport, polar vegetation, and other variables that permitted low-latitude tropical and high-latitude polar
warmth may help climate scientists grapple with anthropogenic global warming. Unlocking deep-time hothouse
secrets can offer clues to where our climate system may be heading in the future.
8. Summary of Cretaceous Rivers
Average Cretaceous river length was shortest (~2500 km) during the mid-Cretaceous high sea level stand during
the Cenomanian-Santonian (95-80 Ma). River length peaked around 5000 km during the earliest Cretaceous
(Berriasian-Valanginian, 145-140 Ma) when sea level was relatively low. Surprisingly, the number of major rivers
(>250 km) remained fairly constant throughout the Cretaceous, averaging around 50 rivers during each stage
regardless of land area.
Climate exerts a strong control on river extent, with major river systems existing within the everwet climates and
temperate rainfall belts. Occasionally, the simulated drainage patterns predict rivers in arid regions that should
have been dry, like the proto-Congo during the Aptian.
Plate tectonics profoundly impacted Cretaceous river system geometries. Andean-style subduction zones created
foredeep basins down which rivers flowed, like the ancient Mackenzie and Amur systems. Continental collisions
during the Aptian diverted the Amur River eastward after the Mongol-Okhotsk Ocean closed. Uplift related to the
opening South Atlantic created drainage patterns that radiated away from eastern Brazil and west-central Africa.
Nearly half of Earth's 25 longest modern rivers had Cretaceous precursors, including versions of the Nile, Amazon,
Amur, and Niger. Exceptions include systems created by later tectonics or areas flooded that were flooded during
the Cretaceous sea level highstand. Intriguing Cretaceous rivers without modern equivalents, flowed across now
vanished basins and drainage pathways.
The speculative Cretaceous river systems provide a glimpse into the potential arrangements of fluvial regimes
transporting sediments across ancient landscapes. Their geometries were dictated primarily by the elevations and
climates of those prehistoric earth surfaces.
187
9. Summary of Cretaceous Ocean Circulation
Ocean surface circulation is primarily wind-driven, with tropical easterlies and temperate westerlies pushing
currents that are deflected poleward upon hitting continents due to the Coriolis force. This creates clockwise
rotating gyres in the northern hemisphere and counter-clockwise gyres in the southern hemisphere.
During the Early Cretaceous, many of the modern continents were still joined together. A strong east-to-west
flowing Equatorial Current hit the east African coast and split north and south. The northern branch flowed along
the southern Tethys for most of the Cretaceous until this sea narrowed in the Campanian.
In the south, a powerful current flowed west-to-east along the Indo-Australian coast of Gondwana from the Early
Cretaceous until the Cenomanian-Turonian (95-90 Ma). After this time, cooling waters flowed northward from
Antarctica through the widening seaway between India and Australia.
There was never a strong connection between Tethys and the Central Atlantic due to narrow seaways and
opposing wind patterns. Instead, two stable gyres occupied these oceans starting in the Aptian until the
Campanian.
Additional key Cretaceous circulation features include:
- A clockwise Arctic gyre from 125-75 Ma
- A counter-clockwise South Atlantic gyre developing by 90 Ma
- Throughflow between Atlantic and Pacific from 115-65 Ma
- India beginning to block the Equatorial Current by 70 Ma
In summary, Cretaceous circulation patterns followed basic wind-driven models, but were heavily influenced by
the changing configurations of oceans and continents as Pangea fragmented. Key surface current pathways
opened and closed during the Cretaceous, with stable gyres developing in ocean basins as they widened. The
breakup of Gondwana resulted in cooling water exchanges between high southern latitudes and the Tethys by the
mid-Cretaceous.
10. Summary of Cretaceous Precipitation Patterns
Cretaceous rainfall regimes shared similarities with the modern climate system, dominated by a tropical everwet
equatorial belt, strong western continent rainfall from temperate westerlies, and eastern continent subtropical
humid belts.
188
The tropical everwet belt persisted throughout the Cretaceous, but at times was segmented by landmasses
straddling the equator. The temperate rainfall belt impinged on western North and South America, Australia, and
southern Africa. Subtropical humidity characterized eastern North America and Asia.
These overall patterns masked important shifts through the Cretaceous driven by continental motions across
climate zones and the opening of new oceanic gateways (e.g., connection between the South and Central Atlantic).
During the Early Cretaceous, prior to the opening of the South Atlantic, a sharp, Sahel-like boundary developed
across northern South America and north-central Africa.
As South Atlantic rifting progressed, southwestern Africa gained humidity from Atlantic sources during the Albian
and Cenomanian. Northern continentality decreased with extensive flooding, enabling a through-going everwet
tropical belt by the Cenomanian at 95 Ma. During the early and mid-Cretaceous, northeastern North America
became wetter as it moved northward out of the arid subtropics and into more humid warm, temperate latitudes.
The increase in rainfall was also, in part, due to the widening of the Central Atlantic and the increased transport of
moisture into the North American continent.
India’s initial location was in the southern arid belt. During the Early Cretaceous it moved southward into the
temperate rainy belt. India reversed direction in the mid-Cretaceous and by the Late Cretaceous its northward
movement brought it into the region of subtropical monsoons. Australia similarly became wetter during its south
to north traverse, ending the Cretaceous well within the temperate rainy belt.
In general, Cretaceous precipitation patterns recorded in the geological proxies are in good agreement with
computer simulations of climate. Some mismatches occurred however, with models underestimating subtropical
rainfall in North Africa and regions of northern South America.
While robust similarities remained with modern rainfall zones, the Cretaceous saw dramatic continental-scale
shifts in moisture transport as landmasses traveled across global climate belts in their tectonic voyages.
E. Discussion
Map Projection. The reader may have noticed that most of the maps in this chapter use the same map projection,
the equirectangular map projection. The Equirectangular map projection is the simplest map projection; latitude
and longitude are Cartesian coordinates, where latitude = y, and longitude = x. For that reason, paleo-latitude and
paleo-longitude can be read directly from the map. The main disadvantage of the equirectangular project is that it
189
greatly distorts and exaggerates features at high latitudes. For example, the polar ice caps shown in Figures 49 – 52
(145 Ma – 130M) appear relatively larger than they actually are (Figure 116).
The most significant advantage of the equirectangular projection is that an equirectangular image can be easily
mapped onto the surface of a sphere. This can be done using standard geographic information software (GIS)
software (e.g., ArcGIS™ or QGIS) or with the mapping software, G.Projector v. 3.2.3, developed by Robert B.
Schmunk, NASA Goddard Institute for Space Studies. G.Projector allows users to make maps in over 130 map
projections including: Mercator, Mollweide, Orthographic, Robinson, and various polar projections. G.Projector
was used to produce Figure 116 as well as the maps in the Supplemental Materials. It can be freely downloaded at
https://www.giss.nasa.gov/tools/gprojector/.
190
Figure 116. Earliest Cretaceous (145 Ma) Polar Orthographic projection illustrating actual size of south polar icecap
(~ 4 million km2).
191
Supplemental Materials. The reader, no doubt, has noticed the many references to the Supplemental Materials
that are sprinkled throughout the chapter. In many the text in this essay is simply a brief introduction to the wealth
of cartographic and digital data that we are providing in the Supplemental Material. One of the important goals of
this essay is to provide the original digital data that was used to produce the various plate tectonic,
paleogeographic, and paleoclimatic maps in this chapter. We also provide ancillary materials, such as animations
and supplemental reports that may be of interest to the reader. Because of the limitations of the equirectangular
projection, we have also provided versions of the principal figures in several complementary map projections (i.e.,
Mollweide, North and South Polar Orthographic, and Robinson). For easy reference, the Supplemental Materials
that are mentioned throughout this chapter have been compiled into a single, comprehensive table (Table 14).
The Supplemental Materials are available in two on-line archives. One archive,
https://doi.org/10.5281/zenodo.10659104, deals exclusively with Cretaceous material. The other archive,
https://doi.org/10.5281/zenodo.10659112, contains material that is Phanerozoic in scope.
F. Conclusions
1. Some Conclusions about the Cretaceous
We can draw several important conclusions about the Cretaceous from the synthesis presented in this paper. It is
clear that the tectonics, geography, and climate of the Cretaceous was a very different from the modern world and
very different from the last 35 million years of the Cenozoic.
During the Late Cenozoic, the continents were separated by wide ocean basins. In contrast, at the start of the
Cretaceous the supercontinent of Pangea had just begun to break apart and only a few small ocean basins
separated Laurasia (Eurasia and North America), West Gondwana (South America, Africa, and Arabia), and East
Gondwana (Madagascar, India, Antarctica, Australia, and Zealandia).
The Late Cenozoic was a time characterized by multiple continent-continent collisions and extensive mountain
building. In contrast there were no significant continent-continent collisions during the Cretaceous and the
continents were low-lying and easily flooded.
Both the late Cenozoic and the Cretaceous had large ocean basins (Pacific-Panthallasic) covering more than half of
the surface area of the globe. The Panthallasic Ocean was significantly larger (~20%) than the Pacific Ocean and
contained numerous continental and oceanic island arcs (Wrangellia, North Slope, Okhotsk-Bering Sea, Oman,
Ladakh-Gangesi, Philippines, and Zealandia). Several of these island arcs collided with adjacent continents during
the Cretaceous.
192
The transition from a Pangea-like to a more dispersed continental configuration had important effects on global
sea level. During the Early Cretaceous, the length of mid-ocean ridges and subduction zones was approximately
equal. As a consequence, ridges had to spread relatively quickly to keep up with the amount of oceanic lithosphere
removed by subduction. Over time, these high rates of spreading produced large areas of young, high-standing
ocean floor which displaced water out of the ocean basins and onto the continents. The resulting continental
flooding was amplified by the fact that as the continents rifted apart, the new continental rifts were transformed
into young ocean basins. The oceanic lithosphere in these young ocean basins was thermally elevated, which
further boosted sea level. Sea level was highest during the mid-Cretaceous (90 Ma – 80 Ma), with a subsidiary peak
~ 120 million years ago (early Aptian).
We can now conclude that, overall, the Cretaceous was much warmer than the Late Cenozoic (> 10˚C warmer). The
Cenomanian-Turonian Thermal Maximum (95 – 80 Ma) achieved temperatures (28˚C) that were only exceeded by
the Permo-Triassic Thermal Maximum (32˚). These very warm times caused several episodes of oceanic anoxia
(OAEs) and high temperatures in equatorial regions sometimes made terrestrial and shallow marine ecosystems
uninhabitable (temperatures > 40˚C). This is unlike anything we have seen in the last 35 million years and may
presage the eventual results of man-made global warming.
There were, however, times during the Early Cretaceous when the climate was similar to the late Cenozoic. During
parts of the Early Cretaceous, there was a small, permanent ice cap on Antarctica. Though half the size of
Antarctica (~ 4 million km2), the waxing and waning of this ice cap contributed to changes in global sea level (~ 30
m). During the Early Cretaceous, there were cool conditions in both hemispheres at temperate latitudes (>55˚ N
and S) as evidenced by dropstones, glendonites, rare tillites, and temperate forest biomes.
2. How well do we know the Cretaceous?
It depends. If we subdivide our knowledge of the Cretaceous into five categories: time control, life, climate,
tectonics, and geography, we can give very high marks to the paleontologists and stratigraphers who, during the
last 200 years, have assembled a detailed description of the evolution of life and the correlation of Cretaceous rock
units (Gradstein et al., 2020). The grades we give to the climatologists, tectonicists, and paleogeographers are
more mixed.
Supercomputer models of paleoclimate are now highly precise and give excellent results for the modern world
and the recent global ice age, but the climate simulations still seem to struggle when modelling ancient hothouse
worlds. There has been a long-running debate as to whether the Cretaceous was a hothouse world with warm
polar regions or, as early climate simulations have suggested, was characterized by cool temperate regions and
undetected polar ice caps. This controversy has largely been resolved by incremental improvements to the climate
193
models that has resulted in warmer polar regions. With more lithologic indicators of climate, we can now make
more nuanced interpretations of changing Cretaceous climates. The Cretaceous was neither a hothouse nor a mild
icehouse, but rather a combination of both: cool early on (145 Ma – 130 Ma), warming (125 Ma – 105 Ma), then
very warm (100 Ma – 85 Ma), mild towards the end (80 Ma – 67 Ma), with a chilling punctuation mark at the
Cretaceous/Paleogene boundary (66 Ma) as a consequence of the end-Cretaceous Impact Winter.
Our confidence in the plate tectonic story for the Cretaceous is very mixed. Overall, the paleomagnetic record for
the Cretaceous is good and we can place the continents within 5˚ of their correct latitudes. The plate tectonic
development of the intra-Pangean ocean basins (Atlantic and Indian) is well-documented and precise. The plate
tectonic history of the Tethys Ocean, however, is little more than an educated guess. Our understanding of the
plate tectonic development of the vast extra-Pangean, Pacific-Panthallasic ocean basin degrades rapidly as we go
back through time. We can reimagine subducted ocean floor with some confidence back to the mid-Cretaceous (84
Ma; chron 34); prior to that time, all plate tectonic reconstructions in the vast Panthallasic ocean are largely
guesswork. Once we get back into the pre-Cretaceous, though paleomagnetic data constrains changing orientation
of the continents, we have no clue what was happening in the vast Panthallasic Ocean.
The paleogeographic maps presented here, taken as a whole, are the most complete and the best synthesis to
date. As a paleogeographer (CRS), I wish I could say that they are the best that we can do. Unfortunately, they are
not. The paleogeographic maps presented here are the work of one man. They are a rough, initial synthesis, based
on a limited amount of geological data and a lot of tectonic inference. But there is hope. In this age of big data and
AI assistants, I am confident that, in the future, we will do much better.
This is important that we improve our paleogeographic mapping because paleogeography, i.e. the ancient
topography and bathymetry, is a key input to paleoenvironmental reconstructions and paleoclimate models. The
paleotemperature, paleorainfall, paleo-Köppen, paleo-oceanic circulation, and paleo-river simulations presented in
this paper all depend on accurate estimates of paleobathymetry and paleotopography.
194
Figure 117. Connections within the Earth System
3. Future Earth System History Research and the Use of Artificial Intelligence (AI)
Since the time of Nicholas Steno, more than 350 years ago, field geologists, paleontologists, stratigraphers, and
geophysicists have fastidiously and painstakingly described and mapped the surficial and buried layers of the
Earth’s crust, revealing its detailed and exquisite history. This story emerged slowly at first, then accelerated as
Earth Historians were able to calibrate their correlations using absolute radiometric age-dating. Using this robust
chronometric framework, we are beginning to tell the exciting story of the Earth during the Cretaceous, in all of its
interwoven tectonic, geographic, oceanographic, atmospheric, geochemical, climatic, and biologic complexity
(Figure 117).
195
During the last 30 years, energy companies and national geological surveys have compiled 3D exploration
databases and have built 3D stratigraphic models covering the globe. An immense amount of high-resolution,
digital stratigraphic data are now on hand. So what does the future hold?
The task before us is two-fold: first, a new Earth System History “Library of Alexandria” must be built from this
great volume of digital Earth System History data to preserve it for future generations and, more importantly, to
create a comprehensive, coherent, and authoritative digital data source for the emerging artificial intelligence
systems (e.g., GPT5). The Earth Sciences, in particular geochronology, geochemistry, paleoclimate, tectonics,
stratigraphy, and paleontology, should begin to build a large language model (LLM) dedicated to helping earth
scientists unravel the long and complex history of the Earth.
Figure 118. Earth System History Machine
The second task, though daunting, is to use this Earth System History database and LLM, to build an “Earth System
History Machine” (Figure 118). This Earth System History Machine, with the help of AI, will dynamically simulate
the plate tectonic, paleogeographic, sedimentologic, paleoceanographic, geochemical, paleoclimatic evolution of
the Earth from its beginnings, to the present, and forward into the future. Numerous researchers, around the
globe, are now actively working towards this lofty goal They include researchers and students at the PALEOMAP
Project (Northwestern University), EarthByte (University of Sydney), CEED (Norway), University of Geneva,
196
University of Chicago, Leeds University, Chronosphere Project (Erlangen), University of Bristol, Texas A&M
University, NMNH, DDE Project, Purdue University, Chinese Academy of Sciences, Peking University, University of
Geosciences (Wuhan), Nanjing University, Paleobiology Database, Macrostrat (University of Wisconsin), Utrecht
University, University of Western Australia, CNRS, ETH, University of Lisbon, and many others. This research should
be encouraged and supported because understanding how the Earth System has evolved in the past will give us
the best opportunity to understand the present-day workings of the Earth System as well as tackle future
challenges. As T.S. Eliot wrote,
“We shall not cease from exploration
And the end of all our exploring
Will be to arrive where we started
And know the place for the first time.”
Little Gidding, Four Quartets (1943)
G. Acknowledgements
CRS received support to produce the maps in the Paleogeographical Atlas from the PALEOMAP Project industrial
consortium (2003 – 2013) and would like to thank innumerable friends, mentors, and enthusiastic supporters
through the years, in particular A.M. Ziegler, W.S. McKerrow, R. van der Voo, S. Snelson, J. Westrich, P. Unternehr,
C.P. Summerhayes, and A.J. Boucot. ATK assembled the on-line Supplemental Materials and was supported by the
Deutsche Forschungsgemeinschaft (Ko 5382/2-1). Alexander Tenesaca helped with the preparation of some of the
figures. Malcolm Hart, Michael Wagereich, and Irek Walaszczyk provided valuable critiques and corrections.
Special thanks to Phyllis Richmond for correcting grammar, clarifying prose, and corralling verbage. This chapter is
dedicated to the life-long work and research in the Cretaceous period of Erle G. Kauffman, Roger L. Larson, Bruce
W. Sellwood, and Robert A. Spicer.
197
H. References Cited
Advokaat, E.L., van Hinsbergen, D.J.J., 2023. Finding Argoland: Reconstructing a microcontinental
archipelago from the SE Asian accretionary orogen, Gondwana Research (2023), doi: https://
doi.org/10.1016/j.gr.2023.10.005
Ahlberg, A., Herman, A.B., Raikevich, M., Rees, A., and Spicer, R., 2002. Enigmatic Late Cretaceous high
paleo-latitude lonestones in Chukotka, northeasternmost Asia, GFF, v. 124, p. 197-199.
Algeo TJ, Seslavinsky KB. 1995. Reconstructing eustasy and epeirogenic trends from Paleozoic continental
flooding records. In: ed. Haq, BU, Sequence Stratigraphy and Depositional Response to Eustatic, Tectonic and Climate
Forcing, pp. 209-246, Netherlands:Kluwer Academic Publishers
Ali, J.R., and Kraus, D.W., 2011. Late Cretaceous bioconnections between Indo-Madagascar and
Antarctica: refutation of the Gunnerus Ridge causeway hypothesis, Journal of Biogeography, 38:1855-1872.
Allaby, M., and Lovelock, J., 1985. The Great Extinction, Paladin, Granado Publishing, London, 192 p.
Alley, N.F., and Frakes, L.A., 2003. First known Cretaceous glaciation: Livingstone Tillite Member of the
Cadna-owie formation, Australian Journal of Earth Sciences, v. 50, p. 139-144.
Alroy J, Aberhan M, Bottjer DJ, Foote M, Fürsich FT, Harries PJ, Hendy AJW, Holland SM, Ivany LC, Kiessling
W, Kosnik MA, Marshall CR, McGowan AJ, Miller AI, Olszewski TD, Patzkowsky ME, Peters SE, Villier L, Wagner PJ,
Bonuso N, Borkow PS, Brenneis B, Clapham ME, Fall LM, Ferguson CA, V.Hanson VL, Krug AZ, Layou KM, Leckey EH,
Nürnberg S, Powers CM, Sessa A, Simpson C, Tomašových A, Visaggi CC. 2008. Phanerozoic trends in the global
diversity of marine invertebrates. Science 321:97-100.
Alsenz, H., Regnery, J., Ashckenazi-Polivoda, S., Meilijson, A., Ron-Yankovich, L., Abramovich, S., Illner, P.,
Almogi-Labin, A., Feinstein, S., Berner, Z., Püttmann, W., 2013. Sea surface temperature record of a Late
Cretaceous tropical southern Tethys upwelling system. Palaeogeogr. Palaeoclimatol. Palaeoecol. 392, 350–358.
Alvarez, L.W., and Alvarez, W., Asaro, F., and Michel, H.V., 1980. Extraterrestrial cause for the Cretaceous-
Tertiary extinction, Science, 208:1095-1107, doi: 10.1126/science.208.4448.1095
Antretter, M., 2001, PhD thesis at LMU Munich: https://edoc.ub.uni-
muenchen.de/427/1/Antretter_Maria.pdf.
Archibald, J.D., 1996. Dinosaur Extinction and the End of an Era: What the Fossils Say, Columbia University
Press, New York, 237 p.
Artemieva, I., 2006. Global 1˚ x 1˚ thermal model TC1 for the continental lithosphere: Implications for
lithosphere secular evolution, Tectonophysics, 416:245-277.
Arthur, M.A., and Sageman, B.B., 1994. Marine black shales: Depositional mechanisms and environments
of Ancient Deposits, Annual Review of Earth and Planetary Sciences, 22: 499-552.
Audley-Charles, M., 1983. Reconstruction of eastern Gondwanaland, Nature, 306:48-50,
doi.org/10.1038/306048a0
Audley-Charles, M., 1984. Cold Gondwana, warm Tethys, and Tibetan Lhasa Block, Nature 310:165,
doi:10.1038/310165.0
Audley-Charles, M., 1988. Evolution of the southern margin of Tethys (North Australian region) from early
Permian to late Cretaceous, in M. Audley-Charles and A. Hallam (editors), Gondwana and Tethys, Geological
Society of London, Special Publication, 37: 79-100.
Baatsen, M., van Hinsbergen, D. J. J., von der Heydt, A. S., Dijkstra, H. A., Sluijs, A., Abels, H. A., and Bijl, K.,
2015. A generalized approach for reconstructing geographical boundary conditions for palaeoclimate modeling.
Climate of the Past Discussions, 11, 4917-4942.
Bao, X.J., and Hu, Y.Y., 2024. Quantitative estimation of global mean precipitation throughotu the
Phanerozoic era, Science China Earth Sciences, 67, pp. 1-9,https://doi.org/10.1007/s11430-023-1263-x,
Barker, P., Hill, I., and Osmaston, M., 1981. Back-arc extension in the Scotia Sea, Philosophical
Transactions of the Royal Society of London, 300:249-262, doi: 10.10989/rsta.1981.0063
Barral, A, Gomez, B., Legendre, S., Lécuyer, C., 2017. Evolution of the carbon isotope composition of
atmospheric CO2 throughout the Cretaceous. Palaeogeography, Palaeoclimatology, Palaeoecology,
http://dx.doi.org/10.1016/j.palaeo.2017.01.034
198
Barrera, E. and Johnson, C., 1999. Evolution of the Cretaceous Ocean-Climate System, Geological Society
of America Special Paper 332, Boulder, Colorado, pp. 445.
Barron, E.J., 1983. A Warm, Equable Cretaceous: the Nature of the Problem, Earth-Sc ience Reviews,
19:305-335.
Barron, E.J., 1984. Climatic implicationsof the variable obliquity explanation of Cretaceous-Paleogene
high-latitude floras, Geology, 12:595-598
Barron, E.J., Fawcett, P.J., and Peterson, W.H., 1995. A “simulation” of mid-Cretaceous climate,
Paleoceanography, v. 10, no. 5, pp. 953-962.
Barron, E. J., Peterson, W.H., et al., 1993. Past climate and the role of ocean heat transport: Model
simulations for the Cretaceous. Paleoceanography 8(6): 785-798.
Barron, E.J., Thompson, S.L., and Schneider, S.H., 1981. An Ice-free Cretaceous? Rsults from Climate
Model Simulations, Science, 212:501-508.
Barron, E. J. and Washington, W.M., 1982a. Atmosphereic circulation during warm geologic periods: Is the
equator-to-pole surface-temperature gradient the controlling factor?, Geology, 10:633-636.
Barron, E. J. and Washington, W.M., 1982b "Cretaceous climate: A comparison of atmospheric simulations
with the geologic record." Palaeogeography, Palaeoclimatology, Palaeoecology 40(1-3): 103-133.
Barron, E. J. and Washington, W.M., 1984. The Role of Geographic VAriables in Explaining Paleoclimates:
Results from Cretaceous Climate Model Sensitivity Studies, Journal of Geophysical Research, v. 89, no. D1, pp.
1267-1279.
Barron, E. J. and Washington, W.M., 1985. Warm Cretaceous climates - High atmospheric CO2 as a
plausible mechanism. The Carbon Cycle and Atmospheric CO2 : Natural Variations Archean to Present. Washington
D.C., American Geophysical Union: 546-553.
Bakker, R.T., 1986. The Dinosaur Heresies: New Theories Unlocking the Mystery of the Dinosaurs and
Their Extinction, Zebra Books, Kensington Publishing Corp., New York, New York, 482 p.
Bakker, R.T., 1995. Raptor Red, Bantam Books, New York, New York, 247 p.
Bebout, G.E., Scholl, D.W., Kirby, S.H., Platt, J.P., 1996. (editors), Subduction: Top to Bottom, American
Geophysical Union, Geophysical Monograph 96, Washington, D.C., 384 p.
Beccaluva, L., Bianchini, G., and Wilson, M., 2011. Volcanism and Evolution of the African Lithosphere,
Geological Society of America Special Paper 478, Goelogical Society of America, Boulder, CO, 331 p.
Beerling, D.J., and Woodward, F.I., 2001. (editors) Chapter 7, The Cretaceous, in Vegetation and the
Terrestrial Carbon Cycle: Modelling the first 400 million years, Cambridge University Press, Cambridge, pp. 183-
237.
Beerling, D.J., Fox, A., Stevenson, D.S., Valdes, P.J., 2011. Enhanced chemistry-climate feedbacks in past
greenhouse worlds. P Natl Acad Sci USA 108, 9770-9775.
Benton, M.J., 2003. When Life Nearly Died: The Greatest Mass Extinction of All Time, Thames and Hudson,
London, 336 p.
Berggren, W.A., and Van Couvering, J.A., 1984. (editors), Catastrophes in Earth History: The New
Uniformitarianism, Princeton University Press, Princeton, NJ, 464 p.
Bernardi, M., Gianolla, P., Petti, F.M., Mietto, P., and Benton, M.J., 2018. Dinosaur diversification linked
with the Carnian Pluvial Episode, Nature Communications, v. 9 :1499, doi: 10.1038/s41467-018-03996-1.
Berner, R. A., and Kothavala, Z., 2001, Geocarb III: A Revised Model of Atmospheric CO2 Over Phanerozoic
Time, American Journal of Science, vol. 301, p. 182-204
Berner R. A., 1994. GEOCARB II: A revised model of atmospheric CO2 over Phanerozoic time. American
Journal Science 294(1):56–91.
Berner, R.A., 2004. The Phanerozoic Carbon Cycle: CO2 and O2. Oxford University Press, New York.
Berner, R.A., Lasaga, A.C., and Garrels, R.M. (1983). The carbonate silicate geochemical cycle and its effect
on atmospheric carbon dioxide over the past 100 million years. American Journal Science, 283, 641-683.
Bice, K.L., and Norris, 2002, Possible atmospheric CO2 extremes of the middle Cretaceous (late Albian to
Turonian), Paleoceanography, 17:1-17.
Bice, K.L., Bralower, T.J., Duncan, R.A., Huber, B.T., Leckie, R.M., and Sageman, B.B., 2002, Cretaceous
climate-ocean dynamics: future directions for IODP, A workshop sponsored by JOI/USSSP and NSF, The Nature Place,
Florissant, Colorado, July 14-17, 2002. (www.usssp-iodp.org/Science_Support/Workshops/past_workshops.html)
199
Bice, K.L., Huber, B.T., and Norris, R.D., 2003, Extreme polar warmth during the Cretaceous greenhouse?
Paradox of the late Turonian D 18O record at Deep Sea Drilling Project Site, Paleoceanography, 18:1013-1029.
Biddle, K.T., 1991. (editor), Active Margin Basins, American Association of Petroleum Geologists, Memoir
52, Tulsa, OK, 324 p.
Bijl, P.K., Schouten, S., Sluijs, A., Reichart, G.J., Zachos, J.C., and Brinkhuis, H., 2009, Early palaeogene
temperature evolution of the southwest Pacific Ocean: Nature, v. 461, p. 776–779, doi:10.1038/nature0839
Bird, J.M., and Dewey, J.F., 1970. Lithosphere plate-continental margin tectonics and the evolution of the
Appalachian orogen, Geological Society of America Bulletin, 81:1031-1060.
Blakey, R. C., 2002. Global Paleogeography, Rectilinear Projection, Colorado Plateau Geosystems, Inc.,
Flagstaff, AZ,, DVD
Blakey, R. C., 2008. Gondwana paleogeography from assembly to breakup - A 500 m.y. odyssey,
Geological Society of America Special Papers, v. 441, p. 1-28
Blakey, R. C., 2011. Paleogeography of Europe, Colorado Plateau Geosystems, Inc., Flagstaff, AZ,, DVD
Blakey, R. C., 2013. Key Time Slices of North American Geologic History, Colorado Plateau Geosystems,
Inc., Flagstaff, AZ, DVD
Bohor, B.F., Triplehorn, D.M., Nichols, D.J., and Millard, H.T.,Jr., 1987. Dinosaurs, spherules, and the
"majic" layer: A new K-T boundary clay site in Wyoming, Geology,15:896-899.
Bond, D.P.G., and Grasby, S.E., 2017. On the causes of mass extinctions, Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 478, p. 3–29, doi.org/10.1016/j.palaeo.2016.11.005
Bond, D.P.G., and Wignall, P.B.,2014. Large igneous provinces and mass extinction: An update, in
Volcanism, Impacts, and Mass Extinctions: Causes and Effects, G. Keller and A. Kerr, (editors), Geological Society of
America, Special Paper 505, pp. 29 -55, doi: 10:1130/2014.2505(02)
Boschman, LM, 2022. Andean mountain building since the Late Cretaceous: A paleoelevation
reconstruction, Earth-Science Reviews v. 220, 103640
Boucot, A. J., Chen Xu, and Scotese, C. R., 2013. Phanerozoic Paleoclimate: An Atlas of Lithologic
Indicators of Climate, SEPM Concepts in Sedimentology and Paleontology, (Print-on-Demand Version), No. 11, 478
p, ISBN 978-1-56576-289-3, October 2013, Society for Sedimentary Geology, Tulsa, OK.
Boutelier, D., and Cruden, A., 2013. Slab rollback rate and trench curvature controlled by arc deformation,
Geology, v. 41, no. 8, pp. 911-914. Doi: 10.1130/G34338.1
Bozhko, N. A., & Khain, V. E., 1987. Gondwana Paleotectonic Maps, Ministry of Higher and Secondary
Special education of the U.S.S.R., and Ministry of Geology, U.S.S.R., produced by the Geological Complex of Central
Regions of the U.S.S.R., 30 pp.
Bradley, D.C., 2008. Passive margins through Earth history, Earth-Science Reviews, 91:1-26,
doi:10.1016/jearsirev2008.08.001
Bradley, D.C., 2011. Secular trends in the geologic record and the supercontinent cycle, Earth-Science
Reviews, 1008 (1-2), pp. 16-33.
Brady, E.C., DeConto, R., Thompson, S.L., 1998, Deepwater formation and poleward ocean heat transport
in the warm climate extreme of the Cretaceous (80 Ma), Geophysical Research Letters, 25: 4205-4208.
Brass, G.W., Southam, J.R., Peterson, W.H., 1982. Warm Saline Bottom Water in the Ancient Ocean.
Nature 296, 620-623.
Brassell, S.C., 2009. Steryl ethers in a Valanginian claystone: molecular evidence for cooler waters in the
central Pacific during the Early Cretaceous? Palaeogeogr. Palaeoclimatol. Palaeoecol. 282, 45–57.
Briggs, J.C., 1987. Biogeography and Plate Tectonics, Developments in Palaeontology and Stratigraphy, no.
10, Elsevier, Amsterdam, 204 p.
Brown, J.H., and Lomolino,M.V., 1998. Biogeography, second edition, Sinauer Associates, Sunderland, MA,
692 p.
Brusatte, S., 2018. The Rise and Fall of the Dinosaurs: New History of Their Lost World, William MOrrow,
an impront of NarperCollins Publishers, New York, New York, 404 p.
Buffan, L., Jones, L.A., Domeier, Scotese, C.R., Zahirovic, S., and Varela, S., 2023. Mind the Uncertainty:
Global plate model choice impacts deep-time palaeobiological studies, Methods in Ecology and Evolution, Wiley
Online Library. DOI: 10.1111/2041-210X.14204
Buffler, R.T., and Sawyer, D.S., 1983. Distribution of crust and early history, Gulf of Mexico basin, Gulf
Coast Association, Geological Society Transactions, 35:334-344.
200
Bullard, E., Everett, J.E., and Smith, A.G., 1965. The Fit of the Continents Around the Atlantic, Philosophical
Transactions of the Royal Society pf London, Series A., Mathematical and Physical Sciences, October 28, 1965, No.
1088, A Symposium on Continental Drift, pp. 41-51.
Burchfiel, B.C., Cowan, D.S., and Davis, G.A., 1992. Tectonic overview of the Cordilleran orogen in the
western United States, in The Cordilleran Orogen: Conterminous U.S., B.C. Burchfiel, P.W. Lipman, and M.L. Zoback
(eds.), Decade of North American Geology, The Geology of North America, v. G-3, Geological Society of America,
Boulder, Colorado, p. 407-480.
Burchfiel, B.C., Lipman. P.W., and Zoback, M.L., 1992. The Cordilleran Orogen: Conterminous U.S., The
Geology of North America, Volume G-3, Geological Society of America, Boulder, CO, 724 p.
Burg, J.P., 2011. The Asia–Kohistan–India Collision: Review and Discussion. In: Arc-Continent Collision.
Frontiers in Earth Sciences. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-540-88558-0_10
Burg, J.P., and Ford, M., 1997. (editors), Orogeny Through Time, Geological Society Special Publication,
No. 121, Geological Society, London, 270 p.
Burgener, L., Hyland, E., Reich, B.J., and Scotese, C., 2023. Cretaceous climates: Mapping paleo-Löppen
climatic zones using a statistical analysis of lithologic, paleontologic, and geochemical proxies, Palaeogeography,
Paleoclimatology, and Palaeoecology, 613, 111373, https://doi.org/10.1016/j.palaeo.2022.111373
Burk, C.A., and Drake, C.L., 1974. (editors), The Geology of Continental Margins, Springer-Verlag, New
York, 1009 p.
Burrett, C., 1974. Plate tectonics and the fusion of Asia, Earth and Planetary Science Letters, 21: 181-189.
Burrett, C., Zaw, K., Meffre, S., Lai, C.K., Somboon, K, and Pol,C., 2014. The configuration of Greater
Gondwana – Evidence from LA ICPMS, U-Pb geochronology of detrital zircons from the Palaeozoic and Mesozoic of
Southeast Asia and China, Gondwana Research, 26: 31-51, dx.doi.org/10.1016/j.gr.2013.05.020
Burke, K., and Dewey, J.F., 1975. The Wilson Cycle. In: Geological Society of America, Northeastern
Section, 10th Annual Meeting, Syracuse, NY, Abstracts with Programs, Boulder, CO, 48.
Bush, A.B.G., and Philander, S.G.H., 1997. The late Cretaceous: Simulation with a coupled atmosphere-
ocean general circulation model, Paleoceanography, v. 12, no. 3, pp. 495-516.
Butler, R.F., 1998. Paleomagnetism: Magnetic Domains to Geologic Terranes,
https://www.geo.arizona.edu/Paleomag/tocpref.pdf
Cande, S.C. and Mutter, J.C., 1982. A revised identification of the oldest sea-floor spreading anomalies
between Australia and Antarctica, Earth and Planet. Sci. Letters, 58:151-160.
Cande, S.C., and Stegman, D.R., 2011. Indian and African plate motions driven by the push force of the
Reunion plume head, Nature, v. 475, 47-52 p. et al., (2015)
Cande, S., LaBrecque, J.L., and Haxby, W.B., 1988. Plate kinematics of the South Atlantic: Chron 34 to
present, J. Geophys. Res., 93 (B11): 13479-13492
Cande, S.C., LaBrecque, J.L., Larson, R.L., Pitman, W.C., III, Golovchenko, X., and Haxby, W.F., 1989,
Magnetic lineations of the world's ocean basins: Tulsa, Oklahoma, American Association of Petroleum Geologists,
Tulsa, OK, scale 1:27,400,000 at the equator.
Cantrill, D.J., and Poole, I., 2012. The Vegetation of Antarctica Through Geological Time, Cambridge
University Press, Cambridge, England, 480 pp.
Cao, W., C.-T. A. Lee, and J. S. Lackey (2017). Episodic nature of continental arc activity since 750 Ma: A
global compilation. Earth and Planetary Science Letters, 461, 85-95.
Cao, W., Williams, S., Flament, N., Zahirovic, S., Scotese, C., and Müller, R. D., 2019. Palaeolatitudinal
distribution of lithologic indicators of climate in a palaeogeographic framework. Geological Magazine, 156(2), 331–
354. https://doi.org/10.1017/S0016756818000110
Carlisle, D.B., 1995. Dinoasaurs, Diamonds. and Things from Outer Space: The Great Extinction, Stanford
University Press, Stanford, CA., 241 p.
Casey, R,m and Rawson, P.F., 1973. The Boreal Lower Cretaceous, Geological Journal Special Issue, No. 5,
Seel House Press, Liverpool, 448 p.
Catalan, J.R.M., Hatcher, R.D., Jr., Arenas, R., a Garcia, F.D., 2002. Variscan-Appalachian dynamics: The
building of the late Paleozoic basement, Geological Society of America, Special Paper 364, Boulder, CO, 305 p.
Chamberlin, T.C., 1898. The ulterior basis of time divisions and the classification of geological history,
Journal of Geology, vol. 6, pp. 449-462.
201
Chamberlin, T.C., 1906. On a possible reversal of deep-sea circulation and its influence on geologic
climates, Proceedings of the American Philosophical Society, v. 45, n. 182, p. 33-43.
Chamberlin, T.C., 1909. Diastrophism as the ultimate basis for correlation. Journal of Geology, vol. 17, pp.
685-693.
Chatterjee, S., 1997. The Rise of Birds: 225 Million Years of Evolution, Johns Hopkins University Press,
Baltimore, Maryland, 312 p.
Chatterjee, S., and Scotese, C.R.,1999. The breakup of Gondwana and the evolution and biogeography of
the Indian plate. Proceedings of Indian National Science Academy 65A:397-425.
Chatterjee, S., Guven, N., Yoshinobu, A., and Donofrio, R., 2006. Shiva structure: A possible KT boundary
impact cater on the western shelf of India, Museum of Texas Tech University, Special Publication, 40 pp.
Chen, B., Joachimski, M. M., Shen, S.-Z., Lambert, L. L., Lai, X.-L., Wang, X.-D., Chen, J., and Yuan, D.-X.,
2013, Permian ice volume and palaeoclimate history: Oxygen isotope proxies revisited, Gondwana Research, v. 24,
no. 1, p. 77-89.
Chenet, A. L., Fluteau, F., Courtillot, V., Gerard, M., and Subbarao, K. V., 2008. Determination of rapid
Deccan eruptions across the Cretaceous- Tertiary boundary using paleomagnetic secular variation: Results from a
1200-m-thick section in the Mahabaleshwar escarpment. Journal of Geophysical Research, 113, B04101.
https://doi.org/10.1029/2006JB004635
Chiarenza, A.A., Mannion, P.D., Farnsworth, A., and Carrano, M.T., 2022. Climatic constraints on the
history of Mesozoic dinosaurs, Current Biology, 32: 570-585, doi.org/10/1016/j.cub.2021.11.061
Chumakov, N.M., 1997. The Warm Biosphere of the Cretaceous Period, Journal of Journals, Review of
Global Scientific Achievments, International Academic Publishing Cp., NAUKA, v.1, pp.69-74.
Chumakov, N. M., Zharkov, M. A., Herman, A. B., Doludenko, M. P., Kalandadze, N. N., Lebedev, E. L.,
Ponomarenko, A. G., and Rautian, A. S., 1995. Climatic Belts of the Mid-Cretaceous Time. Stratigraphy and
Geological Correlation, 3(3), 241–260.
Clapham, M.E., and Renne, P.R., 2019. Flood Basalts and Mass Extinctions. Annual Review of Earth
Planetary Sciences, v. 47, p. 275–303, doi: 10.1146/annurev-earth-053018-060136
Clark, P. U. et al., 2016: Consequences of twenty-first-century policy for multi-millennial climate and sea-
level change. Nature Climate Change, 6, 360–369, doi:10.1038/nclimate2923.
Clennett, E.J., Holt, A.F., Tetley, M.G., Becker, T.W., and Faccenna, C., 2023. Assessing plate reconstruction
models using plate driving force consistency test, Nature, Scientific Reports, 13:10191, 15 p., doi/10.1038/s41598-
023-37117-w
Cobbold, P.R., and Rossello, E.A., 2003. Aptian to recent compressional deformation, foothills of the
Neuquen Basin, Argentina, Marine and Petroleum Geology, 20:429-443.
Cocks, L.R.M., and Torsvik, T.H., 2020. Ordovician palaeogeography and climate change, Gondwana
Research, X:XX, doi.ord/10.1016/j.gr.2020.09.008
Coffin, M.F., and Eldholm, O., 1991. (editors), Large Igneous Provinces, JOIO/USSAC Workshop Report,
University of Texas at Austin, Institute for Geophysics Technical Reports, No. 114, 79 p.
Coffin, M.F., and Eldholm, O., 1994. Large igneous provinces: Crustal structure, dimensions, and external
consequences, Reviews of Geophysics, 32: 1-36.
Coffin, M.F., and Rabinowitz, P.D., 1988. Evolution of the conjugate East African – Madagascan margins and
the western Somali Basin, Geol. Soc. America Special Paper 226, 78 p.
Cohen, K.M., Finney, S.C., Gibbard, P.L., and Fan, J.-X., 2013. The International Chronostratigraphic Chart,
(version 2023/06), Episodes 36: 199-204
Colbert, E.H., 1973. Continental Drift and the Distributions of Fossil Reptiles, in Implications of Continental
Drift to the Earth Sciences, volume 1, D.H. Tarling and S.K. Runcorn (editors), Academic Press, London, pp. 395-412
Coleman, R.G., 1977. Ophiolites: Ancient Oceanic Lithosphere ?, Springer-Verlag, Berlin, 229 p.
Collins, A. S., M. L. Blades, A. S. Merdith, and J. D. Foden (2021). Closure of the Proterozoic Mozambique
Ocean was instigated by a late Tonian plate reorganization event. Communications Earth & Environment, 2(1), 75,
https://doi.org/10.1038/s43247-021-00149-z.
Condie, K.C., 2001. Mantle Plumes and Their Record in Earth History, Cambridge University Press,
Cambridge, 306 p.
Conrad, C.P., and Lithgow-Beretelloni, C., 2006. Influence of continental roots and asthenosphere on
plate-mantle coupling, Geophysical Research Letters, v.33., L053312, doi:10.1029/2005GL025621
202
Conybeare, W.D., And Phillips, W., 1822. Outlines of the Geology of England and Wales, with an
Introduction Compendium of the General Principles that Science, and Comparative Views of the Structure of
Foreign Countries. Part 1., London: William Phillips, 470 p.
Cook, P. J., 1990. Australia: Evolution of a Continent. Bureau of Mineral Resources, BMR, Paleogeographic
Group, Australian Government Publishing Service, Canberra.
Cook, T. D., & Bally, A. W., 1975. Stratigraphic Atlas of North America and Central America, Princeton
University Press, Princeton, New Jersey, 272 pp.
Cope, J. C. W., Ingham, J. K., & Rawson, P. F., 1992. Atlas of Paleogeography and Lithofacies, Geological
Society of London, Memoir 13, 153 pp.
Cordani, U.G., Milani, E.J., Thomaz-Filho, A., and Campos, D.A., 2000. (editors), Tectonic Evolution of
South America, Proceedings of the 31st International Geological Congress, Rio de Janiero, Brazil, August 6 -17,
2000, 856 p.
Corfu, F., Gasser, D., and Chew, D.M., 2014. (editors), New Perspectives on the Caledonides of Scandinavia
and Related Areas. Geological Society Publication, No. 390, Geological Society, London, 718 p.
Courtillot, V., 1999, Evolutionary Catastrophes: The Science of Mass Extinction, Cambridge University Press,
173 pages.
Courtillot, V., and Renne, X., 2003. On the ages of flood basalt events, Comptes Rendus Geosciences,
335:113-140.
Courtillot, V., Feraud, G., Maluski, H., Vandamme, D., Moreau, M.G., and Besse, J., 1988. Deccan flood
basalts and the Cretaceous-Tertiary boundary, Nature,333:843-846.
Courtillot, V., Vandamme, D., Besse, J., Jaeger, J.J., and Javoy, M. 1990. Deccan volcanism at the
Cretaceous/Tertiary boundary; Data and inferences, in V.I. Sharpton and P.D. Ward, (eds.), Global Catastrophes in
Earth History, Geol. Soc. America, Special Paper 247:401-409.
Coward. M.P., Dietrich, D., and Park, R.G., 1989. (editors), Alpine Tectonics, Geological Scoiety Special
Publication, No. 45, Geological Society, Blackwell Scientific Publications, Oxford, 450 p.
Cox, A. and Hart, R.B., 1986. Plate Tectonics: How It Works, Blackwell Scientific Publications, Inc., Palo
Alto, CA., 392 p.
Creer, K. M., 1954. The measurement of geological intervals of time by paleomagnetic methods. Journal
of Geology, 62(2), 234-238.
Crough, T.S., Morgan, W.J., and Hardgraves, R.B., 1980. Kimberlites: Their relation to mantle hot spots,
Earth and Planetary Science Letters, 50:260-274.
Crowley, T.J., Short, D.A., Mengel, J.G., and North, G.R., 1986. Role of seasonality in the evolution of
climate during the last 100 million years, Science, 231:579-584.Cuvier, G., 1831. A Discourse on the Revolutions of
the Surface of the Globe and the Changes thereby produced in the Animal Kingdom, translated from French with
illustrations and a glossary, Carey and Lea, Philadelphia, 260 p.
Cui, Y., and L. R. Kump (2015). Global warming and the end-Permian extinction event: Proxy and modeling
perspectives. Earth-Science Reviews, 149, 5-22.
D’Halloy, J.G.J., 1822. Observations sur un essai de cartes geologiques de la France. des Pays-Bas, et des
contrees voisines. Annales de Mines, 7 :353-376.
D’Lemos, R.S., Strachan, R.A., and Topley, C.G., 1990. (editors) The Cadomian Orogeny, Geological Society
Special Publication No. 51, The Geological Society, London, 423 p.
D’Orbigny, A. 1840. Paleontologie francaise. Terrains cretaces. 1. Cephalpodes. Paris, 622 p.
Dal Corso, J., Benton, M. J., Bernardi, M., Franz, M., Gianolla, P., Hohn, S., and Zhang, Y., 2018a. First
Workshop on the Carnian Pluvial Episode (Late Triassic): a report, Albertiana, v. 44, p. 49-57.
Dal Corso, J., Gianolla, P., Rigo, M., Franceschi, M., Roghi, G., Mietto, P., Manfrin, S., Raucsik, B., Budai, T.,
Jenkyns, H.C., Reymond, C.E., Caggiati, M., Gattolini, G., Breda, A., Merico, A., Preto, N., 2018b. Multiple negative
carbon-isotope excursions during the Carnian Pluvial Episode (Late Triassic), Earth-Science Reviews, v. 185, p. 732–
750, doi:org/10.1016/j.earscirev.2018.07.004
Davies, A. and Simmons, M.D., 2023. Placing constraints on the nature of short-term eustatic curves, Basin
Research, 17 p., doi:10.1111/bre.12832
DeCelles, P.G., Ducea, M.N., Carrapa, B., and Kapp, P.A., 2015. Geodynamics of the Cordilleran Orogenic
System: The Central Andes of Argentina and Northern Chile, Memoir 212, Geological Society of America, Boulder,
CO, 490 p.
203
DeConto, R.M., Hay, W.W., Thompson, S.L., and Bergengren, J., 1999. Late Cretaceous climate and
vegetation interactions: Cold continental interior paradox, in Evolution of the Cretaceous Ocean-Climate System, E.
Barrera and C. Johnson, (editors), Geological Society of America, Special Paper 332:391-406.
de Lurdes Fonseca, M., Scotese, C.R., and Cachao, M., 2019. Late Cretaceous paleobiogeography
of Braarudosphaera bigelowi, Marine Micropaleontology, 152, 21 p., doi.org/10.1016/j.marmicro.2019.03.010
De Lurio, J.L. and Frakes, L.A., 1999. Glendonites as a paleoenvironmental tool: Implications for early
Cretaceous high latitude climates in Australia, Geochimica et Cosmochimica Acta, Vol. 63, No. 7/8, pp.
Demaison, G.J., and Moore, G.T., 1980. Anoxic environments and oil source bed genesis, American
Association of Petroleum Geologists Bulletin, 64:1179-1209.
Dengo, G., and Case, J.E., 1990. The Caribbean Region, Geology of North America, Volume H., Geological
Society of America, Boulder CO, 528 p.
Dera, G., Brigaud, B., Monna, F., Laffont, R., Pucéat, E., Deconinck, J.-F., Pellenard, P., Joachimski, M.M.,
and Durlet, C., 2011. Climatic ups and downs in a disturbed Jurassic world, Geology, v. 39, no. 3., p. 215-218, doi:
10.1130/G31579.1
Dercourt, J., Gaetani, M., Vrielynck, B., Barrier, E., Biju-Duval, B., Brunet, M.F., Cadet, J.P., Crasquin, S., and
Sandulescu, M., 2000. Atlas PeriTethys Palaeogeographic Maps, Commision de la Carte Geologique du Monde,
Paris.
Dercourt, J., Ricou, L. E., and Vrielynick, B.,1993. Atlas Tethys, Palaeoenvironmental Maps. Paris: Gauthier-
Villars, 307 p.
Dercourt, J., Zonenshain, L., Ricou, Kazmin, V. G., Le Pichon, X., Knipper, A. L., Grandjaquet, C.,
Sbortshikov, I. M., Boulin, J., Sorokhtin, O., Geyssant, J., Lepvrier, Biju-Duval, B., Sibuet, J. C., Savostin, L. A.,
Westphal, M., and Lauer, J., 1985. Presentation de 9 cartes paleogeographiques au 1/20.000.000 s’etendant de
l’Atlantique au Pamir pour la periode du Lias a l’Actuel. Bull. Soc. geo. France, 8(5), 637-652.
Dettmann, M.E., 1989. Antarctica: Cretaceous cradle of austral temperate rainforests?, in Origins and
Evolution of the Antarctic Biota, Geological Society of London Special Publication, J.A. Crame (editor), v. 47, p. 89-
105.
Dewey, J.F., and Burke, K.C.A., 1973. Tibetan, Variscan, and Precambrian basement reactivation: products
of continental collision, Journal of Geology, 81:683-692.
Dewey, J.F., Pitman, W.C. III, Ryan, W.B.F., and Bonin, J., 1973. Plate Tectonics and the Evolution of the
Alpine System, Geological Society of America Bulletin, 84:3137-3180.
Dewey, J.F., and Burke, K.C.A., 1974. Hot Spots and Continental Break-up: Implications for Collisional
Orogeny, Geology, 2:57-60.
Dewey, J.F., Gass, I.G., Curry, G.B., Harris, N.B.W., and Sengor, A.M.C., 1991. (editors), Allochthonous
Terranes, Cambridge University Press, Cambridge, 199 p.
Dilek, Y., and Furnes, H., 2014. (editors), Ophiolites. Elements, v. 10, No.2.
Dilek, Y., Moores, E.M., Elthon, D., and Nicolas, A., 2000. Ophiolites and Oceanic Crust: New Insights from
Field Studies and the Ocean Drilling Program, Geological Society of America, Special Paper 349, Boulder, CO, 552 p.
Dilek, Y., and Robinson, P.T., 2003. (editors), Ophiolites in Earth History, Geological Society Special
Publication 218, Geological Society, London, 717 p.
Dilek, Y., and Robinson, P.T., 2009. (editors), Special Issue on Mantle Dynamics and Crust-Mantle
Interactions in Collisional Orogens, Lithos, v. 113, Nos. 1-2, 434 p.
Dingus, L., and Rowe, T., 1998. The Mistaken Extinction: Dinosaur Evolution and the Origin of Birds,
W.H.Freeman, New York, 332 p.
Donnadieu, Y., Puceat, E., Moiroud, M., Guillocheau, F., Deconinck, J.F., 2016. A better-ventilated ocean
triggered by Late Cretaceous changes in continental configuration. Nat Commun 7, 12.
Donovan, S.K., 1989. Mass Extinctions: Process and Evidence, Belhaven Press, London, 266 p.
Duncan, R.A., and Storey, M., 1992. The Life Cycle of Indian Ocean Hotspots, in Duncan, R.A., Rea, D.K.,
Kidd, R.B., von Rad, U., and Weissel, J.K., 1992. Synthesis of Results from Scientific Drilling in the Indian Ocean,
American Geophys. Union, Geophysical Monograph 70: 91 – 103
Dunhill, A.M., Bestwick, J., Narey, H., and Sciberras, J., 2016. Dinosaur biogeographical structure and
Mesozoic continental fragmentation: a network-based approach, Journal of Biogeography, 43:1691-1704.
Edwards, J.D., and Santogrossi, P.A., 1989. (editors), Divergent/Passive Margin Basins, American
Association of Petroleum Geologists, Memoir 48, Tulsa, OK, 252 p.
204
Eglington, B., Evans, D., Pehrsson, S., and Huston, D., 2017. Reconstructing Phanerozoic and Proterozoic
Earth Evolution: building on pre-existing efforts and data, in R. Butler, M. Daly, G. Roberts, J. Turner, T. Watts
(convenors), Plate Tectonics at 50, William Smith Meeting, 2017, 3-5 October, The Geological Society, Burlington
House, London, (abstract book), pp. 28-29.
Eliot, T.S., 1943. Little Gidding, in Four Quartets, lines 239 -242.
Elling, R.P., 2022. Continental Rifting and Precambrian Tectonics: Insights from Gravity Modeling and
Apparent Polar Wander Paths, Ph.D. Thesis, Department of Earth and Planetary Science, Northwestern University,
April 2022, 124 p.
Ellis, R., 2003. Sea Dragons: Predators of the Prehistoric Oceans, University Press of Kansas, Lawrence,
Kansas, 313 p.
Engebretson, D.C., Cox, A., and Gordon, R.G., 1985. Relative Motions between Oceanic and Continental
Plates in the Pacific Basin, Geological Society of America Special Paper 206, Boulder, CO, 59 p.
Erba, E., Bartolini, A., and Larson, R.L., 2004. Valanginian Weissert oceanic anoxic event, Geology
32(2):149-152, doi 10.1130/G20008.1
Erba, E., Duncan, R.A., Bottini, C., Tiraboschi, D., Weissert, H., Jenkyns, H.C., and Malinverno, A., 2015,
Environmental consequences of Ontong Java Plateau and Kerguelen Plateau volcanism, in Neal, C.R., Sager, W.W,
Sano, T., and Erba, E., eds., The Origin, Evolution, and Environmental Impact of Oceanic Large Igneous Provinces:
Geological Society of America Special Paper 511, p. 271–303, doi:10.1130/2015.2511(15).
Ernst, R.E., 2014. Large Igneous Provinces, Cambridge University Press, Cambridge, UK, 653 pp.
Ernst, R.E., and Youbi, N., 2017. How Large Igneous Provinces affect global climate, sometimes cause mass
extinctions, and represent natural markers in the geological record, Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 478, p. 30–52., doi:10.1016/j.palaeo.2017.03.014
Ernst, W.G., 1975. (editor), Subduction Zone Metamorphism, Benchmark Papers in Geology 19, Dowden,
Hutchinson, and Ross, Stroudsburg, PA, 445 p.
Ernst, W.G., 2004. (editor). Serpentine and Serpentinites: Mineralogy, Petrology, Geochemistry, Ecology,
Geophysics, and Tectonics, A Tribute to Robert G. Coleman, International Book Series, volume 8, Geological Society
of America, Bellweather Publishing, Columbia, Md, 606 p.
Erwin, D.W., 1993. The Great Paleozoic Crisis: Life and Death in the Permian, Critical Moments in
Paleobiology and Earth History Series, Columbia University Press, New York, 327 p.
Erwin, D.H., 2006. Extinction: How Life on Earth Nearly Ended 250 Million Years Ago, Princeton University
Press, Princeton, 296 p.
Evans, D., Graham, C., Armour, A., & Bathurst, P., 2003. The Millenium Atlas: Petroleum Geology of the
Central and Northern North Sea, Geological Society of London, 389 pp.
Evans, D.A.D., 2003. True polar wander and supercontinents, Tectonophysics, 362:303-320.
Evans, D.A.D., and Eglington, B.M., 2022. Continuous Quantitative Model of Global Paleogeography
Through 2.0 Billion Years, (abstract), American Geophysical Union, Fall Meeting, Chicago, IL, 12-16 December,
2022, GP43A-02
Evans, D.A.D., and 16 others, 2022, An expanding list of reliable paleomagnetic poles for Precambrian
tectonic reconstructions, in Pesonen, L., Salminen, J., Elming, S-A., Evans, D., and Veikkolainen, T., eds., 2022,
Ancient supercontinents and the paleogeography of Earth: Elsevier, Amsterdam, Netherlands, 646 p.
Everhart, M.J., 2005. Oceans of Kansas: A Natural History of the Western Interior Sea, J.O. Farlow, (editor),
Life of the Past, Indiana University Press, Bloomington, Indiana, 323 p.
Fairhead, J.D., 1988. Mesozoic plate tectonic reconstructions of the South Atlantic Ocean: the role of the
West and Central African Rift System, in Mesozoic and Cenozoic Plate Reconstructions, C.R. Scotese and W.W.
Sager (eds.), Tectonophysics, 155:181-192.
Farnsworth, A., Lunt, D.J., O’Brien, C., Foster, G.L., Inglis, G.N., Markwick, P., Pancost, R.D., and Robinson,
S.A., 2019. Climate sensitivity on geological timescales controlled by non-linear feedbacks and ocean circulation,
Geophysical Research Letters.
Fastovsky, D.E., and Weishampel, D.B., 1996. The Evolution and Extinction of the Dinosaurs, Cambridge
University Press, 461 p.
Fischer, A. G., 1981. Climatic oscillations in the biosphere, in M.H. Nitecki (editor), Biotic Crises in
Ecological and Evolutionary Time, Academic Press, pp. 103-131.
205
Fischer, A. G., 1982. Long-term Climatic Oscillations Recorded in Stratigraphy, in Studies in Geophysics:
Climate in Earth History, W.H. Berger and J.C. Crowell (editors), Chapter 9., National Academy Press, Washington,
D.C., pp. 97-104.
Fischer, A. G., 1984. The Two Phanerozoic Supercycles, in Catastrophes and Earth History: The New
Uniformitarianism, W.A. Berggren and J.A. Van Couvering, (editors), Chapter 7, Princeton University Press,
Princeton, N.J., pp. 129 – 150.
Fisher, R.A., 1953. Dispersion on a sphere. Proceedings of the Royal Scoiety,v. A217, pp. 295-305.
Fisher, N.I., Lewis, T., Embleton, B.J.J, 1987. Statistical analysis of spherical data. Cambridge University
Press.
Flögel, S., Wallman, K., Poulsen, C.J., Zhou, J., Oschlies, A., Voigt, S., and Kuhnt, W., 2011. Simulating the
biogeochemical effects of volcanic CO2 degassing on the oxygen-state of the deep ocean during the Cenomanian-
Turonian Anoxic Evcent (OAE2), Earth and Planetary Science Letters, 305:371-384.
Flügel,, E. and Kiessling, W., 2002. Patterns of Phanerozoic reef crises. In Kiessling, W., Flügel,, E. and
Golonka, J. (eds): Phanerozoic Reef Patterns. SEPM (Society for Sedimentary Geology) Special Publications 72, 691-
733.
Foster, G.L., Royer, D.L., and Lunt, D.J., 2017. Future Climate Forcing Potentially without precedent in the
last 420 million years, Nature Communications, v. 8, 14845, doi:10.1038/ncomms14845. Forsyth, D., and Uyeda, S.,
1975. On the Relative Importance of the Driving Forces of Plate Motion, Geophysical Journal of the Royal
astronomical Society, v.43, pp. 163-200.
Foulger, G.R., and Jurdy, D.M., 2007. (editors), Plates, plumes, and planetary processes, Geological Society
of America, Special Paper 430, Boulder, CO, 997 p.
Frakes, L.A., 1999. Estimating the global thermal state from Cretaceous sea surface and continental
temperature data, in Evolution of the Cretaceous Ocean-Climate System, E. Barrera, E. and C. Johnson (eds.),
Geological Society of America Special Paper 332, Boulder, Colorado, .p. 49-57.
Frakes, L.A., Alley, N.F., and Deynoux, M., 1995. Early Cretaceous ice rafting and climate zonation in
Australia, International Geology Review, v. 37, p. 567-583.
Frakes, L.A., and Francis, J.E., 1988. A guide to Phanerozoic cold polar climates from high latitude ice-
rafting in the Cretaceous. Nature 333, 547–549.
Frakes L. A. and Francis J. E., 1990. Cretaceous palaeoclimates. In Cretaceous Resources, Events, and
Rhythms (ed. R. N. Ginsburg and B. Beaudoin), pp. 373–287. Kluwer Academic Publishers.
Frakes, L.A., Francis, J.E., and Sykes, J.I., 1992. Climate Modes of the Phanerozoic, Cambridge University
Press, Cambridge, 274 pp.
Francis J. E. and Frakes L. A., 1993. Cretaceous climates. Sediment. Rev. 1, 17–30.
Francois, C., Publellier, M., Robert, C., Bulois, C., Jamaludin, S.N.F., Oberhänsli, R., Faure, M., St-Onge,
M.R., and the IGCP 667 Team, 2021. Temporal and spatial evolution of orogens: a guide for geological mapping,
Episodes, v. 45, No. 3, pp. 265-283.
Franke, W., Haal, V., Oncken, O., and Tanner, D., 2000. (editors), Orogenic Processes: Quantification and
Modelling in the Variscan Belt, Geological Society Special Publication No. 179, The Geological Society, London, 459
p.
Friedman. R.M., 1983. Accretionary History of Western North America during the Mesozoic and Cenozoic
Eras, Masters thesis, Department of Geophysical Sciences, University of Chicago, Chicago, Illinois, August 1983, 512
p.
Friedrich, O., Norris, R.D., Erbacher, J., 2011. Evolution of Cretaceous oceans: A 55 million year record of
Earth's temperature and carbon cycle. Grzyb Found Spec Pub 17, 85-85.
Friedrich, O., Norris, R.D., Erbacher, J., 2012. Evolution of middle to Late Cretaceous oceans-A 55 m.y.
record of Earth's temperature and carbon cycle. Geology 40, 107-110.
Furon, R., 1941. La Paleogeographie, Essai sur l’evolution des continents et des oceans, Payot, Paris, 530
pp.
Gabrielse, H., and Yorath, C.J., 1992. (editors), Geology of the Cordilleran Orogen in Canada, Geology of
Canada, no. 4, Volume G-2, Geology of North America, 1
Gahagan, L., Scotese, C.R., Royer, J.-Y., Sandwell, D.T., Winn, J.K., Tomlins, R.I., Ross, M. I., Newman, J.S.,
Müller, R.D., Mayes, C.L., Lawver, L.A., and Heubeck, C.E., 1988. Tectonic fabric map of the ocean basins from
206
Satellite altimetry data, in C.R. Scotese and W. Sager, (editors), Mesozoic and Cenozoic Plate Tectonic
Reconstructions, Tectonophysics, pp. 1-26.
Gaina, C., Mueller, R.D., Royer, J.-Y., Stock, J., Hardebeck, J., and Symonds, P.A., The tectonics history of
the Tasman Sea: a puzzle with 13 pieces, J. Geophys. Res., 103:12413-12433.
Gaina, C., Roest, W., Mueller, R.D., and Symonds, P.A., 1998b. The opening of the Tasman Sea: a gravity
anomaly grid animation, Earth Interactions, 2-004, 1-23.
Gansser, A., 1964. Geology of the Himalayas, Interscience Publishers, London, 289 p.
Gaskell, D. E., Huber, M., O’Brien, C. L., Inglis, G. N., Acosta, R. P., Poulsen, C. J., and Hull, P. M. (2022). The
latitudinal temperature gradient and its climate dependence as inferred from foraminiferal δ 18 O over the past 95
million years. Proceedings of the National Academy of Sciences, 119(11).
https://doi.org/10.1073/pnas.2111332119
Gass, I.G., Lippard, S.J., and Shelton, A.W., 1984. Ophiolites and Oceanic Lithosphere, Geological Society
Special Publication 13, Blackwell Scientific Publications, Oxford, 413 p.
Gasson, E.G.W., and Kiessling, B.A., 2020. The Antarctic Ice Sheet, A Paleoclimate Modeling Perspective,
Oceanography, June 2020 Online Release, pp. 11, doi.org/10.5670/oceanog.2020.208
Gayer, R.A., 1989. (editor), The Caledonide Geology of Scandinavia, Graham and Trotman, London, 312 p.
Gee, D.G., and Sturt, B.A., 1985. The Caledonide Orogen – Scandinavia and Related Areas, Parts 1 and 2,
Wiley-Interscience, Chichester, UK, 1266 p.
Gibbs, M.T., Rees, P.Mc., Kutzbach, J.E., Ziegler, A.M., Behling, P.J., and Rowley, D.B., 2002. Simulations of
Permian climate and comparisons with climate-sensitive sediments, Journal of Geology, 110:33-55.
Gloaguen, R., and Ratschbacher, L., 2011. Growth and Collapse of the Tibetan Plateau, Geological Society
Special Publication, No. 353, Geological Society, London, 255 p.
Goddéris, Y., Donnadieu, Y., Lefebvre, V., Le Hir, G., and Nardin, E., 2012. Tectonic control of continental
weathering, atmospheric CO2 , and climate over Phanerozoic times: Comptes Rendus Geoscience, 344, 652–662.
Gold, T., 1965. Instability of the Earth’s axis of rotation, Nature, v. 175, pp. 526-529.
Goldreich,P. and Toomre. A., 1969. Some Remarks on Polar Wandering, Journal of Geophysical Research,
v. 74, no. 10, pp. 2555-2567.
Golonka, J., Ross, M. I., & Scotese, C. R., 1994. Phanerozoic Paleogeographic and Paleoclimatic Modeling
Maps, in A. F. Embry, B. Beauchamp, and D.J. Glass, editors, Pangea, Global Environments and Resources, Canadian
Society of Petroleum Geologists, Memoir 17, p. 1-47.
Golonka, J., 2000. Cambrian-Neogene Plate Tectonic Maps, Rozprawy Habilitacyine No. 350,
Wydawnictwo Uniwersytetu Jagiellonskiego, Krakow, 123 pp.
Gomez-Tuena, A., Straub, S.M., and Zellmer, G.F., 2014. Orogenic Andesites and Crustal Growth,
Geological Society Special Publication, No. 385, Geological Society, London, 414 p.
Gordon, C., Cooper, C., Senior, C.A., Banks, H., Gregory, J.M., Johns, T.C., Mitchell, J.F.B., Wood, R.A.,
2000. The simulation of SST, sea ice extents and ocean heat transports in a version of the Hadley Centre coupled
model without flux adjustments. Clim. Dyn. 16, 147-168.
Gornitz, V., 2008. Sea level change, post-glacial. In Encyclopedia of Paleoclimatology and Ancient
Environments (V. Gornitz, Ed.) Springer, p. 887-893.
Gough, D.O., 1981. Solar Interior Structure and Luminosity Variations. Sol Phys 74, 21.
Grabau, 1924. Principles of Stratigraphy, Volume 1 and 2, 1185 p., Dover, New York.
Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G. (Eds.). (2020). Geological Time Scale 2020 (Vol. 2).
Elsevier.
Granot, R., and Dyment, J., 2015. The Cretaceous Opening of the South Atlantic, Earth and Planetary
Science Letters, 414:156-163, http://dx.doi.org/10.1016/j.epsl.2015.01.015
Grasby, S.E., McCune, G.E., Beauchamp, B., Galloway, J.M., 2017. Lower Cretaceous cold snaps led to
widespread glendonite occurrences in the Sverdrup Basin, Canadian High Arctic. Geol. Soc. Am. Bull. 129, 771–787.
http://dx.doi.org/10.1130/B31600.1.
Green, A.R., Kaplan, A.A., and Vierbuchen, R.G., 1986. Circum-Arctic Petroleum Potential, in Future
Petroleum Provinces of the World, Proceedings of the Wallace E. Pratt Memorial Conference, Phoenix, AZ,
December, 1984, M.T. Halbouty (ed.), American Association of Petroleum Geologists, Memoir 40, pp. 101-130.
Grossman, E. L., 2012a. Applying Oxygen Isotope Paleothermometry in Deep Time. In L. Ivany and B.
Huber (Eds.), Reconstructing Earth’s Deep-Time Climate: State of the Art in 2012, Paleontological Society Short
207
Course (Vol. 18, pp. 39–67). The Paleontological Society Papers.
Grossman, E. L., 2012. Oxygen Isotope Stratigraphy. In F. M. Gradstein, J. G. Ogg, M. D. Schmitz, and G. M.
Ogg (Eds.), The Geologic Time Scale 2012 (Vol. 1, pp. 181–206). Elsevier.
Grossman, E. L., and Joachimski, M. M.,2020. Oxygen Isotope Stratigraphy. In Geologic Time Scale 2020
(pp. 279–307). Elsevier. https://doi.org/10.1016/B978-0-12-824360-2.00010-3
Grossman, E. L., and Joachimski, M. M.,2022. Ocean temperatures through the Phanerozoic reassessed.
Scientific Reports, 12(1), 8938. https://doi.org/10.1038/s41598-022-11493-1
Habicht, J. K. A., 1979. Paleoclimate, Paleomagnetism, and Continental Drift, AAPG Studies in Geology No.
9. American Association of Petroleum Geologists.
Haggart, J.W., Enkin, R.J., and Monger, J.W.H., 2006. (editors), Paleogeography of the North American
Cordillera: Evidence for and Against Large-Scale Displacements, Geological Association of Canada Special Paper 46,
Toronto, 420 p.
Hallam, A., 1973. (editor), Atlas of Palaeobiogeography, Elsevier Scientific Publishing Co,, Amsterdam, 531
p.
Hallam, A., 1984. Continental humid and arid zones during the Jurassic and Cretaceous, Palaeogeography,
Palaeoclimatology. and Palaeoecology, 47:195-223.
Hallam A., 1984. Pre-Quaternary sea-level changes. Annual Review of Earth and Planetary Sciences 12:
205-243, http://dx.doi.org/10.1146/annurev.ea.12.050184.001225.
Hallam, A., 2004. Catastrophes and Lesser Calamities: The Causes of Mass Extinctions, Oxford University
Press, Oxford, 226 p.
Hallam, A., and Wignall, P.B., 1997. Mass Extinctions and Their Aftermath, Oxford University Press, 320 p.
Hambrey, M. J., & Harland, W. B., 1981. Earth’s pre-Pleistocene glacial record, Cambridge University
Press, Cambridge, 1004 pp.
Hamilton, W., 1970. The Uralides and the Motion of the Russian and Siberian Platforms, Geological
Society of America Bulletin, 81:2553-2576.
Haq, B. U., 2014. Cretaceous eustasy revisited. In Global and Planetary Change (Vol. 113, pp. 44–58).
https://doi.org/10.1016/j.gloplacha.2013.12.007
Haq, B. U., and Schutter, S. R., 2009. A chronology of Paleozoic sea-level changes. Science, 322, 64-68.
Haq, B. U., Hardenbol, J., and Vail, P. R., 1987. Chronology of fluctuating sea levels since the Triassic.
Science, 235, 1156-1167.
Haq, B.U., Hardenbol, J., and Vail, P.R., 1988, Mesozoic and Cenozoic chronostratigraphy and cycles of sea-
level change. Sea-level changes: An integrated approach. SEPM, Special Publication 42:71-108.
Harper, D.A.T., and Servais, T., 2013. (editors), Early Paleozoic Biogeography and Palaeogeography,
Geological Society Memoir No. 39, Geological Society, London, 490 p.
Harper, D.A.T., Lefebvre, B., Percival, I.G., and Servais, T.,2023. (editors), A Global Synthesis of the
Ordovician System: Part I., Geological Society of London, Special Publications 535, doi.org/10.114/SP535-2022-168
Harper, D.A.T., Rasmussen, C.M.O., Liljeroth, M., Blodgett, R.B., Candela, Y., JIn. J.S., Percival, I.G., Rong,
J.Y., Villas, E., and Zhan, R.B, (2013), in D.A.T. Harper and T. Servais (editors), Early Paleozoic Biogeography and
Palaeogeography, Geological Society Memoir No. 39, Geological Society, London, pp. 127-144,
dx.doi.org/10.1144/ME38
Harris, A.L., Holland, C.H., and Leake, B.E., 1979. (editors), The Caledonides of the British Isles – reviewed,
Scottish Academic Press, Edinburgh, 768 p.
Harris, A.L., and Fettes, D.J., 1988. The Caledonian-Appalachian Orogen, Geological Society Special
Publication No. 38, Geological Society, Blackwell Scientific Publications, Oxford, 643 p.
Hart, M.B., 1996. (editor),Biotic Recovery from Mass Extinction Events, Geological Society Psoecial
Publication, No. 102, Geological Society, London, 392 p.
Hart, M.B., 2000. Climate modelling in the Cretaceous using the distribution of planktonic Foraminiferida,
in Climates: Past and Present, M.B. Hart (editor), Geological Society Special Publication, No. 181, 33-41.
Hart, M.B., and Koutsoukos, E.A.M., 2015. Paleoecology of Cretaceous Foraminifera: Examples from the
Atlantic Ocean and Gulf of Mexico Region. Transactions of the Gulf Coast Association of Geological Societies, 65,
175–199.
Hart, M.B., Harries, P.J., and Cardenas, A.L., 2013. The Cretaceous/Paleogene boundary events in the Gulf
Coast: comparisons between Alabama and Texas. Transactions of the Gulf Coast Association of Geological
208
Societies, 63, 235-255.
Hart, M.B., Leighton, A., Yancey, M. Hampton, Liu, C., Miller, B., Smart, C., and Twitchett, R., 2014.
Cretaceous-Paleogene boundary events in Texas: new sections, revised micropalaeontological interpretations and
clarification of the stratigraphy. STRATI, 2013, 37-41.
Hart, M.B., Yancey, T.E., Leighton, A.D., Miller, B., Liu, C., Smart, C.W., and Twitchett, R.J., 2012. The
Cretaceous-Paleogene boundary on the Brazos River, Texas: New stratigraphic sections and revised
interpretations, Journal of the Gulf Coast Association of Geological Societies, 1, 69-80.
Hartman, J.H., Johnson, K.R., and Nichols, D.J., 2002. The Hell Creek Formation and the Cretaceous-
Tertiary Boundary in the Northern Great Plains, Special Paper 361, Geological Society of America, Boulder, CO, 520
p.
Hasegawa, H., Tada, R., Jiang, X., Suganuma, Y., Imsamut, S., Charusiri, P., Ichinnorov, N., and Khand, Y.
(2012). Drastic shrinking of the Hadley circulation during the mid-Cretaceous Supergreenhouse. Climate of the
Past, 8(4), 1323–1337. https://doi.org/10.5194/cp-8-1323-2012
Hatcher, R.D., Jr., Carlson, M.P., and McBride, J.H., 2007. (editors), 4-D Framework of Continental Crust,
Geological Society of America, Memoir 200, Boulder, CO, 641 p.
Hatcher, R.D., Jr., Williams, H., and Zietz, I., 1983. (editors), Contributions to the Tectonics and Geophysics
of Mountain Chains, Geological Society of America, Memoir 158, Boulder, CO., 223 p.
Hawkesworth, C.J., Cawood, P.A., Dhuime, B., and Kemp, T.I.S., 2017. Earth’s continental lithosphere
through time, Annual Reviews of Earth and Planetary Sciences, v. 45, 169-198.
Hay, W. W., DeConto, R. M., de Boer, P., Flögel, S., Song, Y., and Stepashko, A. (2019). Possible solutions to
several enigmas of Cretaceous climate. International Journal of Earth Sciences, 108(2), 587–620.
https://doi.org/10.1007/s00531-018-1670-2
Hay, W. W., and Flögel, S. (2012). New thoughts about the Cretaceous climate and oceans. In Earth-
Science Reviews (Vol. 115, Issue 4, pp. 262–272). https://doi.org/10.1016/j.earscirev.2012.09.008
Hay, W.W., 2008. Evolving ideas about the Cretaceous climate and ocean circulation, Cretaceous
Research, 29: 725-753.
Hay, W.W., 2016. Experimenting on a Small Planet: A History of Scientific Discoveries, a Future of Climate
Change and Global Warming, Springer, Switzerland, 819 pp.
Hay, W.W., DeConto, R.M., de Boer, P, Flögel, Song, Y., and Stepashko, A., 2019. Possible solutions to
several enigmas of Cretaceous Climate, International Journal of Earth Science, v. 108: 587-620
Hayes, P.A., Francis, J.E., Cantrill, D.J., and Crame, J.A., 2006. Palaeoclimate analysis of Late Cretaceous
angiosperm leaf floras, James Ross Island, Antarctica, in Cretaceous-Tertiary High-Latitude Palaeoenvironments,
James Ross Basin, Antarctica, J.E., Francis, D. Pirrie, and J.A. Crame, (editors), Geol. Soc. London Special Publications
258:49-62.
Haywood AM, Valdes PJ, Aze T, Barlow NA, Dolan AM, von der Heydt AS, Hill DJ, Jamieson SSR, Otto-
Bliesner B., Salzmann U, Saup E., Voss J. 2019. What can Palaeoclimate Modelling do for you? Earth Systems and
Environment 3(1):1-18
Henderson, C.M., Davydov, V.I., Wardlaw, B.R., 2012. The Permian Period, Chapter 24, in The Geologic
Time Scale 2012, volume 2, F.M. Gradstein, J.G. Ogg, M.D. Schmitz, and G.M. Ogg (editors), Elsevier, Amsterdam,
p.653-679.
Henderson, L.J., 1985. Motion of the Pacific Plate Relative to the Hot Spots Since the Jurassic and Model
of Oceanic Plateaus of the Farallon Plate, Ph.D. dissertation, June, 1985, Northwestern University, 311 p.
Hendrix, M.S., and Davis, G.A., 2001. (editors), Paleozoic and Mesozoic tectonic evolution of central Asia:
From continental assembly to intracontinental deformation, Geological Society of America, Memoir 194, Boulder,
CO, 447 p.
Henkes, G.A., Passey, B.H., Grossman, E.L., Shenton, B. J., Yancey, T. E., and Pérez-Huerta, A., 2018.
Temperature evolution and oxygen isotope composition of Phanerozoic oceans from carbonate clumped isotope
thermometry, Earth and Planetary Science Letters, v. 490, pp. 40-50, doi:10.1016/j.epsl.2018.02.001.
Herman, A. B. and Spicer,R.A., 1996. "Paleobotanical evidence for a warm Cretaceous Arctic Ocean."
Nature 380(28 March 1996): 330-333.Herrle, J.O., Schröder-Adams, C.J., Davis, W., Pugh, A.T., Galloway, J.M., and
Fath, J., 2015. Mid-Cretaceous High Arctic stratigraphy, climate, and Oceanic Anoxic Events. Geology, 43(5): 403-
406
209
Herman, A.B., and Spicer, R.A., 1997. New quantitative paleoclimate data for the Late Cretaceous Arctic:
Evidence for a warm polar ocean, Palaeogeography, Palaeoclimatology, and Palaeoecology, 128:227-251
Herrle, J.O., Schröder-Adams, C.J., Davis, W., Pugh, A.T., Galloway, J.M., and Fath, J., 2015. Mid-
Cretaceous High Arctic stratigraphy, climate, and Oceanic Anoxic Events. Geology, 43(5): 403-406.
Hibbard, J.P., Pollock, J.C., Murphy, J.B., van Staal, C.R., and Greenough, J.D., 2015. Reeltime Geological
Syntheses: Remembering Harold “Hank” Williams, Geoscience Canada Reprint Series 10, Geological Association of
Canada, Ottawa, 486 p.
Hildebrand, A.R., Penfield, G.T., Kring, D.A., Pilkington, M., Camargo, Z.A., Jacobsen, S.B., and Boynton, W.V.,
1991. Chicxulub crater: A possible Cretaceous-Tertiary boundary impact crater on the Yucutan Peninsula, Mexico,
Geology, 19:867-871.
Hillhouse, J.W., 1989. (editor), Deep Structure and the Past Kinematics of Accreted Terranes, Geophysical
Monograph 50, IUGG v. 5, American Geophysical Union, International Union of Geodesy and Geophysics,
Washington, D.C., 283 p.
Hoffman, P.F., 1988. United Plates of America, Birth of a Craton: Early Proterozoic Assembly and Growth
of Laurentia, Annual Reviews of Earth and Planetary Sciences, 16:543-603.
Hönisch, B., Royer, D., Breecker, D.O., Polissar, P.J., Bowen, G., Henehan, M.J., Cui, Y., and 80 others,
2023. Towards a Cenozoic History of Atmospheric CO2 , The Cenozoic CO2 Proxy Integration Project (Cen CO2 PIP)
Consortium, Science, (in press), 45 pp.
Horrell, M.A., 1991. Phytogeography and palaeoclimatic interpretation of the Maastrichtian,
Palaeogeography, Palaeoclimatology, and Palaeoecology, 86:87-138.
Howell, D.G., 1989. Tectonics of Suspect Terranes: Mountain Building and Continental Growth, Chapman
and Hall, London, 232 p.
Hsü, K.J., 1982. (editor) Mountain Building Processes, Academic Press, London, 263 p.
Hsü, K.J., 1986. The Great Dying, Harcourt Brace Jovanovich Publishers, San Diego, 292 p.
Huber, B.T., 1998. Tropical paradise at the Cretaceous poles?, Science, v. 282, p. 2199-2200.
Huber, B.T., MacLeod, K.G., and Wing, S.L., 2000. Warm Climates in Earth History, Cambridge University
Press, 462 pp.
Huber, M., 2012. Progress in Greenhouse Climate Modeling. In L. Ivany and B. Huber (Eds.),
Reconstructing Earth’s Deep-Time Climate – The State of the Art in 2012, Paleontological Society Short Course: Vol.
Volume 18 (pp. 213–262). Paleontological Society Papers.
Hughes, N.F., 1973. Organisms and Continents through Time, Special Papers in Palaeontology, no. 12,
Palaeontological Association, London, 334 p.
Hulver, M., 1985. Cretaceous Marine Paleogeography of Africa, Master’s Thesis, University of Chicago,
Chicago
Hunter, S. J., P. J. Valdes, et al. , 2008. Modelling Maastrichtian climate: investigating the role of
geography, atmospheric CO2, and vegetation. Climates of the Past Discussion 4: 981-1019.
Hutchison, C.S., 1989. Geological Evolution of South-east Asia, Oxford Monographs on Geology and
Geophysics, No. 13, Clarendon Press, Oxford, 368 p.
IPCC 2007: Summary for Policymakers. IN: Climate Change 2007: The Physical Science Basis. Contribution
of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Solomon,
S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor, and H.L. Miller (eds.)]. Cambridge University
Press, Cambridge, United Kingdom and New York, USA.
IPCC 2018: Summary for Policymakers. In: Global Warming of 1.5˚C. An IPCC Special Report on the impacts
of global warming of 1.5˚C above pre-industrial levels and related greenhouse gas emission pathways, in the
context of strengthening the global response to the threat of climate change, sustainable development, and efforts
to eradicate poverty [Masson-Delmotte, V., P. Zhai, H.-O. Portner, D. Roberts, J. Skea, P.R. Sukla, A. Pirani, W.
Moufouma-Okia, C. Pean, R. Pidcock, S. Connors, J.B.R. Matthews, Y. Chen, X. Zhou, M.I. Gomis, E. Lonnoy, T.
Maycock, M. Tignor, and T. Waterfield (eds.)] Cambridge University Press, Cambridge, United Kingdom and New
York, USA.
IPCC 2019: Summary for Policymakers. In: IPCC Special Report on the Ocean and Cryosphere in a Changing
Climate [H.-O. Portner, D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, M.
Nicolai, A. Okem, J. Petzold, B. Rama, N. Weyer(eds.)] Cambridge University Press, Cambridge, United Kingdom and
New York, USA.
210
Isozaki, Y., 2014. Memories of pre-Jurassic lost oceans: How to retrieve them from extant lands.
Geoscience Canada: Journal of the Geological Association of Canada/Geoscience Canada: journal de l’Association
Géologique du Canada, 41(3), 283-311.
Isozaki, Y., Aoki, K., Nakam, T., and Yanai, S., 2010. New insight into a subduction-related orogen: a
reappraisal of the geotectonic framework and evolution of the Japanese Islands, Gondwana Research, 18 (1), 82-
105.
Isozaki, Y., Maruyama, S., Furuoka, F., 1990. Accreted oceanic materials in Japan, Tectonophysics, 131, (1-
4), 179-205.
Jabri, N., Scotese, C.R., and D.J. Main, 2010. Dinosaur biogeography: migration pathways and refugia,
Geological Society of America, 2010 annual meeting, abstracts with programs, v. 42:249.
Jacobs, L.L., Polcyn, M.J., Taylor, L.H., and Ferguson, K., 2005, Sea-surface temperatures and
paleoenvironments of dolichosaurs and early mosasaurs, Netherlands Journal of Geosciences, Geologie en
Mijnbouw, 84-3:269-281.
Jenkyns, H.C., Jones, C.J., Groeke, D.R., Hesselbo, S.P., and Parkinson, D.N., 2002. Chemostratigraphy of
the Jurassic System: applications, limitations and implications for palaeoceanography, Journal of the Geological
Society, London, v. 159, p. 351–378.
Jenkyns, H.C., Schouten-Huibers, L., Schouten, S., and Sinninghe Damste, J.S., 2012. Warm Middle
Jurassic–Early Cretaceous high-latitude sea-surface temperatures from the Southern Ocean, Climate of the Past,
8:215-226, doi:10.5194/cp-8-215-2012
Johnson, H., Dore, T., Gatliff, R.W., Holdsworth, R.W., Lundin, E.R., and Ritchie, J.D., 2008. The Nature and
Origin of Compression in Passive Margins, Geological Society Special Publication, No. 306, Geological Society,
London, 220 p.
Johnston, S.T., and Thorkelson, D.J., 2000. Continental flood basalts: Episodic magmatism above long-lived
hotspots, Earth and Planetary Science Letters, 175:247-256.
Jones, C.E., and Jenkyns, H.C., 2001. Seawater strontium isotopes, oceanic anoxic events, and seafloor
hydrothermal activity in the Jurassic and Cretaceous, American Journal of Science, 301:112-149.
Judd, E.J., Tierney, J.E., Huber, B., Wing, S.L., Lunt, D.J., Ford, H.L., Inglis, G.N., McClymont, E.L., O’Brien,
C.L., and 17 others, 2023. The PhanSST global database of Phanerozoic sea surface temperature proxy data,
Nature, Scientific Data, December 6, 2022, 37 p.
Kauffman, E.G., and Walliser, O.H., 1990. (editors), Extinction Events in Earth History, Springer-Verlag,
Berlin, 432 p.
Kay, S.M., and Ramos, V.A., 2006. Evolution of an Andean margin: A tectonic and magmatic view from the
Andes to the Neuquen Basin (35˚ - 39˚ S lat), Geological Society of America, Special Paper 407, Boulder, CO, 359 p.
Kazmin, V. G., & Natapov, L. M., 1998. The Paleogeographic Atlas of Northern Eurasia: Paleogeographic
Maps on the Palinspastic Reconstruction, Orthographic Projection, Institute of Lithospheric Plates, Russian
Academy of Natural Sciences, Moscow.
Keller, G., 2011. Defining the Cretaceous-Tertiary boundary: a practical guide and return to first principles.
In: Keller, G., Adatte, T. (Eds.), The KT Mass Extinction and the Chicxulub impact in Texas. SEPM Special Publication,
100, 23-42.
Keller, G., 2014. Deccan volcanism, the Chicxulub impact, and the end-Cretaceous mass extinction:
Coincidence? Cause and effect? In: Keller, G., Kerr, A. (Eds.), Volcanism, Impacts and Mass Extinctions: Causes and
Effects. Geological Society of America Special Papers, 505, 57-89, doi:10.1130/2014.2505(03).
Keller, G., Adatte, T., Gardin, S., Bartolini, A., and Bajpai, S., 2008. Main Deccan volcanism phase ends near
the K-T boundary: Evidence from Krishna-Godavari Basin, SE India, Earth and Planetary Science Letters, 268:293-
311.
Keller, G., Mateo, P., Punekar, J., Khozyem, H., Gertsch, B., Spangenberg, J., Bitchong, A., and Adatte, T.,
2017. Environmental Changes during the Cretaceous-Paleogene Mass Extinction and Paleocene-Eocene Thermal
Maximum: Implications for the Anthropocene, Gondwana Research, v. 56, p. 69-89,
doi.org/10.1016/j.gr.2017.12.002
Kennett, J.P., and Stott, L.D., 1991. Abrupt deep-water warming, palaeoceanographic changes and benthic
extinctions at the end of the Paleocene, Nature, v. 353, p.225-229.
Kent, W., Saunders, A.D., Kempton, P.D., and Ghose, N.C., 1997. Rajmahal Basalts, Eastern India: Mantle
Sources and Melts Distribution at a Volcanic Rifted Margin, in J.J. Mahoney and M.F. Coffin (editors), Large Igneous
211
Provinces: Continental, Oceanic, and Planetary Flood Volcanism, Geophysical Monograph 100, American
Geophysical Union, Washington, D.C., pp. 145-182.
Kerr, A.C., Tarney, J., Marriner, G.F>, Nivia, A., and Saunders, A.D., 1997. The Caribbean-Colombian
Cretaceous Igneous Province: The Internal Anatomy of an Oceanic Plateau, in J.J. Mahoney and M.F. Coffin
(editors), Large Igneous Provinces: Continental, Oceanic, and Planetary Flood Volcanism, Geophysical Monograph
100, American Geophysical Union, Washington, D.C., pp. 123-144.
Khain, V. Ye., Ronov, A. B., & Balukhovsky, A. N., 1976. Cretaceous lithologic associations of the world,
International Geology Review, v. 18, no.11, p. 1269-1295, English translation from Russian Sovetskaya Geologiya,
1976, v. 11, p. 10-39.
Kidder, D.L. and Worsley, T.R., 2010. Phanerozoic Large Igneous Provinces (LIPs), HEATT (Haline Euxinic
Acidic Thermal Transgression) episodes, and mass extinctions, Palaeogeography, Palaoeclimatology,
Palaeoecology, v. 295, pp. 162-191.
Kidder, D.L., and Worsley, T.T., 2012. A human-induced hothouse climate?, GSA Today, v.22, no. 2, pp. 4-
11, doi:10.1130/G131A.1
Kiehl, J.T., Shields, C.A., 2013. Sensitivity of the Palaeocene-Eocene Thermal Maxim XXX
Kiessling, W., 2001. Paleoclimatic significance of Phanerozoic Reefs, Geology, v. 29, no. 8, p. 751-754.
Kiessling W., Flügel E, Golonka J. 2002. Phanerozoic Reef Patterns. Tulsa: SEPM (Society for Sedimentary
Geology), Special Publication Number 72:1-775
Kiessling, W., and Flugel, E., 2002. PaleoReefs: A Database on Phanerozoic Reefs, in Phanerozoic Reef
Patterns, W. Kiessling, E. Flugel, and J. Golonka (editors), SEPM (Society for Sedimentary Geology) Special
Publication No. 72, p.77 -92.
King, C., A., 2016. A Revised Correlation of Tertiary Rocks in the British Isles and adjacent areas of NW
Europe, A.S. Gale and T.L. Barry (editors), Geological Society Special Report no. 27, 719 pp.
Kirschvink, J.L., Ripperdan, R.L., and Evans, D.A., 1997. Evidence for a Large-Scale Reorganization of Early
Cambrian Contienntal Masses by Inertial Interchange True Polar Wander, Science, 277: 541- 545.
Kocsis, A.T. and Scotese, C.R., 2020. Mapping paleocoastlines and continental flooding during the
Phanerozoic, Earth Science Reviews, v. 213, 1-15.
Köppen, W., 1918. Klassification der Klimate nach Temperatur, Niederschlag and Jahreslauf, Petermanns
Geographische Mitteilungen, v. 64, pp.193-248.
Köppen, W., 1936. Das geographisca System der Klimate. In W. Köppen and G. Gieger (Eds.), Handbuch
der Klimatologie (Vol. 1, pp. 1–44). Gebrüder Borntraeger.
Koppers, A.A.P., Becker, T.W., Jackson, M.G., Konrad, K., Müller, R.D., Romanowicz, Steinberger, B.,
Whittaker, J.M., 2021. Mantle plumes and their role in Earth processes, Nature Reviews, v. 2, June 2021, 382-401,
https://doi.org/10.1038/s43017-021-00168-6.
Kraus, D.W., Sertich, J.J.W., O’Connor, P.M., Rodgers, K.C., and Rogers, R.R., 2019. The Mesozoic
Biogeographic History of Gondwanan Terrestrial Vertebrates: Insighhts from Madagascar’s Fossil Record, Annual
Reviews of Earth and Planetary Sciences, 47:519-553
Kriest, J., 1991, Plate-Tectonic Atlas, Exploration Bulletin, no. 258, 1995/5, Shell Exploration Company,
Den Hague, 8 pp.
Kumar, P., Yuan, X.H., Kumar, M.R., Kind, R., Li, X.Q., and Chadha, R.K., 2007. The rapid drift of the Indian
tectonic plate, Nature Letters, volume 449, October 2007, pp. 894-897, doi:10.1038/nature06214
Kump, L.R., Pollard, D., 2008. Amplification of Cretaceous warmth by biological cloud feedbacks. Science
320, 195.
Kusky, T., B. Windley, I. Safonova, K. Wakita, J. Wakabayashi, A. Polat, and M. Santosh (2013). Recognition
of ocean plate stratigraphy in accretionary orogens through Earth history: A record of 3.8 billion years of sea floor
spreading, subduction, and accretion. Gondwana Research, 24(2), 501-547.
Ladant, J. B., and Donnadieu, Y. (2016). Palaeogeographic regulation of glacial events during the
Cretaceous supergreenhouse. Nature Communications, 7. https://doi.org/10.1038/ncomms12771
Ladant, J.B., Poulsen, C.J., Fluteau, F., Tabor, C.R., MacLeod, K.G., Martin, E.E., Haynes, S.J., Rostami, M.A.,
2020. Paleogeographic controls on the evolution of Late Cretaceous ocean circulation. Clim. Past. 16, 973-1006.
Lamb, S., 2004. Devil in the Mountain: A Search for the Origin of the Andes, Princeton University Press,
335 p.
212
Landwehrs, J., Feulner, G., Petri, S., Sames, B., and Wagreich, M., 2021. Investigating Mesozoic CLimate
Trends and Sensitivities With a Large Ensemble of Climate Simulations, Paleoceanography and Paleoclimatology,
Research article, 36, e2020PA004134, doi:10.1029/2020PA004134
Larson, R.L., 1975. Late Jurassic seafloor spreading in the eastern Indian Ocean, Geology, 3:69-71.
Larson, R.L., 1991. Geological consequences of superplumes, Geology, v.19, pp. 963-966.
Larson, R.L., 1991. Latest pulse of Earth: Evidence for a mid-Cretaceous superplume, Geology, v. 19, pp.
547-550.
Larson, R.L., and Erba, E., 1999. Onset of the mid-Cretaceous greenhouse in the Barremian-Aptian: Igneous
events and biological sedimentary, and geochemical responses, Paleoceanography, 14:663-678.
Larson, R.L., Mutter, J.C., Diebold, J.B., Carpenter, G.B., and Symonds, P., 1979. Cuvier basin: A product of
ocean crust formation by Early Cretaceous rifting off western Australia. Earth and Planet. Sci. Letters, 45:104-114.
Larson, R.L., Pitman, W.C., III, Golovchenko, X., Cande, S.C., Dewey, J.F., Haxby, W.F., and LaBrecque, J.L.,
1985, The bedrock geology of the world: New York, W.H. Freeman, scale 1:23,230,300 at the equator.
Lawver, L.A., and Müller, R.D.M., 1994. The Iceland Hotspot Track, Geology, 22:311-314.
Lawver, L.A., and Scotese, C.R., 1987. A revised reconstruction of Gondwanaland, in Gondwana Six:
Structure, Tectonics and Geophysics, G.D. McKenzie, (editor), Geophysical Monograph 40, American Geophysical
Union, Washington, D.C., pp.17-24.
Lawver, L.A., and Scotese, C.R., 1990. A review of tectonic models for the evolution of the Canada Basin, in
The Arctic Ocean Region, Geology of North America, volume L, A. Grantz, L. Johnson, and J.F.Sweeney (eds.),
Geological Society of America, Boulder, CO, pp. 593-618.
Le Pichon, X., Sengor, A.M.C., Jellinek, M., Lenardic, A., and Imren, C., 2023. Breakup of Pangea and the
Cretaceous Revolution, Tectonics, 42. E2022TC007489. Https://doi.org/10/1029/2022TC007489. 30 p.
Legates, D.R., and Wilmott, C.J., 1990. Mean seasonal and spatial variability in global surface air
temperature, Theoretical and Applied Climatology, v. 41, pp. 11-21.
Leggett, J.K., 1982. Trench-Forearc Geology: Sedimentation and Tectonics on Modern and Ancient Active
Plate Margins, Geological Society of London, Blackwell Scientific Publications, Oxford, 576 p.
Li, X., Hu, Y., Guo, J., Lan, J., Lin, Q., Bao, X., Yuan, S., Wei, M., Li, Z., Man, K., Yin, Z., Han, J., and Zhang, J.
(2022). A high-resolution climate simulation dataset for the past 540 million years. Scientific Data, 9.
https://doi.org/10.1038/s41597-022-01490-4
Li, Z.-X., Y. Liu, and R. Ernst, 2023. A dynamic 2000–540 Ma earth history: from cratonic amalgamation to
the age of supercontinent cycle. Earth-Science Reviews, 104336, https://doi.org/10.1016/j.earscirev.2023.104336.
Lieberman, B.S., 2000. Paleobiogeography: Using Fossils to Study Global Change, Plate Tectonics, and
Evolution, Topics in Geobiology, no. 16, N.H. Landman and D.S. Jones, (editors), Kluwer Academic / Plenum
Publishers, NewYork, 208 p.
Lieberman, B.S., 2003. Paleobiogeography: The Relevance of Fossils to Biogeography, Annual Reviews of
Ecology, Evolution, and Systematics, 34:51-69.dooi: 10.1146/annrev.ecolsys.34.121101.153549
Linnemann, U., Nance, R.D., Kraft, P., and Zulauf, G., 2007. (editors), The Evolution of the Rheic Ocean:
From Avalonian-Cadomian Active Margin to Alleghenian-Variscan Collision, Geological Society of America Special
Paper 423, Boulder, CO, 630 p.
Lonsdale, P., 2005. Creation of the Cocos and Nazca plates by fission of the Farallon plate. Tectonophysics
404, 237–264.
Lowrie, W., 2007. Fundamentals of Geophysics, 2nd edition, Cambridge University Press, Cambridge, UK,
381 p.
Lunt, D.J., Farnsworth, A., Loptson, C., Foster, G.L., Markwick, P., O'Brien, C.L., Pancost, R.D., Robinson, S.A.,
Wrobel, N., 2016. Palaeogeographic controls on climateXX
Lunt , D.J., Valdes, P., and Scotese, C., 2023. Changes in climate sensitivity and polar amplification over the
last 500 million years, (abstract), CL 1.1.4, Deep-time climate change: insights from models and proxies, European
Geophysical Union (EGU) General Assembly 2023,EGU23-9574, Vienna, Austria .
Luyendyk, B., 1995, Hypothesis for Cretaceous rifting of East Gondwana caused by subducted slab capture:
Geology, v. 23, p. 373–376, doi: 10.1130/0091-7613(1995)023 2.3.CO;2.
Lyell, C. 1830. Principles of Geology, Being an Attempt to Explain the Former Changes of the Earth’s
Surface, by Reference to Causes Now in Operation, Volume 1, 1st edition, John Murray: London, 481 p.
213
Lyell, C. 1832. Principles of Geology, Being an Attempt to Explain the Former Changes of the Earth’s
Surface, by Reference to Causes Now in Operation, Volume 2, 1st edition, John Murray, London, 332 p.
Lyell, C. 1833. Principles of Geology, Being an Attempt to Explain the Former Changes of the Earth’s
Surface, by Reference to Causes Now in Operation, Volume 3, 1st edition, John Murray, London, 109 p.
Macfarlane, A., Sorkhabi, R.B., Quade, J., 1999. Himalaya and Tibet: Mountain Roots to Mountain Tops,
Geological Society of America Special Paper 328, Boulder, CO, 330 p.
MacLeod, N., 2013. The Great Extinctions: What Causes them and How they Shape Life, Firefly Books,
Buffalo, NY, 298 p.
Mahoney, J.J., and Coffin, M.F., 1997. (editors), Large Igneous Provinces: Continental Oceanic, and
Planetary Flood Volcanism, Geophysical Monograph 100, American Geophysical Union, Washington, D.C., 438 p.
Malchus, N., and Steuber, T., 2002. Stable isotope records (O, C) of Jurassic aragonitic shells from England
and NW Poland: palaeoecologic and environmental implications, Geobios, v. 35, p. 29–39.
Mallory, W. W., 1972., Editor, Geological Atlas of the Rocky Mountain Region, Rocky Mountain
Association of Geologists, Denver, 331 pp.
Mann, P., 1995. Geologic and Tectonic Development of the Caribbean Plate Boundary in Southern Central
America, Geological Society of America, Special Paper 295, 349 p.
Mansour, A. and Wagreich, M., 2021. Earth system changes during the cooling greenhouse phase of the
Late Cretaceous: Coniacian – Santonian OAE3 subevents and fundamental variations in organic carbon deposition,
Earth-Science Reviews, doi:10.1016/j.earscirev.2022.10422
Markwick, P.J., 2004. Palaeo-digital elevation models for use as boundary conditions in coupled ocean-
atmosphere GCM experiments: a Maastrichtian (late Cretaceous) example, Palaeogeography, Palaeoclimatology,
and Palaeoecology, 213:37-63.
Markwick, P. J., 2007.. The palaeogeographic and palaeoclimatic significance of climate proxies for data-
model comparisons. In M. Williams, A. M. Haywood, F. J. Gregory, and D. N. Schmidt (Eds.), Deep-Time
perspectives on climate change: marrying the signal from computer models and biological proxies (pp. 251–312).
Geological Society of London.
Markwick P.J. 2019. Palaeogeography in Exploration, in Special Issue, “Advances in Palaeogeography,
Geological Magazine, 156(2):366-407
Matthews, K.J., Williams, S.E., Whittaker, J.M., Müller, R.D., Seton, M., and Clarke, G.L., 2015. Geologic
and kinematic constraints on the Late Cretaceous to mid-Eocene plate boundaries in the southwest Pacific, Earth-
Science Reviews, 140:72-107.
McArthur, J.M., Janssen,N.M.M, Reboulet, S., Leng, M.J., Thirwall, M.F., van de Schootbrugge, B., 2007.
Palaeotemperatures, polar ice-volume, and isotope stratigraphy (Mc/Ca, ∂18O, ∂13C, 87Sr/86Sr): The Early
Cretaceous (Berriasian, XXX
McCall, G.J.H., 1983. (editor), Ophiolitic and Related Melanges, Benchmark Papers in Geology 66,
Hutchinson Ross Publishing Company, Stroudsburg, PA, 446 p.
McClay, K.R., and Price, N.J., 1981. (editors), Thrust and Nappe Tectonics, Geological Society Special
Publication No. 9, Blackwell Scientific Publications, Oxford, 539 p.
McCrossan, R. G., Glaister, R. P., Austin, G. H., & Nelson, S. J., 1964. Geological History of Western Canada,
Alberta Society of Petrleum Geologists, Calgary, Alberta, 232 pp.
McElhinny, M.W., 1973. Palaeomagnetism and Plate Tectonics, Cambridge Earth Science Series, W.H.
Harland (editor), Cambridge University Press, Cambridge, 358 p.
McElhinny, W.W., and McFadden, P.L., 2000. Paleomagnetism: Continents and Oceans, International
Geophysics Series, R. Dmowska, volume 73, J.R. Holton, and H.T. Rossby (editors). Academiic Press, San Diego, 386
p.
McGhee, G.R., Clapham, M.E., Sheehan, P.M., Bottjer, D.J., and Droser, M.L., 2013. A new ecological-
severity ranking of major Phanerozoic biodiversity crises, Palaeogeography, Palaeoclimatology, Palaeoecology 370:
260–270 pp.
McInerney, F.A., Wing, S.L., 2011. The Paleocene-Eocene Thermal Maximum: a perturbation of carbon
cycle, climate, and biosphere with implications for the future. Annual Review of Earth Planetary Sciences,v. 39, p.
489–516.
214
McKenna, M.C., 1973. Sweepstakes, Filters, Corridors, Noah’s Arks, and Beach Viking Funeral Ships in
Paleogeography, in Implications of Continental Drift to the Earth Sciences, volume 1, D.H. Tarling and S.K. Runcorn
(editors), Academic Press, London, pp. 295-308
McKenzie, N. R., B. K. Horton, S. E. Loomis, D. F. Stockli, N. J. Planavsky, and C.-T. A. Lee (2016).
Continental arc volcanism as the principal driver of icehouse-greenhouse variability. Science, 352(6284), 444-447,
https://doi.org/10.1126/science.aad5787.
McKenzie, N.R., and Jiang,H., 2019. Earth’s outgassing and climatic transitions: The slow burn towards
environmental “catastrophes”?, Elements, v.15, p.325-330.
McKerrow, W.S., and Scotese, C.R. (eds.), Paleozoic Paleogeography and Biogeography, Memoir 12,
London: Geological Society of London, 435 p.
Meissner, R., T. Wever, and P. Sadowiak (1991). Continental collisions and seismic signature. Geophysical
Journal International, 105(1), 15-23.
Menzies, M.A., Klemperer, S.L., Ebinger, C.J., and Baker, J., 2002. (editors), Volcanic Rifted Margins,
Geological Society of America, Special Paper 362, Boulder, CO, 230 p.
Merdith, A. S., S. E. Williams, S. Brune, A. S. Collins, and R. D. Müller (2019.S). Rift and plate boundary
evolution across two supercontinent cycles. Global and Planetary Change, 173, 1-14.
Merdith, A. S., S. E. Williams, A. S. Collins, M. G. Tetley, J. A. Mulder, M. L. Blades, A. Young, S. E.
Armistead, J. Cannon, and S. Zahirovic (2021). Extending full-plate tectonic models into deep time: Linking the
Neoproterozoic and the Phanerozoic. Earth-Science Reviews, 214, 103477,
https://doi.org/10.1016/j.earscirev.2020.103477.
Metcalfe, I., 1984. Stratigraphy, palaeontology and palaeogeography of the Carboniferous of Southeast
Asia, Memoires Societe Geologiques France (New series), 174:107-118.
Metcalfe, I., 1993. Southeast Asian terranes: Gondwanaland origins and evolution, in Gondwana Eight:
Assembly, evolution and dispersal, R.H. Findlay, R. Uhrug, M.R. Banks, and J.J. Veevers, (editors), Balkema,
Rotterdam-Brookfield, pp. 181-200.
Metcalfe, I., 1999. Gondwana dispersion and Asian accretion: An overview, in I. Metcalfe, R. Jishun, J.
Charvet and S. Hada (editors), Gondwana Dispersion and Asia Accretion, A.A. Balkema, Rotterdam/Brookfield, pp.
9-28.
Metcalfe, I., 2011. Tectonic framework and Phanerozoic evolution of Sundaland, Gondwana Research,
19:3-21. Doi:10.1016/j.gr.2010.02.016
Metzler, M., 2006. The Formation and Obduction of Ophiolites Since the Late Precambrian, Master Thesis,
University of Texas at Arlington, 282 p.
Meyer, K.M., and Kump, L.R., 2008. Oceanic Euxinia in Earth History: Causes and Consequences, Annual
Review of Earth and Planetary Sciences, 36:251-288.
Middlemiss, F.A., Rawson, P.F., and Newall, G., 1971. Faunal Provinces in Space and Time, Proceedings of
the 17th Inter-University Geological Congress, Queen Mary College, University of London, 17,18,19 December,
1969, Geological Journal, Special Issue No. 4, Seel House Press, Liverppool, 236 p.
Mihalynuk, M.G., Nelson, J., and Diakow, L.J., 1994. Cache Creek terrane entrapment: Oroclinal paradox
with the Canadian Cordillera, Tectonics, 13: 575-595.
Miller, K.G., Browning, J.V., Scmetz, W.J., Kopp, R.E., Mountain, G.S., and Wright, J.D., 2020. Cenozoic sea-
level and cryosphere evolution from deep-sea geochemical and continental margin records, Science Advances, 6,
eaaz1346.
Miller, C.S., Petersen, F., da Silva, A.-C., Baranyi, V., Reichart, G.J., and Kürschner, W.M., 2017.
Astronomical age constraints and extinction mechanisms of the Late Triassic Carnian crisis. Nature, Scientific
Reports 7: doi:10.1038/s41598-017-02817-7.
Miller, K.G., Barrera, E., Olsson, R.K., Sugarman, P.J., Savin, S.M., 1999. Does ice drive early Maastrichtian
eustasy? Geology 27, 783–786.
Miller, K.G., Kominz, M.A., Browning, J.V., Wright, J.D., Mountain, G.S., Katz, M.E., Sugarman, P.J., Cramer,
B.S., Christie-Blick, N., Pekar, S.F., 2005a. The Phanerozoic record of global sea-level change. Science 310, 1293–
1298.
Miller, K.G., Sherrell, R.M., Browning, J.V., Field, M.P., Gallagher, W., Olsson, R.K., Sugarman, P.J., Tuorto,
S., and Wahyudi, H., 2010. Relationship between mass extinction and iridium across the Cretaceous-Paleogene
boundary in New Jersey, Geology, 38:867-870.
215
Miller, K.G., Sugarman, P.J., Browning, J.V., Kominz, M.A., Olsson, R.K., Feigenson, M.D., Hernández, J.C.,
2004. Upper Cretaceous sequences and sea-level history, New Jersey coastal plain. Geol. Soc. Am. Bull. 116, 368–
393.
Miller, K.G., Wright, J.D., J. V. Browning, J.V., 2005b. Visions of ice sheets in a greenhouse world. Mar.
Geol. 217, 215–231.
Mills, B.J.W., Krause, A.J., Scotese, C.R., Hill, D.J., Shields, G.A., and Lenton, T.M., 2019. Modelling the
long-term carbon cycle, atmospheric CO2 , and Earth surface temperature from late Neoproterozoic to present-
day, Gondwana Research, v. 67, p. 172-186, doi: 10.1016/j.gr.2018.12.001 1342-937.
Molnar, P., and Tapponnier, P., 1975. Cenozoic tectonics of Asia: Effects of a continental collision, Science,
189:419-426.
Monger, J.W.H., 2008. The Evolution of Canada’s Western Mountains, Geological Survey of Canada Open
File Report 504, 1 (poster), Vancouver, BC.
Monger, J.W.H., and Nokleberg, W.J., 1996. Evolution of the northern North American Cordillera:
generation, fragmentation, displacement and accretion of successive North American plate margin arcs, in Coyner,
A.R., and Fahey, P.L., (eds.), Geology and Ore deposits of the American Cordillera, Geol. Soc. Nevada Symposium
Proceedings, Reno/Sparks, Nevada, April 1995, pp. 1133-1152.
Moore, T. L., & Scotese, C. R., 2012. Ancient Earth: Breakup of Pangea, Vers. 1.0, iOS Mobile Application,
retrieved from http://itunes.apple.com/
Moores, E. M. (2002). Pre–1 Ga (pre-Rodinian) ophiolites: Their tectonic and environmental implications.
Geological Society of America Bulletin, 114(1), 80-95.
Moreno, T., Wallis, S., Kojima, T., and Gibbons, W., 2016. The Geology of Japan, Geological Society,
London, 522 p.
Morgan, P., and Baker, B.H., 1983. Processes of Continental Rifting, Selected papers from the Lunar and
Planetary Institute Topical Conference on the Processes of Planetary Rifting, held in St. Helena, CA, December 3-5,
1981, Developments in Geotectonics 19, reprinted from Tectonophysics, v.94, Amsterdam, 680 p.
Morgan, J.P., Reston, T.J., and Ranero, C.R., 2004. Contemporaneous mass extinctions, continental flood
basalts, and ‘impact signals’: are mantle plume-induced lithospheric gas explosions the causal link? Earth and
Planetary Science Letters, 217:263-284.
Morley, R.J., 2000. Origin and Evolution of Tropical Rain Forests, John Wiley and Sons, Ltd., Chichester, UK,
362 p.
Morley, R.J., 2011. Cretaceous and Tertiary climate change and the past distribution of megathermal
rainforests, in Tropical Rainforest Responses to Climatic Change, M.B. Bush, J.R. Flenley, W.D. Gosling, (editors), p.
1-34, Springer-Verlag, Berlin.
Mossop, G., & Shetson, I., 1994. Geological Atlas of Western Canada Sedimentary Basins, Canadian
Society of Petroleum Geologists, Calgary, 510 pp.
Müller, R. D., Cannon, J., Qin, X., Watson, R. J., Gurnis, M., Williams, S., et al. 2018. GPlates: Building a
virtual Earth through deep time. Geochemistry, Geophysics, Geosystems, 19. doi:10.1029/2018GC007584.
Müller R.D., Roest W.R., Royer J.Y., Gahagan L.M., Sclater J.G. 1993. A Digital Age Map of the Ocean Floor:
Scripps Institute of Oceanography (SIO) Reference Series 93-30, pp. 3, San Diego
Müller R.D., Roest WR, Royer J., Gahagan LM, Sclater JG. 1997, Digital isochrons of the world’s ocean
floor: Journal of Geophysical Research 102: 3211–3214
Müller, R.D., Royer, J.Y., and Lawver, L.A., 1993. Revised plate Motions relative to hotspots combined
Atlantic and Indian Ocean hotspot tracks, Geology, 21:275-278.
Müller R.D., Sdrolias M, Gaina C, Roest WR. 2008. Age, spreading rates, and spreading asymmetry of the
world’s ocean crust. Geochemistry Geophysics Geosystems 9(4):Q04006, https://doi.org/10.1029/2007GC001743
Müller R.D., Sdrolias M, Gaina C, Steinberger B, Heine C. 2008b. Long-term sea-level fluctuations driven by
ocean basin dynamics. Science 319:1357-1362. http://dx.doi. org/10.1126/science.l 151540
Mutterlose, J., Pauly, S., and Steuber, T., 2009. Temperature controlled deposition of early Cretaceous
(Barremian–early Aptian) black shales in an epicontinental sea, Palaeogeography, Palaeoclimatology,
Palaeoecology, 273:330-345, doi:10.1016/j.palaeo.2008.04.026
Muttoni, G., Dallanave, E., Channell, J.E.T., 2013. The drift history of Adira and Africa from 280 Ma to
present, Jurassic true polar wander, and zonal climate control on Tethyan sedimentary facies, Palaeogeography,
Palaeoclimatology, and Palaeoecology, 386:415-435, doi.org/10.1016/j.palaeo.2013.06.011
216
Muttoni, G., and Kent, D.V., 2091. Jurassic Monster Polar Shift Confirmed by Sequential Paleopoles from
Adria, Promontory of Africa, Journal of Geophysical Research: Solid Earth, 124:3288-3306,
doi.org/10/1029/2018JB017199
Nakanishi, M., Sager, W.W., and Korenaga, J., 2015. Reorganization of the Pacific-Izanagi-Farallon triple
junction in the Late Jurassic: Tectonic events before the formation of the Shatsky Rise, in C.R. Neal, W.W. Sager, T.
Sano, and E. Erba, (editors), The Origin, Evolution, and Environmental Impact of Oceanic Large Igneous Provinces,
Geological Society of America, Special Paper 511, Boulder, CO, 339 pp. 85-101, doi:10.1130/2015.2511(05)
Nance, R.D., and Murphy, J.B., 2013. Origins of the supercontinent cycle, Geoscience Frontiers, 4:439-448,
https://doi.org/10.1016/j.gsf.2012.12.007
Nance, R.D., and Thompson, M.D., 1996. Avalonian and Related Peri-Gondwanan Terranes of the Circum-
North Atlantic, Geological Society of America Special Publication 304, 390 p.
Nance, R.D., Murphy, J.B., and Santosh, M., 2014. The supercontinent cycle: A retrospective essay,
Gondwana Research, 25:4-29
Nance, R.D., Worsley, T.R., and Moody, J.B., 1988. The Supercontinent Cycle, Scientific American, v. 259,
No. 1, pp. 72-79. Https://www.jstor.org.stable/10.2307/24989160
Navarro-Ramirez, J.P., Bodin, S., Consorti, L., and Immenhauser, A., 2017. Response of western South
American epeiric-neritic ecosystem to middle Cretaceous Oceanic Anoxic Events, Cretaceous Research, v. 75, p.
61–80
Neal, C.R., Mahoney, J.J., Kroenk, L.W., Duncan, R.A., and Petterson, M.G., 1997. The Ontong Java Plateau,
in J.J. Mahoney and M.F. Coffin (editors), Large Igneous Provinces: Continental, Oceanic, and Planetary Flood
Volcanism, Geophysical Monograph 100, American Geophysical Union, Washington, D.C., pp. 183-216.
Neal, C.R., Sager, W.W., Sano, T., and Erba, E., 2015. (editors) The Origin, Evolution, and Environmental
Impact of Oceanic Large Igneous Provinces, Geological Society of America, Special Paper 511, Boulder, CO, 339 p.
Nichols, D.J., and Johnson, K.R., 2008. Plants and the K-T Boundary, Cambridge University Press, 280 p.
Nitecki, M.H., 1984. (editor), The University of Chicago Press, Chicago, 354 p.
Nokleberg, W.J., Parfenov, L.M., Monger, J.W. H., Norton, I.O., Khanchuk, A.I., Stone, D.B., Scotese, C.R.,
Scholl, D.W., and Fujita, K., 2001. Phanerozoic tectonic evolution of the Circum-North Pacific, U.S.G.S. Professional
Paper 1626, pp. 122.
O’Brien, C.L., Robinson, S.A., Pancost, R.D., Sinninghe Damsté, J. S., Schouten, S., Lunt, D. J., Alsenz, H.,
Bornemann, A., Bottini, C., Brassell, S. C., Farnsworth, A., Forster, A., Huber, B. T., Inglis, G. N., Jenkyns, H. C.,
Linnert, C., Littler, K., Markwick, P., McAnena, A., Mutterlose, J., Naafs, B. D. A., Püttmann, W., Sluijs, A., van
Helmond, N. A. G. M., Vellekoop, J., Wagner, T., and Wrobel, N. E., 2017. Cretaceous sea-surface temperature
evolution: Constraints for TEX 86 and planktonic foraminiferal Oxygen isotopes, Earth-Science Reviews, v. 172, p.
224-247, doi: 10.1016/j.earscirev.2017.07.012. 2018
Officer, C.B., and Drake, C.L., 1985. Terminal Cretaceous environmental events, Science, 227:1161-1167.
Officer, C.B., and Page, J., 1996. The Great Dinosaur Extinction Controversy, Helix Books, Reading, MA, 209
p.
Ogg, J.G., Ogg, G.M., and Gradstein, F.M., 2016. A Concise Geological Time Scale 2016, Elsevier, Amsterdam,
234 pp.
Olsson, R.K., Miller, K.G., Browning, J.V., Wright, J.D., and Cramer, B.S., 2002. Sequence stratigraphy and
sea level change across the Cretaceous-Tertiary boundary on the New Jersey passive margin, in C. Koeberl and
McLoed, K.G., (eds.), Catastrophic Events and Mass Extinctions: Impacts and Beyond, Geol. Soc. America Special
Paper 356:97-108.
Oncken, O., Chong, G., Fraunz, G., Giese, P., Götze, H.-J., Ramos, V.A., Strecker, M.R., and Wigger, P.,
2006. (editors), The Andes: Active Subduction Orogeny, Springer, Berlin, 569 p.
Ostrom, J.H., and Orville, P.M., 1975. (editors), Tectonics and Mountain Ranges, A special volume of the
American Journal of Science Dedicated to John Rodgers, American Journal of Science, v. 275-A, New Haven, CT,
518 p.
Otto-Bliesner, B. L. and G. R. Upchurch, 1997. Vegetation-induced warming of high latitude regions during
the Late Cretaceous period. Nature 385: 804-807.
Oxburgh, E.E., 1968. The Geology of the Eastern Alps. The Geologists’ Association, London, 127 p.
217
Panayioyou, A., 1980. (editor), Ophiolites: Proceedings of the International Ophiolite Symposium, Cyprus,
1979, Ministry of Agriculture and Natural Resources, Geological Survey Department, Republic of Cyprus, Nicosia
781 p.
Pankhurst, R.J., and Rapela, C.W., 1998. The Proto-Andean Margin of Gondwana, Geological Society
Special Publication No. 142, The Geological Society, London, 383 p.
Parfenov, L.M., Berzin, N.A., Badarch,G., Belichenko, V.G., Bulgatov, A.N., Dril, Khanchuk, A.I., Kirillova,
G.L., Kuzmin, M.I., Nokleberg, W.J., Ogasawara, M., Oblenskiy, A.A., Prokopiev, A.V., Rodionov, Scotese, C.R.,
Timofeev, V.F., Tomurtogoo, O. Yan, H.Q., 2010. Metallogenesis and Tectonics of Northeast Asia Edited by Warren
J. Nokleberg U.S. Geological Survey Professional Paper 1765 Chapter 9 Tectonic and Metallogenic Model for
Northeast Asia, USGS, Washington, D.C., 55 p.
Parfenov, L.M., Nokleberg, W.J., Berzin, N.A., Badarch, Gombosuren, Dril, S.I., Gerel, Ochir, Goryachev,
N.A., Khanchuk, A.I., Kuz'min, M.I., Obolenskiy, A.A., Prokopiev, A.V., Ratkin, V.V., Rodionov, S.M., Scotese, C.R.,
Shpikerman, V.I., Timofeev, V.F., Tomurtogoo, Onongin, and Yan, Hongquan; Nokleberg, W.J., ed, 2011, Tectonic
and metallogenic model for northeast Asia: U.S. Geological Survey Open-File Report 2011-1026., Washington, D.C.,
630 p.
Parrish, J.T., 1998. Interpreting Pre-Quaternary Climate from the Geologic Record, Columbia University
Press, New York, 338 pp.
Parrish, J.T., and Spicer, R.A., 1988, Late Cretaceous terrestrial vegetation: A near polar temperature
curve, Geology, 16:22-25.
Parrish, J.T., A.M. Ziegler, and Scotese, 1982. Rainfall patterns and the distribution of coals and evaporites
in the Mesozoic and Cenozoic, Palaeogeography, Palaeoclimatology, Palaeoecology, v.40:67-101.
Parsons B, Sclater JG. 1977. An analysis of the variation of ocean floor bathymetry and heat flow with age.
Journal of Geophysical Research 82(5):803-827
Patriat, P., and Segoufin, J., 1988. Reconstruction of the Central Indian Ocean, in Mesozoic and Cenozoic
Plate Reconstructions, C.R. Scotese and W.W. Sager (eds.), Tectonophysics, 155:211-234.
Paul, G.S., 2016. The Princeton Field Guide to Dinosaurs, 2nd edition, Princeton University Press,
Princeton, 360 p.
Paul, G.S., 2022. The Princeton Field Guide to Mesozoic Sea Reptiles, Princeton University Press,
Princeton, 208 p.
Paul, G.S., 2022. The Princeton Field Guide to Pterosaurs, Princeton University Press, Princeton, 184 p.
Peat, D.W., 1997. The Parana-Entendeka Province, in J.J. Mahoney and M.F. Coffin (editors), Large Igneous
Provinces: Continental, Oceanic, and Planetary Flood Volcanism, Geophysical Monograph 100, American
Geophysical Union, Washington, D.C., pp. 217- 245.
Peltier WR. 2004. Global Glacial Isostasy and the Surface of the Ice-Age Earth: The ICE-5G (VM2) Model
and GRACE. Annual Review of Earth and Planetary Sciences 32:111-149
Percival, L.M.E., Matsumoto, H., Callegaro, S., Erba, E., Kerr, A.C., Mutterlose, J., and Suzuki, K., 2024.
Cretaceous large igneous provinces: from volcanic formation to environmental catastrophes and biological crises,
in M.B. Hart, S.J. Batenburg, B.T. Huber, G.D. Price, N. Thibault, M. Wagreich, and I. Walaszczyk, (editors),
Cretaceous Project 200, Volume 1, The Cretaceous World, Geological Society, London, Special Publications, 544,
https://doi.org.10.1144.SP544-2023-88
Petersen, S.V., Dutton, A., Lohmann, K.C., 2016. End-Cretaceous extinction in Antarctica linked to both
Deccan volcanism and meteorite impact via climate change, Nature Communications, 7:12079 doi:
10.1038/ncomms12079
Pfiffner, O.A., 2014. Geology of the Alps, revised and updated translation of Geologie der Alpen, second
edition, Wiley – Blackwell, Chichester, UK, 376 p.
Pindell, J., and Barrett, S.F., 1990. Geological evolution of the Caribbean region: a plate tectonic perspective,
in Dengo, G., and Case, J.E., (editors), The Caribbean Region, Geology of North America, DNAG v. H, Geological
Society of America, Boulder, Colorado, pp. 405-432.
Pindell, J., Cande, S.C., Pitman III, W.C., Rowley, D.B., Dewey,J.F., LaBreque, J., and Haxby, W., 1988. A plate
kinematic framework for models of Caribbean evolution, in Scotese, C.R., and Sager, W.W., (editors), Mesozoic and
Cenozoic Plate Reconstructions, Tectonophysics, 155: 121-138.
218
Pindell, J., and Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean, and norhern South
America in the mantle reference frame: an update, Geological Society of London Special Publication 328:1-55,
https://doi.org/10.1144/SP328.1
Pindell, J., Kennan, L., Maresch, W.V., Stanel, K.-P., Draper, G., Higgs, R., 2005. Plate-kinematics and crustal
dynamics of Circum-Caribbean arc-continent interactions: Tectonic controls on basin development in Proto-
Caribbean margins, in Ave-Lallemant, H.G., and Sisson, V.B., (editors),Caribbean-South America Plate Interactions,
Venezuela, Geological Society of America Special Paper 394:7-52.
Pirrie, D., Doyle, P., Marshall, J., Ellis, G., 1995. Cool Cretaceous climates: new data from the Albian of
Western Australia. J. Geol. Soc. 152: 739–742.
Pirrie, D., Marshall, J., Doyle, P., Riccardi, A., 2004. Cool early Albian climates; new data from Argentina.
Cretaceous Research 25:, 27–33.
Pisarevsky, S. A., 2005. New edition of the Global Paleomagnetic Database (GPMDB4.6b), EOS, v. 86 (17),
170 pp.
Pisarevsky, S.A., Li, Z.X., Tetley, M.G., Liu, Y., Beardmore, J.P., 2022. An updated internet-based Global
Paleomagnetic Database, Earth-Science Review, 235 (2022) 104258,
https://doi.org/10.1016/j.earscirev.2022.104258
Poblete F, Dupont-Nivet G, Licht A, van Hinsbergen DJJ, Roperch P, Mihalynuk MG, Johnston S.,
Guillocheau F, Baby G, Fluteau F, Robin C, Baatsen M. 2020. Global paleogeographic maps through the
Greenhouse-Icehouse transition at 60, 40 and 20 Ma (in press.)
Podlaha, O.G., Mutterlose, J., and Veizer, J., 1998. Preservation of d18O and d13C in belemnite rostra
from Jurassic/early Cretaceous successions, American Journal of Science, v. 298, p. 324-347.
Pope, V.D., Gallani, M.L., Rowntree, P.R., Stratton, R.A., 2000. The impact of new physical
parametrizations in the Hadley Centre climate model: HadAM3. Clim. Dyn. 16, 123-146.
Poulsen, C.J., Barron, E.J., Arthur, M.A., Peterson, W.H., 2001. Response of the mid-Cretaceous global
oceanic circulation to tectonic and CO2 forcings. Paleoceanography 16, 576-592.
Poulsen, C.J., Barron, E.J., Peterson, W.H., and Wilson, P.A., 1999. A reinterpretation of mid-Cretaceous
shallow marine temperatures through model-data comparison, Palaeoceanography, v. 14, no. 6, pp. 679-697.
Poulsen, C. J., Pollard, D., and White, T. S., 2007. General circulation model simulation of the δ18O
content of continental precipitation in the middle Cretaceous: A model-proxy comparison. Geology, 35(3), 199–
202. https://doi.org/10.1130/G23343A.1
Powell, C. McA., Roots, S.R., and Veevers, J.J., 1988. Pre-breakup continental extension in East
Gondwanaland and the early opening of the eastern Indian Ocean, in Mesozoic and Cenozoic Plate
Reconstructions, C.R. Scotese and W.W. Sager (eds.), Tectonophysics, 155:261-283.
Powell, J.L., 1998. Night Comes to the Cretaceous: Comets, Craters, Controversy, and the Last Days of the
Dinosaurs, Harcourt Brace and Company, San Diego, 250 p.
Preto, N., Kustatscher, E., and Wignall, P.B. 2010. Triassic climates—state of the art and perspectives.
Palaeogeography, Palaeoclimatology, Palaeoecology, 290:1–10.
Price, G.D., 1999. The evidence and implications of polar ice during the Mesozoic. Earth Sci. Rev. 48, 183–
210.
Price, G.D., Mutterlose, J., 2004. Isotopic signals from late Jurassic–early Cretaceous (Volgian–Valanginian)
sub-Arctic belemnites, Yatria River, Western Siberia. J. Geol. Soc. 161, 959–968.
Price, G.D., Nunn, E.V., 2010. Valanginian isotope variation in glendonites and belemnites from Arctic
Svalbard: transient glacial temperatures during the Cretaceous greenhouse. Geology 38, 251–254.
Price, G.D., Passey, B.H., 2013. Dynamic polar climates in a greenhouse world: evidence from clumped
isotope thermometry of Early Cretaceous belemnites. Geology 41, 923–926.
Price, G.D., Valdes, P.J., and Sellwood, B.W., 1998. A comparison of GCM simulated Cretaceous
‘greenhouse’ and ‘icehouse’ climates: implications for the sedimentary record, Palaeogeography,
Palaeoclimatology, Palaeoecology, 142:123-138.
Price, R.A., and Monger, J.W.H., 2000. A transect of the southern Canadian Cordillera from Calgary to
Vancouver, Annual Meeting of the Cordilleran Section of the Geological Society of America, Vancouver, April 2000,
GeologicalAssociation of Canada, Vancouver, 164 p.
219
Prokoph, A., Shields, G. A., and Veizer, J., 2008, Compilation and time-series analysis of a marine
carbonate δ18O, δ13C, 87Sr/86Sr and δ34S database through Earth history, Earth-Science Reviews, v. 87, no. 3, p. 113-
133.
Pucéat, E., Lécuyer, C., Sheppard, S.M., Dromart, G., Reboulet, S., Grandjean, P., 2003. Thermal evolution
of Cretaceous Tethyan marine waters inferred from oxygen isotope composition of fish tooth enamels.
Paleoceanography 18, 1029. http://dx.doi. org/10.1029/2002PA000823.
Punekar, J., Mateo, P., and Keller, G., 2014. Effects of Deccan volcanism on paleoenvironment and planktic
foraminifera: A global survey, Geological Society of America Special Paper, v. 505, p. 91-116,
doi:10.1130/2014.2505(04)
Rae, J.W.B., Zhang, Y.G., Liu, X., Foster, G.L., Stoll, H.M., and Whiteford, 2021. Atmospheric CO2 over the
Past 66 Million Years from Marine Archives, Annual Review of Earth and Planetary Sciences, 49: 609-41.
Rampino, M.R., and Self, S., 2015. Large Igneous Provinces and Biotic Extinctions, in The Encyclopedia of
Volcanoes, second edition, H. Sigurdsson, B. Houghton, H. Rymer, J. Stix, and S. McNutt (editors), p.1049-158,
Elsevier, Amsterdam.
Rampino, M.R., and Stothers, R.B., 1988. Flood Basalt Volcanism during the Past 250 Million Years,
Science, 241: 663-668.
Raup, D.M., 1986. The Nemesis Affair: A Story of the Death of the Dinosaurs and the Ways of Science,
W.W. Norton and Company, New York, 220 p.
Raup, D.M., 1991. Extinction: Bad Luck or Bad Gene?, W.W. Norton, New York, 210 p.
Rea, D.K., Zachos, J.C., Owen, R.M., and Gingerich, P.D., 1990. Global change at the Paleocene-Eocene
boundary: climatic and evolutionary consequences of tectonic events, Palaeogeography, Palaeoclimatology,
Palaeoecology 79:117-128.
Rees, P.Mc., Gibbs, M.T., Ziegler, A.M., Kutzbach, J.E., Behling, P.J., and Rowley, D.B., 2002. Permian
phytogeographic patterns and climate data/model comparisons, Journal of Geology, 27:891-894.
Renaut, R.W., and Ashley, G.M., 2002. (editors), Sedimentation in Continental Rifts, SEPM (Society for
Sedimentary Geology), Special Publication 73, Tulsa, OK, 326 p.
Renne, P.R., Sprain, C.J., Richards, M.A., Self, S., Vanderkluysen, L., and Pande, K., 2015. State shift in
Deccan volcanism at the Cretaceous-Paleogene boundary, possibly induced by impact, Science 350 (6256):76-78.
Retallack, G., 2013. Permian and Triassic greenhouse crises, Gondwana Research, v. 24, p. 90–103,
doi:10.1016/j.gr.2012.03.003.
Retallack , G.J., Sheldon, N.D., Carr,P.F., Fanning, M., Thompson, C.A., Williams, M.L., Jones, B.J., and
Hutton, A., 2011. Multiple Early Triassic greenhouse crises impeded recovery from Late Permian mass extinction,
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 308, 233–251, doi: 10.1016/j.palaeo.2010.09.022
Richards, M.A., Alvarez, W., Self, S., Karlstrom, L., Renne, P.R., Manga, M., Sprain, C.J., Smit, J., Roche, V.,
and Ringebach, J.C., 2022. The Davie Fracture Zone: A recorder of continents drifts and kinematic changes,
Tectonophysics, v. 823, 229188, https://doi.org/10.1016/j.tecto.2021.229188
Rigo, M., and Joachimski, M.M., 2010. Palaeoecology of Late Triassic conodonts: constraints from oxygen
isotopes in biogenic apatite, Acta Palaeontologica Polonica, v. 55, p. 471–478.
Roberts, D.G., and Bally, A.W., 2012. (editors), Regional Geology and Tectonics :Phanerozoic Rift systems
and Sedimentary Basins, Volume 1B, Elsevier, Amsterdam, 528 p.
Roberts, D.G., and Bally, A.W., 2012. (editors), Regional Geology and Tectonics: Phanerozoic Passive
Margins, Cratonic Basins, and Global Tectonic Maps, Volume 1C, Elsevier, Amsterdam, 1196 p.
Rodgers, J., 1970. The Tectonics of the Appalachians, Wiley-Interscience, New York, 271 p.
Rogov, M. Ershova, V., Gaina, C., Vershchagin, O., Vasileva, K., Mikhailova, K., and Krylov, A., 2023.
Glendonites throughout the Phanerozoic, Earth-Science Reviews, v. 241, 104430.
Ronov, A. B., Khain, V. Ye., & Balukhovsky, A., 1989. Atlas of Lithological-Paleogeographical Maps of the
World, Mesozoic and Cenozoic of the Continents, USSR Academy of Sciences, USSR State Committee for Public
Education, Ministry of Geology of the USSR, 79 pp., Leningrad.
Ross CA, Ross J R. 1985. Late Paleozoic depositional sequences are synchronous and worldwide. Geology,
13:194-197
Ross, M.I., and Scotese, C.R., 1988. A hierarchical tectonic model of the Gulf of Mexico and Caribbean
Region, in Mesozoic and Cenozoic Plate Reconstructions, C.R. Scotese and W.W. Sager (editors), Tectonophysics,
155: 139-168.
220
Rowley D.B. 2018. Oceanic axial depth and age-depth distribution of oceanic lithosphere: Comparison of
magnetic anomaly picks versus age-grid models, Lithosphere, 11(1):21-43, GSA Data Repository Item 2018395,
doi:/10/1130/L1027.1
Rowley, D.B., and Lottes, A.L., 1988. Plate-kinematic reconstructions of the North Atlantic and Arctic: Late
Jurassic to Present, in Mesozoic and Cenozoic Plate Tectonic Reconstructions, C.R. Scotese and W.W. Sager, (eds),
Tectonophysics, 155:73-120.
Rowley D.B., Currie, BS. 2006. Paleo-altimetry of the late Eocene to Miocene Lunpola basin, central Tibet.
Nature, 439:677-681
Rowley D.B., Garzione, CN. 2007. Stable isotope-based paleoaltimetry. Annual Review of Earth and
Planetary Science, 35:463-508
Rowley D.B., Pierrehumbert, RT, Currie, BS. 2001. A new approach to stable isotope-based paleoaltimetry:
implications for paleoaltimetry and paleohypsometry of the High Himalaya since the Late Miocene. Earth and
Planetary Science Letters, 188:253-26
Royden, L.H., and Husson, L., 2006. Trench motion, slab geometry, and viscousstresses in subduction
systems, Geophysical Journal international, 167:881-906, doi:10.1111/j.1365-246X.2006.03079.x.
Royer, J.-Y., Patriat, P., Bergh, H.W. and Scotese, C.R., 1988. Evolution of the Southwest Indian Ridge from
the late Cretaceous (Anomaly 34) to the Middle Eocene (anomaly 20), Tectonophysics, 155: 235-260.
Royer, J.-Y., R.D. Müller, L.M. Gahagan, L.A. Lawver, C.L. Mayes, D. Nürnberg and J.G. Sclater, 1992. A
global isochron chart, University of Texas Institute for Geophysics Technical Report No. 117, pp. 38.
Royer, J.-Y., Sclater, J.G., Sandwell, D.T., Cande, S.C., Schlich, R., Munschy, Dyment, R.D., Fisher, R.L.,
Mueller, R.D., Coffin, M.F., Patriat, P., Bergh, H.W., 1992. Indian Ocean Plate Reconstructions Since the Late
Jurassic, in R.A. Duncan, D.K. Rea, R.B. Kidd, U. von Rad, and J.K. Weissel (eds.), Synthesis of Results from Scientific
Drilling in the Indian Ocean, Geophysical Monograph 70: 471-475, Amer. Geophys. Union, Washington, D.C.
Royer, J.-Y., and Sandwell, D.T., 1989. Evolution of the eastern Indian Ocean since the Late Cretaceous:
Constraints from Geosat altimetry, Journal of Geophysical Research, 94:775-782.
Royer, D. L., Berner, R. A., Montañez, I. P., Tabor, N. J., and Beerling, D. J. (2004). CO2 as a primary driver
of Phanerozoic climate. GSA Today, 14(3), 4. https://doi.org/10.1130/1052-5173(2004)014<4:CAAPDO>2.0.CO;2
Richards, M., Duncan, R., Courtillot, V., 1989, Flood basalts and hot-spot tracks: Science, v. 246, p. 103–
107.
Safonova, I. Y., V. Simonov, M. Buslov, T. Ota, and S. Maruyama (2008). Neoproterozoic basalts of the
Paleo-Asian Ocean (Kurai accretionary zone, Gorny Altai, Russia): geochemistry, petrogenesis, and geodynamics.
Russian Geology and Geophysics, 49(4), 254-271.
Sager, W.W., 2007. Divergence between paleomagnetic and hotspot -model-predicted polar wander for
the Pacific plate with implications for hotspot fixity, in G.R. Foulger and D.M. Jurdy (editors), Plates, plumes, and
planetary processes, Geological Society of America, Special Paper 430, Boulder, CO, pp. 335-357.
Salles, T., Husson, L., Rey, P., Mallard, C., Zahirovic, S., Boggiani, B.H., Coltice, N., and Arnould, M., 2023.
Hundred Million years of landscape dynamics from catchment to global scale, Science, 379:918-923.
Salvador, A., 1991. The Gulf of Mexico, Geology of North America, Volume J, Geological Society of
America, Boulder, CO., 568 p.
Sandwell, D., Garcia, E., Soofi, K., Wessel, P., Chandler, M., Smith, W.H.F., 2013, Towards 1 mGal global
marine gravity from CryoSat-2, Envisat, and Jason-1: The Leading Edge, v. 32, p. 892–899.
Schaer, J.-P., Rodgers, J., 1987. The Anatomy of Mountain Ranges, Princeton University Press, Princeton,
NJ, 298 p.
Schandelmeier, H., & Reynolds, P. O., 1997. Paleogeographic-Palaeotectonic Atlas of North-Eastern Africa,
Arabia, and Adjacent Areas: Late Proterozoic to Holocene, A.A. Balke Ma, Rotterdam, 160 pp.
Schettino, A., and Turco, E., 2011. Tectonic history of the western Tethys since the late Triassic, Geological
Society of America Bulletin, 123 (1/2), pp. 89-105, doi:10.1130/N30064.1
Schettino, A., and Scotese, C. R., 2005, Apparent polar wander paths for the major continents (200 Ma to
the present day): a palaeomagnetic reference frame for global plate tectonic reconstructions: Geophysical Journal
International, v. 163, p. 727–759.
Schlanger, S.O. and Jenkyns, H.C. 1976. Cretaceous oceanic anoxic events: causes and consequences.
Geologie en Mijnbouw, 55, 179–184.
221
Schneider, S.H., Thompson, S.L., and Barron, E.J., 1985. Mid-Cretaceous continental surface temperatures:
Are high CO2 concentrations needed to simulate above freezing winter conditions?, in E.T. Sundquist and W.S.
Broeker (editors), The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to Present, Geophysical
Monograph 31, pp.554-559.
Schobben, M., van de Schootbrugge, B., Wignall, P.B., 2019. Interpreting the carbon isotope record of
mass extnctions, Elements, v. 15, p.331-337.
Schoene B., Eddy, M.P., Samperton, K.M., Keller, C.B., Keller, G., Adatte, T., And Khadri, S.F.R., 2019. U-Pb
constraints on pulsed eruption of the Deccan Traps across the end-Cretaceous mass extinction, Researchm
Science, 363:862-866.
Schouten, S., Forster, A., Panoto, F. E., and Damste, J. S. S. (2007). Towards calibration of the TEX86
palaeothermometer for tropical sea surface temperatures in ancient greenhouse worlds. Organic Geochemistry,
38, 1537–1546. https://doi.org/10.1029/2004PA001041
Schouten, S., Hopmans, E.C., Forster, A., van Bruegel, Y., Kuypers, M.M., and Sinninghe, D., 2003, Extremely
high sea-surface temperatures at low latitudes during the middle Cretaceous as revealed by archael membrane lipids,
Geology, 31:1069-1072.
Schouten, S., Hopmans, E.C., Schefuß, E., Sinninghe Damsté, J.S., 2002. Distributional variations in marine
crenarchaeotal membrane lipids: a new tool for reconstructing ancient sea water temperatures? Earth Planet. Sci.
Lett. 204, 265–274.
Schuchert, C., 1955., published posthumously, Atlas of paleogeographic Maps of North America, C. O.
Dunbar & C. M. Levene, Eds., John Wiley & Sons, New York, and Chap Mann & Hall, Ltd., London, 177 pp.
Schulte, P., Alegret, L., Arenillas, I., et al., 2010. The Chicxulub asteroid impact and mass extinction at the
Cretaceous-Paleogene boundary, Science,327:1214-1218.
Sclater, J.G., Jaupart, C., and Galson, D., 1980. The heat flow through oceanic and continental crust and
the heat loss of the Earth, Reviews of Geophysics and Space Physics, v. 18, p. 269-311
Scotese, C.R., 1976. A continental drift "flip book\", Computers and Geosciences, 2:113-116.
Scotese, C.R., 1983. Plate Tectonic Reconstruction Software, PALEOMAP Fortran program (Subroutines
MAIN.FTN, RECON8.FTN, ADDER.FTN,ROTATE.FTN, MOLL.FTN), Department of Geophysical Science, University of
Chicago, 14 pp.
Scotese, C.R., 1990. Atlas of Phanerozoic Plate Tectonic Reconstructions, PALEOMAP Progress 01-1090a,
pp. 57, University of Texas at Arlington:Arlington
Scotese, C.R., 1991. Jurassic and Cretaceous Plate Tectonic Reconstructions, Palaeogeography,
Palaeoecology, and Palaeoclimatology, 87: 493-501.
Scotese, C. R., 1998. Digital Paleogeographic Map Archive on CD-ROM, PALEO MAP Project, Arlington,
Texas.
Scotese, C. R., 2001. Atlas of Earth History, Volume 1, Paleogeography, PALEOMAP Project, Arlington,
Texas, 52 pp.
Scotese, C.R., 2001. Animation of Plate Motions and Global Plate Boundary Evolution since the Late
Precambrian. Geological Society of America, 2001 Annual Meeting, Boston, (November 2-6), Abstracts with
Programs, 33(6): 85
Scotese, C. R., 2004. Cenozoic and Mesozoic Paleogeography: Changing Terrestrial Biogeographic
Pathways, in Frontiers of Biogeography: New Directions in the Geography of Nature, M. V. Lomolino & L. R.
Heaney, Eds., Sinauer Associates, Inc., Sunderland, Massachusetts, p. 1-26.
Scotese, C.R., 2008. Circum-Arctic Regional Study, Part1. A Plate Tectonic Model of the Arctic Ocean and
Its Borderlands: "Windshield Wiper" and "Rollback" Tectonics, published by the PALEMAP Project, Arlington, Texas,
98 pp.
Scotese, C. R., 2014a. Atlas of Early Cretaceous Paleogeographic Maps, PALEOMAP Atlas for ArcGIS,
volume 2, The Cretaceous, Maps 23-31, Mollweide Projection, PALEOMAP Project, Evanston, IL.
Scotese, C. R., 2014b. Atlas of Late Cretaceous Paleogeographic Maps, PALEOMAP Atlas for ArcGIS,
volume 2, The Cretaceous, Maps 16 - 22, Mollweide Projection, PALEOMAP Project, Evanston, IL
Scotese, C.R., 2014c. Atlas of Plate Tectonic Reconstructions (Mollweide Projection), Volumes 1-6,
PALEOMAP Project PaleoAtlas for ArcGIS, PALEOMAP Project, Evanston, IL.
Scotese, C.R., 2016a. A New Global Temperature Curve for the Phanerozoic, (abstract), Geological Society
of America Annual Meeting, Sunday, September 25, 2016, Session no. 74, #287, Recent Advances in
222
Paleoclimatology and Paleoceanography (Posters). Exhibit Hall E/F (Colorado Convention Center), Geological
Society of America Annual Meeting, Denver, CO.
Scotese, C.R., 2016b. Tutorial: PALEOMAP Paleoatlas for GPlates and the PaleoData Plotter Program,
<http://www.earthbyte.org/paleomap-paleoatlas-for-gplates/>.
Scotese, C.R., 2018. Plate Tectonic Evolution during the last 1.5 Billion Years: The Movie, Paper No. 93-1,
Session no. 93-T180, Plate Tectonics Paradigm and Future Directions of Research -- Discovery in Light of Its 50 -
Year History, (abstract) Geological Society of America Annual Meeting, Abstracts with Program, November 5,
Indianapolis, IN
Scotese, C.R., 2021. An Atlas of Phanerozoic Paleogeographic Maps: The Seas Come In and the Seas Go
Out. Annual Review of Earth and Planetary Sciences, 49(1), 679--728. https://doi.org/10.1146/annurev-earth-
081320-064052
Scotese, C.R., and Baker, D.W., 1975. Continental drift reconstructions and animation, Journal of
Geological Education, 23: 167-171.
Scotese, C.R., and Dammrose, R. 2008. Plate boundary evolution and mantle plume eruptions during the
last Billion Years. Geological Society of America 2008 Annual Meeting, October 5-9, 2008, Houston, TX, Abstracts
with Programs, 40(6), Abstract 233-3, 328
Scotese, C.R., and Denham, C.R., 1988. User's manual for Terra MobilisTM: Plate Tectonics for the
Macintosh®.
Scotese, C.R., and Elling, R.P., 2017. Plate Tectonic Evolution during the last 1.3 billion Years: The Movie, in
R. Butler, M. Daly, G. Roberts, J. Turner, T. Watts (convenors), Plate Tectonics at 50, William Smith Meeting, 2017,
3-5 October, The Geological Society, Burlington House, London, (abstract book), pp. 16-17.
<https://www.youtube.com/watch?v=CnVGFv-1Wqc>
Scotese, C. R., and Golonka J., 1992. Paleogeographic Atlas, PALEOMAP Progress Report 20-0692,
Department of Geology, University of Texas at Arlington, Texas, 34 p.
Scotese, C.R., and McKerrow W.S. 1990. Revised world maps and introduction. In W. S. McKerrow and C.
R. Scotese (eds.), Paleozoic Paleogeography and Biogeography, Memoir 12, pp. 1-21, London: Geological Society of
London
Scotese, C.R., and Sager, W., 1988. (ed.s), Mesozoic and Cenozoic Plate Tectonic Reconstructions,
Tectonophysics, 155, 399 pp.
Scotese, C.R., and Schettino, A., 2017. Late Permian -- Early Jurassic Paleogeography of Western Tethys
and the World, in: Soto, J.I., Flinch, J., Tari, G. (Eds.), Permo-Triassic Salt Provinces of Europe, North Africa and the
Atlantic Margins. Elsevier, pp. 57-95.
Scotese, C.R., and Van der Voo, R., 2017, A paleomagnetic database for GPlates: Paleopoles, declination
arrows, and paleolatitudes: GPlates Tutorial, 52 p.
Scotese, C. R., & Winn, K., 1987. Phanerozoic Paleogeographic Maps, Paleoceanographic Mapping Project,
POMP Progress Report 33-1287, 31 pp., UTIG Technical Report 84
Scotese, C.R., and Wright, N., 2018. PALEOMAP Paleodigital Elevation Models (PaleoDEMS) for the
Phanerozoic. https://www.earthbyte.org/paleodem-resource-scotese-and-wright-2018/.
Scotese, C.R., Gahagan, L.M., and Larson, R.L., 1988. Plate tectonic reconstructions of the Cretaceous and
Cenozoic ocean basins, in C.R. Scotese and W. Sager, (editors), Mesozoic and Cenozoic Plate Tectonic
Reconstructions, Tectonophysics, pp. 27-48.
Scotese, C.R., Illich, H., Zumberge, J, and Brown, S., and Moore, T., 2007. The GANDOLPH Project: Year
One Report: Paleogeographic and Paleoclimatic Controls on Hydrocarbon Source Rock Deposition, A Report on the
Methods Employed, the Results of the Paleoclimate Simulations (FOAM), and Oils/Source Rock Compilation,
Conclusions at the End of Year One: Cenomanian/Turonian (93.5 Ma), Kimmeridgian/Tithonian (151 Ma),
Sakmarian/Artinskian (284 Ma), Frasnian/Famennian (375 Ma), February, 2007. GeoMark Research Ltd, Houston,
Texas, 142 pp.
Scotese, C.R., Illich, H., Zumberge, J, and Brown, S., and Moore, T., 2008. The GANDOLPH Project: Year
Two Report: Paleogeographic and Paleoclimatic Controls on Hydrocarbon Source Rock Deposition, A Report on the
Methods Employed, the Results of the Paleoclimate Simulations (FOAM), and Oils/Source Rock Compilation,
Conclusions at the End of Year Two: Miocene (10Ma), Aptian/Albian (120 Ma), Berriasian/Barremian (140 Ma),
Late Triassic (220 Ma), and Early Silurian (430 Ma), July, 2008. GeoMark Research Ltd, Houston, Texas, 177 pp.
223
Scotese, C.R., Nokleberg, W.J., Parfenov, L.M., Badarch,G., Berzin, N.A., Khanchuk, A.I., Kuzmin, M.I.,
Oblenskiy, A.A., Prokopiev, A.V., Rodionov, S.M., and Yan,H.Q., 2005.Tectonic and metallogenic evolution
of northeast Asia: Key to regional understanding. In: Mineral Deposit Research: Meeting the Global
Challenge. Springer, Berlin, Heidelberg. https://doi.org/10.1007/3-540-27946-6_302
Scotese, C.R., Royer, D., Summerhayes, and C.P., Mills, B., 2022. Atmospheric CO2 during the last 540
million years, (abstract), American Geophysical Union (AGU) Annual Meeting, PP12D-0657, Monday, December 12,
9:00 AM-12:30 PM, McCormick Place, Poster Hall A, Chicago, IL.
Scotese, C.R., Snelson, S.S., and Ross, W.C., 1980. A computer animation of continental drift, J. Geomag.
Geoelectr., 32: suppl. III, 61-70.
Scotese, C.R., Song, H., Mills, B. J. W., and van der Meer, D. G., 2021. Phanerozoic paleotemperatures: The
earth's changing climate during the last 540 million years. Earth-Science Reviews, 215, 103503.
https://doi.org/10.1016/j.earscirev.2021.103503
Scotese, C.R., Totterdell, J.M., Holliday, S. and Langford, R.P, 1985. Paleogeographic Mapping Software for
the Intergraph Work Station. BMR-AMIRA Paleogeographic Project, Special Report, Bureau of Mineral Resources,
Canberra, Australia, 136 pp
Sears, J.W., Harms, T.A., and Evenchick, C.A., 2007. (editors), Whence the Mountains? Inquiries into the
Evolution of Orogenic Systems: A Volume in Honor of Raymond A. Price, Geological Society of America, Special
Paper 433, Boulder, CO, 417 p.
Segoufin, J., and Patriat, 1981. P., Reconstitutions de l’Ocean Indien accidental pour les époques des
anomalies M21,M2 et 34, paleoposition de Madagascar, Bull. Geol. Soc. France, 23:605-607.
Sellwood, B. W., and Price, G. D. (1994). Sedimentary facies as indicators of Mesozoic palaeoclimate. In J.
R. L. Allen, B. J. Hoslins, B. W. Sellwood, R. A. Spicer, and P. J. Valdes (Eds.), Palaeoclimates and their Modelling,
with special reference to the Mesozoic Era (pp. 17–25). Chapman and Hall, London.
Sellwood, B.W., and Valdes, P.J., 2006. Mesozoic climates: General circulation models and the rock record,
Sedimentary Geology 190:269–287, doi:10.1016/j.sedgeo.2006.05.013
Sellwood, B.W., Price, G.D., and Valdes, P.J., 1994. Cooler estimates of Cretaceous temperatures, Nature,
370:453-492.
Şengör, A.M.C, Altiner, D., Cin, A., Ustaomer, T., and Hsu, K., 1988. Origin and assembly of the Tethyside
orogenic collage at the expense of Gondwana Land, in M. Audley-Charles and A. Hallam (editors), Gondwana and
Tethys, Geological Society of London, Special Publication, 37: 119-181.
Şengör, A.M.C., Altiner, D., Zabci, C., Sunal, G., Lom, N, Aylan, E., and Öner, T., 2023. On the nature of the
Cimmerian Continent, Earth-Science Reviews, 247, 104520, https://doi.org/10.1016/j.earscirev.2023.104520
Şengör, A.M.C., and B. A. Natal’in (2001). Rifts of the world. Mantle plumes: Their identification through
time, 352, 389-482.
Şengör, A.M.C, and Atayman, S., 2009. The Permian Extinction and Tethys: An Exercise in GLobal Geology,
Special Paper 448, Geological Society of America, Boulder, CO, 96 p.
Şengör, A.M.C, and Natalin, B.A., 1996. Paleotectonics of Asia: fragments of a synthesis, in The Tectonic
Evolution of Asia, A. Yin and M. Harrison (eds.), Cambridge University Press, Cambridge, p. 486-640.
Şengör, A.M.C, Cin A, Rowley DB, Nie SY, 1993. Space-Time Patterns of Magmatism along the Tethysides:
A Preliminary Study, Journal of Geology, 101:51-84, https://doi.org/10.1086/648196
Şengör, A.M.C, Natal’in, B.A., van der Voo, R., and Sunal, G., 2014a. A new look at the Altaids: A
superorogenic complex in Northern and Central Asia as a factory of continental crust, Part II: Palaeomagnetic data,
reconstructions, crustal growth, and global sea-level, Austrian Journal of Earth Sciences, v. 107/2, p. 131-181.
Şengör, A.M.C, Natal’in, B.A., Sunal, G., and van der Voo, R., 2014b. A new look at the Altaids: A
superorogenic complex in Northern and Central Asia as a factory of continental crust, Part I: Geological data
compilation (exclusive of paleomagnetic observations), Austrian Journal of Earth Sciences, v. 107/1, p. 169-232.
Sepkoski Jr., J.J., 1996. Patterns of Phanerozoic extinction: a perspective from global data bases, in O.H.
Walliser (editor), Global Events and Event Stratigraphy in the Phanerozoic, Springer-Verlag, Berlin, pp. 35–51.
Servais, T., Harper, D.A.T., Kröger, B., Scotese, C.R., Stigall, A.L., and Xhen, Y.Y., 2023. Changing
palaeobiogeography during the Ordovician, in D.A.T. Harper, B. Lefebvre, I.G. Percival, T. Servais, (editors), A Global
Synthesis of the Ordovician System: Part I., Geological Society of London, Special Publications 535,
doi.org/10.114/SP535-2022-168
Seton. M., Muller. R.D., Zahirovic, S., Gaina. C., Torsvik. T., Shephard, G., Talsma, A., Gurnis. M., Turner,
224
M., Maus. S., Chandler. M., 2012. Global continental and ocean basin reconstructions since 200 Ma. Earth-Sci. Rev.
113, 212-270. https://doi.Org/10.1016/j.earscirev.2012.03.002
Seton, M., Müller, R.D., Zahirovic, S., Williams, S., Wright, N.M., Cannon, J., Whittaker, J.M., Matthews, K.,
and McGirr, R., 2020. A Global Data Set of Present-Day Oceanic Crustal Age and Seafloor Spreading Parameters,
Geochemistry, Geophysics, and Geosystems (G3), American Geophysical Union (AGU), v. 21, e2020GC009214.
https://doi.org/ 10.1029/2020GC009214
Seton, M., Williams, S.E., Domeier, M., Collins, A., and Sigloch, K., 2023. Deconstructing plate tectonic
reconstructions, Nature Reviews, Earth and Environment, 19 p.. https://doi.org/10.1038/s43017-00384-8
Sewall, J.O., van de Wal, R.S.W., van der Zwan, K., van Oosterhout, C., Dijkstra, H.A., and Scotese, C.R.,
2007. Climate model boundary conditions for four Cretaceous time slices, Climate of the Past, v. 3, pp. 647-657
Sharman, G.F., and Risch, D.L., 1988. Northwest Pacific tectonic evolution in the middle Mesozoic, in C.R.
Scotese and W.W. Sager (eds.), Mesozoic and Cenozoic Plate Reconstructions, Tectonophysics, 155: 331-344.
Siegesmund, S., Fügenschuh, B., and Froitzheim, N., 2008. Tectonic Aspects of the Alpine-Dinaride-
Carpathian System, Geological Society Special Publication, No. 298, Geological Society, London, 448 p.
Simmons, M.D., 2012. Sequence Stratigraphy and Sea-Level Change. In: Gradstein, F.M., Ogg, J.G.,
Schmitz, M.D. and Ogg, G.M. (eds.), 2012. The Geologic Time Scale 2012. Elsevier, Amsterdam, pp. 239–267.
Skelton, P., 2003. The Cretaceous World, Cambridge University Press, Cambridge, pp. 360.
Skogseid, J., Planke, S., Faleide, J.I., Pedersen, T., Eldholm, O., and Neverdal, F., 2000. NE Atlantic
continental rifting and volcanic margin formation, in A. Nottvedt (ed.), Dynamics of the Norwegian margin, Geol.
Soc. London, Special Publication 167:295-326.
Sloan, L. C. and D. Pollard, 1998. Polar stratospheric clounds: A high latitude warming mechanism in an
ancient greenhouse world, Geophysical Research Letters 25(18): 3517-3520.
Sloss, L.L., 1963. Sequences in the cratonic interior of North America, Geological Society of America
Memoir 39, pp. 91-124.
Smit, J., 1999. The global stratigraphy of the Cretaceous Tertiary boundary impact ejecta, Annual Review
of Earth and Planetary Sciences, 27:75-91, doi:10.1146/annrev.earth.27.1.75.
Smith, A.G., and Hallam, A., 1970. The Fit of the Southern Continents, Nature, 225:139-144.
Smith, A. G., Smith, D. G., & Funnell, B. M., 1994. Atlas of Mesozoic and Cenozoic Coastlines, Cambridge
University Press, 99 pp.,
Smith W. 1815. A Delineation of the Strata of England and Wales and Part of Scotland. Geological Society
of London
Smith, W. H. F., and Sandwell, D. T., 1997. Global sea floor topography from satellite altimetry and ship
depth soundings. Science, 277, 1956-1962.
Snedden J.W., Liu C. 2010. A Compilation of Phanerozoic Sea-Level Change, Coastal Onlaps and
Recommended Sequence Designations. AAPG Search Discovery Article, 40594:3
Snedden J.W., Liu C. 2011. Recommendations for a uniform chronostratigraphic designation system for
Phanerozoic depositional sequences. AAPG Bull. 95:1095-1122. http://dxdoi.org/10.1306/01031110138
Song, H., and Scotese, C.R., 2023. The end-Paleozoic Great Warming, Science Bulletin,
doi:/10.1016/j.scib.2023.09.009, 9 p.
Song, H., Wignall, P. B., Song, H., Dai, X., and Chu, D. (2019). Seawater Temperature and Dissolved Oxygen
over the Past 500 Million Years. Journal of Earth Science, 30(2), 236–243. https://doi.org/10.1007/s12583-018-
1002-2
Spakman, W., and Nolet, G., 1988. Imaging algorithms, accuracy and resolution in delay time tomography,
in Mathematical Geophysics, Springer, Netherlands, pp. 55-187.
Spencer, A.M., 1974. (editor), Mesozoic – Cenozoic Orogenic Belts: Data for Orogenic Studies, Geological
Society Special Publication 4, Scottish Academic Press, Edinburgh, 809 p.
Spicer, R. A. and A. B. Herman , 2010. The Late Cretaceous environment of the Arctic: A quantitative
reassessment based on plant fossils, Palaeogeography, Palaeoclimatology, Palaeoecology: doi:
10.1016/j.palaeo.2010.1002.1025.
Spicer, R.A., Ahlberg, A., Herman, A.B., Hofmann, C.C., Raikevich, M., Valdes, P.J., and Markwick, P.J.,
2008. The Late Cretaceous continental interior of Siberia: A challenge for climate models, Earth and Planetary
Science Letters, v. 267, p. 228-235, doi:10.1016/j.espl.2007.11.049.
Spicer, R.A., Rees, P.M., Chapman, J.L.,1994. Cretaceous phytogeography and climate signals. In: Allen,
225
J.R.L., Hoskins, B.J., Sellwood, B.W., Spicer, R.A., Valdes, P.J. (eds) Palaeoclimates and their Modelling. Springer,
Dordrecht. https://doi.org/10.1007/978-94-011-1254-3_9
Spicer, R. A., Valdes, P. J., Spicer, T. E. V., Craggs, H. J., Srivastava, G., Mehrotra, R. C., and Yang, J. (2009).
New developments in CLAMP: Calibration using global gridded meteorological data. Palaeogeography,
Palaeoclimatology, Palaeoecology, 283(1–2), 91–98. https://doi.org/10.1016/j.palaeo.2009.09.009
Spicer, R. A., Yang, J., Spicer, T. E. V., and Farnsworth, A. (2021). Woody dicot leaf traits as a palaeoclimate
proxy: 100 years of development and application. In Palaeogeography, Palaeoclimatology, Palaeoecology (Vol.
562). Elsevier B.V. https://doi.org/10.1016/j.palaeo.2020.110138
Spray, J., 2020. Earth Impact Database (EID), http://passc.net/EarthImpactDatabase/New%20website_05-
2018/Index.html
Srivastava, R.K., Ernst, R.E., Buchan, K.L., and de Kock, M., 2022. (editors), Large Igneous Provinces and
their Plumbing System, Geological Society Special Publications, No. 518, Geological Society, London, 600 p.
Srivastava, S.P., 1985. Evolution of the Eurasian Basin and its implications to the motion of Greenland
along the Nares Strait, Tectonophysics, 114:29-53.
Srivastava, S.P., and Roest, W.R., 1989. Sea-floor spreading in the Labrador Sea; a new reconstruction,
Geology, 17:1000-1003.
Srivastava, S.P., and Tapscott, C., 1986. Plate kinematics of the North Atlantic, in The Geology of North
America, volume M, The Western North Atlantic region, P.R. Vogt and B.E. Tucholke, (eds.), Geological Society of
America, Boulder, CO, pp. 379-404.
Stampfli, G. M., and Borel, G. D., 2004. The TRANSMED transects in space and time: Constraints on the
paleotectonic evolution of the Mediterranean domain. In W. Cavazza et al. (Eds.), The TRANSMED Atlas: The
Mediterranean Region from Crust to Mantle (pp. 53-80). Berlin: Springer.
Stampfli GM, Kozur HW. 2006. Europe from the Variscan to the Alpine cycles. In: European Lithosphere
Dynamics, Gee DG, Stephenson RA, ed., Memoir 32: 57-82, London: Geological Society of London
Stampfli, G. M., Mosar, J., Favre, Ph., Pillevuit, A. and Vannay, J. C., 2001. Permo-Mesozoic evolution of
the western Tethys realm: the Neo-Tethys East Mediterranean Basin connection. In A. Ziegler, W. Cavazza, A. H. F.
Robertson and S. Crasquin-Soleau (Eds.), Peri-Tethys Memoir 6; Peri-Tethyan Rift/Wrench Basins and Passive
Margins (Vol. 186, pp. 51-108). Paris: Mémoires du Muséum national d'histoire naturelle.
Stanley, S.M., 1987. Extinction, Scientific American Library, New York, 242 p.
Steinberger, B., and O’Connell, R.J., 1998. Advection of plumes in mantle flow: Implications for hot spot
motion, mantle viscosity, and plume distribution, Geophysical Journal International, 132:412-434.
Steinberger, B., and Torsvik, T.H., 2010. Toward and explanation for the present and past locations of the
poles, Geochemistry, Geophysics, and Geosystems (G3), v. 11, no. 6, p.19
Stein C.A., Stein S. 1992. A model for the global variation in oceanic depth and heat flow with lithospheric
age. Nature, 359:123-129
Stern, R.J., 2002. Subduction Zones, Reviews of Geophysics, 40:1-38, doi:10.1029/2001RG000108
Storey, M., Mahoney, J.J., and Saunders, A.D., 1997. Cretaceous Basalts in Madagascar and the Transition
between Plume and Continental Lithosphere Mantle Sources, in J.J. Mahoney and M.F. Coffin (editors), Large
Igneous Provinces: Continental, Oceanic, and Planetary Flood Volcanism, Geophysical Monograph 100, American
Geophysical Union, Washington, D.C., pp. 95-122.
Strand, T., and Kulling, O., 1972. Scandinavian Caledonides, Wiley-Interscience, London, 302 p.
Suarez, C.A., Edmonds, M., and Jones, A.P., 2019. Earth catastrophes and their impact on the Carbon
Cycle, Elements 15:301-306.
Suess, E. 1908. The Face of the Earth (Das Antlitz der Erde), v. III., Clarendon Press, Oxford, viii, 400 p.
Summerhayes, C.P.,2015. Earth’s Climate Evolution, Wiley-Blackwell, Chichester, UK, 394 pp.
Sun, Y., Joachimski, M.M., Wignall, P.B., Yan, C., Chen, Y., Jiang, H., Wang, L., and Lai, X., 2012. Lethally Hot
Temperatures During the Early Triassic Greenhouse, Science, v. 338, p. 366-370.
Sun, Y.D., Wignall, P.B., Joachimski, M.M., Bond, D.P.G., Grasby, S.E., Lai, X.L., Wang, L.N., Zhang, Z.T., Sun,
S., 2016. Climate warming, euxinia and carbon isotope perturbations during the Carnian (Triassic) Crisis in South
China, Earth and Planetary Science Letters, v. 444, p. 88-100, doi:org/10.1016/j.epsl.2016.03.037
Talwani, M., and Pitmann, W.C. III, 1977. Island Arcs, Deep Sea Trenches, and Back-Arc Basins, Maurice
Ewing Series 1, American Geophysical Union, Washingtgon, D.C., 479 p.
226
Tarduno, J.A., Sliter, W.V., Kroenke, E.L., Leckie, R.M., Mayer, H., Mahoney, J.J., Musgrave, R., Storey, M.,
and Winterer, E.L., 1991. Rapid formation of the Ontong Java plateau by Aptian mantle volcanism, Science,
254:399-403.
Tauxe. L., 2002. Paleomagnetic Principles and Practice, Modern Approaches in Geophysics, G. Nolet
(editor), v. 18, Kluwer Academic Publishers, London, UK, 299 p.
Taylor, B., 2006. The single largest oceanic plateau: Ontong Java–Manihiki–Hikurangi. Earth and Planetary
Science Letters, v. 241, p. 372–380.
Tetley, M.G., 2018. Constraining Earth’s plate tectonic evolution through data mining and knowledge
discovery, Ph. D. thesis, School of Geosciences, University of Sydney, Australia, 186 p.
Tetley, M.G., Williams, S.E., Gurnis, M., Flament, N., Müller, R.D.M., 2019. Constraining absolute plate
motions since the Triassic, Journal of Geophysical Research (JGR) Solid Earth, v. 124, 7231-7258,
https://doi.org/10.1029/2019JB017442
Tierney, J. E., and Tingley, M. P. (2014). A Bayesian, spatially-varying calibration model for the TEX86
proxy. Geochimica et Cosmochimica Acta, 127, 83–106. https://doi.org/10.1016/j.gca.2013.11.026
Tierney, J. E., and Tingley, M. P. (2015). A TEX 86 surface sediment database and extended Bayesian
calibration. Scientific Data, 2. https://doi.org/10.1038/sdata.2015.29
Torsvik TH, and Cocks LRM. 2017. Earth History and Palaeogeography. Cambridge: Cambridge University
Press, 317 p.
Torsvik, T.H., Muller, R., Van der Voo, R., and Steinberger, B., 2008, Global plate motion frames: Toward a
unified model: Reviews of Geophysics, v. 46, 44 p.
Torsvik, T.H., Svensen,H. H., Steinberger, B., Royer,D. L., Jerram, D.A., Jones, M.T, and Domeier, M., 2020.
Connecting the Deep Earth and the Atmosphere, in Mantle Convection and Surface Expression S. Cottaar et al.
(editors), AGU Monograph (in press).
Torsvik, T. H., Van der Voo, R., Preeden, U., Mac Niocaill, C., Steinberger, B., Doubrovine, P. V., van
Hinsbergen, D. J. J., Domeier, M., Gaina, C., Tovher, E., Meert, J. G., McCausland, P. J., Cocks, L. R. M., 2012.
Phanerozoic polar wander, palaeogeography and dynamics: Earth-Science Reviews, v. 114, p. 325–368.
Trotter, A.J., Williams, S.I., Nicora, A., Mazza, M., and Rigo, M., 2015, Long-term cycles of Triassic climate
change: A new d18O record from conodont apatite, Earth and Planetary Science Letters, v. 415, p. 165–174.
Uliana, M. A. and L, Leggarreta, 1993, Hydrocarbon habitat in a Triassic-to-Cretaceous Sub-Andean
setting: Neuquen Basin, Argentina; Journal of Petroleum Geology, v. 16, p. 397-420.
Ulmishek, G. F., & Klemme, H. D., 1990. Depositional Controls, Distribution, and Effectiveness of the World’s
Petroleum Source Rocks, U.S.G.S. Bulletin 1931, Denver, 59 pp.
Unternehr, P., Curie, D., Olivet, J.L., Goslin, J., Beuzart, P., 1988. South Atlantic fits and intraplate
boundaries in Africa and South America, in , in Mesozoic and Cenozoic Plate Reconstructions, C.R. Scotese and
W.W. Sager (eds.), Tectonophysics, 155:169-180.
Upchurch, G.R., Kiehl, J., Shields, C., Scherer, J., and Scotese, C.R., 2015. Latitudinal temperature gradients
and high latitude temperatures during the latest Cretaceous: Congruence of geologic data and climate models,
Geology, v. 43, no. 9, p. 683-686.
Upchurch, G.R., Otto-Bliesner, B.L., and Scotese, C.R., 1999, Terrestrial vegetation and its effects on climate
during the latest Cretaceous, in Evolution of the Cretaceous Ocean-Climate System, E. Barrera and C. Johnson,
(editors), Geological Society of America, Special Paper 332:407-426.
Upchurch, P., Hunn, C.A., and Norman, D.B., 2002. An analysis of dinosaurian biogeography: evidence for
existence of vicariance and dispersal patterns caused by geological events, Proceedings of the Royal Society of
London, series B, 269:613-621, doi: 10.1098/rspb.2001.1921
Uyeda, S., and Kanamori, H., 1979. Back-arc opening and the mode of subduction, Journal of Geophysical
Research, 84:1049-1061, doi:10.1029/JB084iB03p01049
Vaes, B., 2023., On pole position: new approaches to quantifying polar wander and relative paleomagnetic
displacements, (Ph.D. thesis), Utrecht Studies in Earth Sciences, No. 382, 269 p.
Vaes, B., van Hinsbergen, D.J.J., and Boschman, L.M., 2019. Reconstruction of subduction and back-arc
spreading in the NW Pacific and Aleutian Basin: Clues to the causes of Cretaceous and Eocene plate
reorganizations, Tectonics, v. 38, no. 4, 1367-1413.
227
Vaes, B., van Hinsbergen, DJJ, van de Lagemaat, SHA, van der Wiel, E, Lom, N, Advokaat, EL, Boschman,
LM, Gallo, LC, Greve, A, Guilmette, C, Li, SH, Lippert, PC, Montheil, L, Qayyum, A, and Langereis, CG, 2023. A global
apparent polar wander path for the last 320 Ma calculated from site-level paleomagnetic data, Earth-Science
Reviews, Volume 245, October 2023, 104547
Vai l, P. R, Mit chu m, R.M , T hom pso n, S ., 1977 a. Sei smic St rat igr aphy and Gl obal Ch ange s of Sea Level: Part 4.
Global Cycles of Relative Changes of Sea Level Vol. 165, pp. 83-97
Vai l, P. R., MItchum, RM., Todd, R.G., Widmier, J.M., Thompson, III, S., Sangree, J.B., Bubb, J.N., Hatlelid,
W. G ., 1 9 7 7 b . S e is m i c s t ra t i g r a p hy a n d g l o b al c h a n g e s i n s e a l e v e l, i n C . E . Pa y t o n ( e di t o r ) , S e i s mi c S t r a ti g r a p h y –
Applications to Hydrocarbon Exploration, American Association of Petroleum Geologists Memoir 26, pp. 49-212.
Valdes, P. J., Armstrong, E., Badger, M. P. S., Bradshaw, C. D., Bragg, F., Crucifix, M., Davies-Barnard, T.,
Day, J. J., Farnsworth, A., Gordon, C., Hopcroft, P. O., Kennedy, A. T., Lord, N. S., Lunt, D. J., Marzocchi, A., Parry, L.
M., Pope, V., Roberts, W. H. G., Stone, E. J., Williams, J. H. T. (2017). The BRIDGE HadCM3 family of climate models:
HadCM3@Bristol v1.0. Geoscientific Model Development, 10(10), 3715–3743. https://doi.org/10.5194/gmd-10-
3715-2017
Valdes, P. J., Scotese, C. R., and Lunt, D. J., 2021. Deep ocean temperatures through time. Climate of the
Past, 17(4), 1483–1506. https://doi.org/10.5194/cp-17-1483-2021
Valdes, P.J., Sellwood, B.W., and Price, G.D., 1996. Evaluating Concepts of Cetaceous Equability,
Paleoclimates, v.2, pp. 139-158.
van de Lagemaat, SHA, Kamp, PJJ, Boschman, LM Van Hinsbergen, DJJ, 2023. Reconciling the Cretaceous
breakup and demise of the Phoenix Plate with East Gondwana orogenesis in New Zealand, Earth-Science Reviews,
Volume 236, 104276
Vanden berghe, N., Hilgen, F. J ., Speijer, R.P., 2012. The Paleogene Period, Chapter 28, in The Geologic Time
Scale 2012, volume 2, F.M . Gradstein, J.G. Ogg, M.D. Schmitiz, and G.M. Ogg, (editors), p. 855-922, Elsevier,
Amsterdam.
van der Meer, D.G., Scotese, C.R., Mills, B.L.W., Sluijs, A., van den Berg van Saparoea, A-P, van de Weg,
R.M.B., 2022, Long-term Phanerozoic global mean sea level: Insights from Strontium isotope variations and
estimates of continental glaciation, Gondwana Research, v. 111, pp. 103-121.
van der Meer, D.G., van den Berg van Saparoea, A.P.H., van Hinsbergen, D.J.J., van de Weg, R.M.B.,
Godderis, Y., Le Hir, G., Donnadieu, Y., 2017a. Reconstructing first-order changes in sea level during the
Phanerozoic and Neoproterozoic using strontium isotopes. Gondwana Res. 44, 22–34.
https://doi.org/10.1016/j.gr.2016.11.002
van der Meer, D.G., van Hinsbergen, D.J.J., Spakman, W., 2018. Atlas of the underworld: Slab remnants in
the mantle, their sinking history, and a new outlook on lower mantle viscosity. Tectonophysics 723, 309–448.
https://doi.org/10.1016/j.tecto.2017.10.004
Vanderkluysen, L., and Gibson, S.A., 2015. Triggering of the largest Deccan eruptions by the Chicxulub
impact, Geological Society of America Bulletin, 127 (11-12):1507-1520,doi:10.1130/B31167.1
van der Linden, T.J.M., Dupont-Nivet, G., and van Hinsbergen, D.J.J., 2020. To wa r ds a quantitative
paleogeography calculator, (powerpoint), EGU2020-2402, session SSS8.10, May 08, 2020, 8:30-10:15.
van der Meer, D.G. van den Berg van Saparoea, A.P.H., Hinsbergen, D.J.J., van de Weg, R.M.B., Godderis,
Y., Le Hir, G., and Donnodieu, Y., 2017. Reconstructing first-order changes in sea level during the Phanerozoic and
Neoproterozoic using strontium isotopes, Gondwana Research, v. 44, p. 22-34.
van der Meer, D.G., Zeebe, R.E., van Hinsbergen, J.J., Sluijs, A., Spakman, W., and Torsvik, T., 2014. Plate
tectonic controls on atmospheric CO2 levels since the Triassic, Proceedings of the National Academy of Sciences, v.
111, no. 12, p. 4380-4385, doi/10.1073/pnas.1315657111
van der Voo, R., 1993, Paleomagnetism of the Atlantic, Tethys, and Iapetus Oceans, Cambridge University
Press, Cambridge, 411 pp.
van Hinsbergen, D.J.J., de Groot, L., van Schaik, S., Spakman, W., Bijl, P., Sluijs, A., Langereis, C., Brinkhuis,
H., 2015, A paleolatitude calculator for paleoclimate studies: PLoS ONE, v. 10, 21 p.
van Hinsbergen, D.J.J., Edwards, M.A., and Govers, R., 2009. (editors), Collision and Collapse at the Africa-
Arabia-Eurasia Subduction Zone, Geological Society Special Publication, No. 311, Geological Society, London, 368 p.
van Hinsbergen, D.J.J., Torsvik, T.H., Schmid, S.M., Matenco, L.C., Maffione, M., Vissers, R.L.M., Gürer, D.,
and Spakman, W., 2020. Orogenic architecture of the Mediterranean region and kinematic reconstruction of its
tectonic evolution since the Triassic, Gondwana Research, 81:79-229.
228
Vaughan, A.P.M., Leat, P.T., and Pankhurst, R.J., 2005. (editors). Terrane Processes at the Margins of
Gondwana, Geological Society Special Publication, No. 246, The Geological Society, London, 446 p.
Veevers, J. J., 1984. Phanerozoic Earth History of Australia, Oxford Monographs on Geology and
Geophysics, no. 2, Oxford University Press, New York, 418 pp.
Veevers, J. J., 2000. Billion-year history of Australia and neighbours in Gondwanaland, GEMOC Press,
Sydney, 388 pp.
Veikkolainen, T., Biggin, A., Pesonen, L., Evans, D., Jarboe, N., 2017, Advancing Precambrian
palaeomagnetism with the PALEOMAGIA and PINT(QPI) databases: Scientific Data, v. 4, 10 p.
Veizer, J., 1995, Reply to the Comment by L.S. Land on “Oxygen and carbon isotopic composition of
Ordovician brachiopods: Implications for coeval seawater”: Geochimica et Cosmochemica Acta, v. 59, p. 2845–
2846, https://doi .org/10.1016/0016-7037(95)00177-2.
Veizer, J., and Hoefs, J., 1976, The nature of 18O/16O and 13C/12C secular trends in sedimentary
carbonate rocks: Geochimica et Cosmochimica Acta, v. 40, p. 1387–1395, https://doi .org/10.1016/0016-
7037(76)90129-0.
Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn, F., Carden, G. A. F., Diener, A., Ebneth, S.,
Godderis, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O. G., and Strauss, H. (1999). 87Sr/86Sr, δ13C and δ18O
evolution of Phanerozoic seawater. Chemical Geology, 161(1–3), 59–88. https://doi.org/10.1016/S0009-
2541(99)00081-9
Veizer, J., and Prokoph, A. (2015). Temperatures and oxygen isotopic composition of Phanerozoic oceans.
Earth-Science Reviews, 146, 92–104. https://doi.org/10.1016/j.earscirev.2015.03.008
Vellekoop, J., Esmeraysenlet, S., Miller, K.G., Browning, J.V., Sluijs, A., De Schootbrugge, B.V., Damste,
J.S.S., Brinkhuis, H., 2016. Evidence for Cretaceous-Paleogene boundary bolide “impact winter” conditions from
New Jersey, USA. Geology, 44(8): 619-622.
Vellekoop, J., Sluijs, A., Smit., J., Schouten, S., Weijers, J.W.H., Sinninghe Damsté, and Brinkhuis, H., 2014.
Rapid short-term cooling following the Chicxulub impact at the Cretaceous-Paleogene boundary, Proceedings of
the National Academy of Science, v. 111, n. 21, pp. 7537-7541.
Vérard C. 2019. Panalesis: Towards global synthetic paleogeographies using integration and coupling
manifold models, Special Issue, “Advances in Palaeogeography, Geological Magazine, 156(2):320-330
Vérard, C., and Veizer, J., 2019. On plate tectonics and ocean temperatures. Geology, 47(9), 881–885.
https://doi.org/10.1130/G46376.1
Vérard C, Hochard C, Baumgartner O, Stampfli GM. 2015. 3D palaeogeographic reconstructions of the
Phanerozoic versus sea-level and Sr- ratio variations. Journal of Palaeogeography, 4(1):64-84
Vérard, C., Stampfli, G. M., Borel, G., Hochard, C., 2017. The Indian Promontory: A bridge between plate
tectonics and life evolution models. Universal Journal of Geoscience 5(2): 25-32,DOI: 10.13189/ujg.2017.050202
Vickers, M.L., Bajnai, D., Price, G.D., Linckens, J., and Fiebig, J., 2019. Southern high-latitude warmth
during the Jurassic–Cretaceous: New evidence from clumped isotope thermometry, Geology,
doi.org/10.1130/G46263.1
Vine, F.J., and Smith, A.G., 1981. (editors) Extensional Tectonics Associated with Convergent Boundaries,
The Royal Society, London, 442 p.
Vinogradov, A. P., Vereshchagin, V. N., Nalivkin, V. D., Ronov, A. B., Khabakov, A. V., & Khain, V. E., 1968b.
Atlas of Lithological-Paleogeographical Maps of the U.S.S.R., 1:7.500,00, Tome III, Triassic, Jurassic, and Cretaceous,
Tome III editors, V.N. Vereshchagin and A.B. Ronov, Ministry of Geology and the U.S.S.R., Academy of Sciences of
the U.S.S.R., Moscow.
Vrielynck, B., & Bouyesse, P., 2001. Le Visage Changeant de la Terre, L’eclatement se la Pangee et la
mobilite des continents au cours des desniers 250 million d’annees en 10 cartes, Commission de la Carte
Geologique du Modne, Paris, 32 pp.
Wagreich, M., Lein, R., and Sames, B., 2014. Eustasy, its controlling factors, and the lino-eustatic
hypothesis – concepts inspired by Eduard Suess, Austrian Journal of Earth Sciences, v. 107, no. 1, p. 115-131.
Wakabayashi, J., and Dilek, Y., 2021. (editors), Plate Tectonics, Ophiolites, and Societal Significance of
Geology: A Celebration of the Career of Eldridge Moores, Geological Society of America, Special Paper 552,
Boulder, CO, 296 p.
229
Walsh, D. B., 1996. Late Jurassic through Holocene Paleogeographic Evolution of the South Atlantic
Borderlands, Master’s Thesis, University of Texas at Arlington, 136 pp.
Walsh, D.B., and Scotese, C.R., 1993. PALEOMAP PC: plate tectonic reconstructions on IBM compatible
computers. Geological Society of America, Northeastern Section, 27th Annual Meeting, Abstracts with Programs, v.
25, Issue 1, pp. 44-45.
Wang, H.Z., 1985. Atlas of the Paleogeography of China, Chinese Academy of Sciences, Wuhan College of
Geology, Cartographic Publishing House, Beijing, 85 pp.
Ward, P.D., 2004. Gorgon: Paleontology, Obsession, and the Greatest Catastrophe in Earth’s History,
Viking, 257 p.
Ward, P.D., and Brownlee, D., 2003. The Life and Death of Planet Earth: How the New Science of
Astrobiology Charts the Ultimate Fate of Our World, Times Books, Henru Holt and Company, New York, 241 p.
Watkins, J.S., Feng, Z.Q., and McMillen, K.J., 1992. (editors) Geology and Geophysics of Continental
Margins, American Association of Petroleum Geologists, Memoir 53, Tulsa, OK, 419 p.
Weirzbowski, H., and Joachimski, M., 2007. Reconstruction of late Bajocian–Bathonian marine
palaeoenvironments using carbon and oxygen isotope ratios of calcareous fossils from the Polish Jura Chain
(central Poland), Palaeogeography, Palaeoclimatology, Palaeoecology, v. 254, p. 523–540,
doi:10.1016/j.palaeo.2007.07.010
Weishampel, D.B., Dodson, P., and Osmolska, H., 1990. (editors), The Dinosauria, University of California
Press, Berkeley, 733 p.
Wellnhofer, P., 1991. The Illustrated Encyclopedia of Pterosaurs, Crescent Books, New York, New York,
192 p.
Westerhold. T., Marwan, N., Drury, A.J., Liebrand, D., Agnini., C., Anagnostou, E., Barnet, J.S.K., Bohaty,
S.M., De Vleeschouwer, D., Florindo, F., Frederichs, T., Hodell, D.A., Holbourn, A.E., Kroon, D., Lauretano, V., Littler,
K., Lourens, L.L., Lyle, M., Pälike, H., Röhl, U., Tian, J., Wilkens, R.H., Wilson, P.A., Zachos, J.C., 2020. An
astronomically dated record of Earth’s climate and its predictability over the last 66 million years, Science 369:
1383-1387.
Wignall, P.B., 2015. The Worst of Times, How life on Earth Survived Eighty Million Years of Extinctions,
Princeton University Press, Princeton, NJ, 199 p.
Wignall, P.B., 2001. Large igneous provinces and mass extinctions. Earth-Science Reviews, v. 53, p.1-33.
Wilf, P., Johnson. K., and Huber, B.T., 2003. Correlated terrestrial and marine evidence for global climate
changes before mass extinction at the Cretaceous – Paleogene boundary, Proceedings of the National Academy of
Sciences, v. 100, no. 2, p. 599-604, doi /10.1073/pnas.0234701100
Willeit, M., Ganopolski, A., Robinson, A., and Edwards, N. R., XXXX, The Earth system model CLIMBER-X
v1.0 – Part 1: Climate model description and validation , Geosci. Model Dev., 15, 5905–5948,
https://doi.org/10.5194/gmd-15-5905-2022, 2022.
Wilford, G. E., 1979. Phanerozoic Paleogeography, in BMR Earth Science Atlas of Australia, E.K. Carter,
editor, Bureau of Mineral Resources, Geology & Geophysics, Australia, 31 pp.
Williams, H., 1995., (editor), Geology of the Appalachian-Caledonian Orogen in Canada and Greenland,
Geology of Canada, no. 6, Geology of North America, volume F-1, Geological Survey of Canada, Ottawa, 944 p.
Willis, K.J., and McElwain, J.C., 2002. The Evolution of Plants, Oxford University Press, Oxford, UK, 378 p.
Wilson, G.P., Clemens, W. A., Horner, J.R., and Hortman, J.H., 2014. Through the End of the Cretaceous in
the Type Locality of the Hell Creek Formation in Montana and Adjacent Areas, Special Paper 503, Geological
Society of America, Boulder, CO, 392 p.
Wilson, J. T.,1966. "Did the Atlantic Close and then Re-Open?". Nature. 211 (5050): 676–
681. Bibcode:1966Natur.211..676W. doi:10.1038/211676a0. ISSN 0028-0836. S2CID 4226266.
Wilson, J. T., 1968. "Static or Mobile Earth: The Current Scientific Revolution". Proceedings of the
American Philosophical Society. 112 (5): 309–320. ISSN 0003-049X. JSTOR 986051
Wilson, R.C.L., Whitmarsh, R.B., Taylor, B., Froitzheim, N., 2001. (editors), Non-volcanic Rifting of
Continental Margins: A Comparison of Evidence from Land and Sea, Geological Society Special Publication, No. 187,
Geological Society, London, 585 p.
Wilson, R.W., Houseman, G.A., McCaffrey, K.J., Dore, A.G., and Buiter, S.J.H., 2019. Fifty Years of the
Wilson Cycle Concept in Plate Tectonics. Geological Society, London, Special Publications, 470, 1–17. First
published online July 25, 2019, https://doi.org/10.1144/SP470-2019-58
230
Winchester, J.A., Pharaoh, T.C., Verniers, J., 2002. Palaeozoic Amalgamation of Central Europe, Geological
Society Special Publication, No. 201, The Geological Society, London, 353 p.
Windley, B.F., Alexeiev, D., Xiao, W.J., Kroner,A., and Badarch, G., 2007. Tectonic models for the accretion
of the Central Asian Orogenic Belt, Journal of the Geological Society, London, 164:31-47.
Witton, M.P., 2013. Pterosaurs: Natural History, Evolution, Anatomy, Princeton University Press,
Princeton, 291 p.
Wolfe, J. A. and G. R. Upchurch, 1987. North American nonmarine climates and vegetation during the Late
Cretaceous, Palaeogeography, Palaeoclimatology, Palaeoecology 61: 33-77.
Wortel, R., and Cloetingh, S., 1981. On the origin of the Cocos-Nazca spreading center, Geology, 9:425-
430.
Wright, J.E., and Shervais, J.W., 2008. (editors), Ophiolites, Arcs, Batholiths: A Tribute to Cliff Hopson,
Geological Society of America, Special Paper 438, Boulder, CO, 572 p.
Wright, N.M., Seton, M., Williams, S.E., Whittaker, J.M., and Müller, R.D., 2020. Sea-level fluctuations
driven by changes in global ocean basin volume following supercontinent break-up, Earth Science Reviews, v. 208,
no. 103293, 21 p., doi.org/10.1026/j.earscirev.2020.103293
Wright, N.M., Seton, M., Williams, S.E., and Müller, R.D., 2016. The late Cretaceous to recent history of
the Pacific Ocean basin, Earth-Science Reviews, v. 154, 138-173.
Yancey, T.E., and Guillemette, R.N., 2008. Carbonate accretionary lapilli in distal deposits of the Chicxulub
impact event: Geol. Soc. America Bulletin, 120:1105-1118.
Yilmaz, O., Norton, I. O., Leary, D., and Chuchla, R. J., 1996. Tectonic evolution and paleogeography of
Europe. In A. Ziegler and F. Horvath (Eds.), Peri-Tethys Memoir 2; Structure and prospects of Alpine basins and
forelands (Vol. 170, pp. 47-60 and enclosures 1-13), Paris: Mémoires du Muséum national d'histoire naturelle.
Yin, A., and Harrison, T.M., 1996. (editors), The Tectonic Evolution of Asia, Cambridge University Press,
New York, 666 p.
Zachos, J., Pagani, M., Sloan, L., Thomas, E., and Billups, K. 2001. Trends, rhythms and abberations in
global climate 65 Ma to present, Science, v. 292, p. 686-693.
Zachos, J.C., Dickens, G.R., and Zeebe, R.E., 2008. An early Cenozoic perspective on greenhouse warming
and carbon-cycle dynamics, Nature, v. 45117, p. 279-283.
Zahirovic, S., Seton, M., and Müller, R. D., (2014). The Cretaceous and Cenozoic tectonic evolution of
Southeast Asia: Solid Earth (EGU).
Zeebe RE, Zachos JC, Dickens GR. 2009. Carbon dioxide forcing alone insufficient to explain Palaeocene-
Eocene Thermal Maximum warming. Nat. Geosci. 2:576–80
Zhang, L., Wang, C., Li, X., Cao, K., Song, Y., Hu, B., Lu, D., Wang, Q., Du, X. and Cao, S. (2016). A new
paleoclimate classification for deep time. Palaeogeography, Palaeoclimatology, Palaeoecology, 443, 98–106.
https://doi.org/10.1016/j.palaeo.2015.11.041
Zhou, J., Poulsen, C.J., Pollard, D., White, T.S., 2008. Simulation of modern and middle Cretaceous marine
delta O-18 with an ocean-atmosphere general circulation model. Paleoceanography 23, 11.
Ziegler, A.M., Bambach, Parrish, J.T., Barrett, S.F., Gierlowski, E.H., Parker, W.C., Raymond, A.L., and
Sepkoski, J.J., Jr., 1981. Paleozoic biogeography and climatology, in K.J. Niklas (editor), Paleobotany, Paleoecology,
and Evolution, v2, pp. 231-266, Praeger, NY.
Ziegler, A.M., Eshel, G., Rees, P. M., Rothfus, T., Rowley, D., & Sunderlin, D. (2003). Tracing the tropics
across land and sea: Permian to present. Lethaia, 36(3), 227–254. https://doi.org/10.1080/00241160310004657
Ziegler, A.M., Raymond, A.L., Gierlowski, T.C., Horrell, M.A., Rowley, D.B., and Lottes, A.L., 1987. Coal,
climate, and terrestrial productivity: the present and early Cretaceous compared, in A.C. Scott (editor), Coal and
Coal-bearing Strata: Recent Advances, Geological Society of America Special Publication 32, pp. 25-49.
Ziegler, A.M., Rees, P.M., Rowley, D.B., Bekker, A., Li, Q., and Hulver, M.L., 1996. Mesozoic Assembly of
Asia: Constraints from fossil floras, tectonics, and paleomagnetism, in The Tectonic Evolution of Asia, A. Yin & M.
Harrison (eds.), Cambridge University Press, Cambridge, p. 371-400.
Ziegler, A. M., Rowley, D. B., Lottes, A. L., Sahagian, D. L., Hulver, M. L., & Gierlowski, T. C., 1985.
Paleogeographic interpretation: With an example from the Mid-Cretaceous. Annual Review of Earth and Planetary
Sciences, 13, 385-425.
Ziegler, A. M., Scotese, C. R., & Barrett, S. F., 1983. Mesozoic and Cenozoic paleogeographic Maps, in Tidal
friction and the Earth's Rotation II, P. Broche and J. Sunder Mann, eds., Springer-Verlag, Berlin, p.241-252.
231
Ziegler, A.M., Scotese, C.R., McKerrow, W.S., Johnson, M.E., and Bambach, R.K., Paleozoic
paleogeography, Annual Review of Earth and Planet Sci., v. 7, p. 473-302.
Ziegler PA., 1982. Geological Atlas of Western and Central Europe, pp. 130, Den Haag: Shell Internationale
Petroleum Maatschappij B.V.
Ziegler, P.A., 1988. Evolution of the Arctic-North Atlantic and the Western Tethys, American Association of
Petroleum Geologists Memoir 43, Tulsa, Oklahoma, 198 pp.
Ziegler, P.A., 1990. Geological Atlas of Western and Central Europe, Second Edition, Geological Society
Publishing House, Bath, Great Britain, 239 pp.
Ziegler, P.A., 1992. (editor), Geodynamics of Rifting, Volume I. Case History Studies of Rifts: Europe and
Asia, Selected papers of the “Geodynamics of Rifting Symposium”, organized by Working Group 3 of the Inter-
Union Commission on the Lithosphere, held at Glion-sur-Montreux, 4-11 November, 1990, Tectonophysics, v. 108,
n. 1-3, Elsevier Science Publishers, Amsterdam, 253 p.
Ziegler PA, Cavazza W, Robertson AHF, Crasquin-Soleau S. 2001. Peri-Tethyan Rift-Wrench Basins and
Passive Margins, Peri-Tethys Memoir 6, Memoires du Museum National d’Histoire Naturelle, Tome 186, Paris, 551
pp.
Ziegler PA, Horvath F. 1996, Structure and Prospects of Alpine Basins and Forelands, Peri-Tethys Memoir
2, Memoires du Museum National d’Histoire Naturelle, Tome 170, Paris, 551 pp.
Zonenshain, L.P., Kuzmin, M.I., Natapov, L.M., 1990. Geology of the USSR: A Plate Tectonic Synthesis, B.
M. Page (editor), Geodynamics Series Volume 21, American Geophysical Union, San Francisco, pp.242.
232
I. List of Figures
Figure 1. Cretaceous Timescale (Cohen et al., 2013; version 2023/06)
Figure 2. Example Cretaceous plate tectonics, paleogeographic, and paleoclimatic maps
Figure 3. Example Cretaceous Paleoclimatic Maps
Figure 4. Tectonic Elements (static polygons), Scotese (2016, 2021), see Table 2 for legend.
Figure 5. Albian/Cenomanian (100 Ma) reconstruction of tectonic elements (Scotese, 2016, 2021)
Figure 6. Plate tectonic hierarchy describing the movement of the 30 tectonic elements that make up Gondwana.
Figure 7. Lines of evidence used to produce a plate tectonic reconstruction.
Figure 8A. Tectonic Map of the World (part 1) (See Figure 9 for legend).
Figure 8B. Tectonic Map of the World (part 2) (See Figure 9 for legend).
Figure 8C. Tectonic Map of the World (part 3) (See Figure 9 for legend).
Figure 8D. Tectonic Map of the World (part 4) (See Figure 9 for legend).
Figure 9. Legend for Tectonic Map of the World.
Figure 10 Apparent Polar Wander Path of the South Pole in African coordinates during the last 320 million years.
Figure 11. Plate Tectonics as a Catastrophic System (Scotese, 2014).
Figure 12. Location of Cretaceous trenches and subduction zone graveyards.
Figure 13. Early Ordovician paleobiogeography with ocean currents.
Figure 14. Braarudisphaera bigelowii migration paths during the Maastrichtian.
Figure 15. Control points used to map the location of the paleocoastline during the early Jurassic (200 Ma)
Figure 16. Continental flooding curve.
Figure 17. Continental Flooding.
Figure 18. Average elevation of the land surface above sea level during the Phanerozoic.
Figure 19. Phanerozoic Global Sea Level derived from continental flooding.
Figure 20. A Comparison of Phanerozoic Sea Level Curves.
Figure 21. Comparison of Phanerozoic Global Average Temperature with Phanerozoic CO2 Levels.
Figure 22. Modern Köppen Belts (Burgener et al., 2023)
Figure 23. A. Mid-Cretaceous (100 Ma) and B. Early Permian (280 Ma) lithologic indicators of climate and paleo-
Köppen belts (Boucot et al., 2013).
Figure 24. Pole-to-Equator temperature gradients during the Cretaceous
Figure 25. Early Campanian (80 Ma) drainage pattern and paleorivers.
Figure 26. Drainage Basin Size (gray line) and Mean Flow Length (red lines) during the Phanerozoic
Figure 27. Plate Tectonic Tree Diagram for the Mesozoic and Cenozoic.
Figure 28. Jurassic/Cretaceous Boundary (145 Ma) , see Figure 1A for legend.
Figure 29. Early Cretaceous (latest Berriasian, 140 Ma), see Figure 1A for legend.
Figure 30. Early Cretaceous (Valanginian, 135 Ma), see Figure 1A for legend.
233
Figure 31. Early Cretaceous (early Hauterivian, 130 Ma), see Figure 1A for legend.
Figure 32. Early Cretaceous (earliest Barremian, 125 Ma), see Figure 1A for legend.
Figure 33. Early Cretaceous (early Aptian, 120 Ma), see Figure 1A for legend.
Figure 34. Early Cretaceous (late Aptian, 115 Ma), see Figure 1A for legend.
Figure 35. Early Cretaceous (early Albian, 110 Ma), see Figure 1A for legend.
Figure 36. Early Cretaceous (middle Albian, 105 Ma), see Figure 1A for legend.
Figure 37. Early Cretaceous (latest Albian, 100 Ma), see Figure 1A for legend.
Figure 38. Mid-Cretaceous (Cenomanian, 95 Ma), see Figure 1A for legend.
Figure 39. Mid-Cretaceous (latest Turonian , 90 Ma), see Figure 1A for legend.
Figure 40. Late Cretaceous (Santonian - Coniacian, 85 Ma), see Figure 1A for legend.
Figure 41. Late Cretaceous (Early Campanian, 80 Ma), see Figure 1A for legend.
Figure 42. Late Cretaceous (Late Campanian, 75 Ma) , see Figure 1A for legend.
Figure 43. Late Cretaceous (Maastrichtian, 70 Ma), see Figure 1A for legend.
Figure 44. Cretaceous/Paleogene Boundary (65 Ma), see Figure 1A for legend.
Figure 45. Cretaceous sea level derived from continental flooding.
Figure 46. The changing area of the continents during the late Neoproterozoic and Phanerozoic (0 – 750 Ma).
Figure 47. Length of mid-ocean ridges (black line) and subduction zones (dashed line) during the Cretaceous.
Figure 48. Latitudinal limits of snow (dotted line) and ice (black line) during the Phanerozoic.
Figure 49. Jurassic/Cretaceous Boundary (145 Ma), see Figure 1B for legend.
Figure 50. Early Cretaceous (latest Berriasian, 140 Ma), see Figure 1B for legend.
Figure 51. Early Cretaceous (Valanginian, 135 Ma), see Figure 1B for legend.
Figure 52. Early Cretaceous (early Hauterivian, 130 Ma), see Figure 1B for legend.
Figure 53. Early Cretaceous (earliest Barremian, 125 Ma), see Figure 1B for legend.
Figure 54. Early Cretaceous (early Aptian, 120 Ma), see Figure 1B for legend.
Figure 55. Early Cretaceous (late Aptian, 115 Ma), see Figure 1B for legend.
Figure 56. Early Cretaceous (early Albian, 110 Ma), see Figure 1B for legend.
Figure 57. Early Cretaceous (middle Albian, 105 Ma), see Figure 1B for legend.
Figure 58. Early Cretaceous (latest Albian, 100 Ma), see Figure 1B for legend.
Figure 59. Mid-Cretaceous (Cenomanian, 95 Ma), see Figure 1B for legend.
Figure 60. Mid-Cretaceous (latest Turonian , 90 Ma), see Figure 1B for legend.
Figure 61. Late Cretaceous (Santonian - Coniacian, 85 Ma), see Figure 1B for legend.
Figure 62. Late Cretaceous (early Campanian, 80 Ma), see Figure 1B for legend.
Figure 63. Late Cretaceous (late Campanian, 75 Ma), see Figure 1B for legend.
Figure 64. Late Cretaceous (Maastrichtian, 70 Ma), see Figure 1B for legend.
Figure 65. Cretaceous/Paleogene Boundary (65 Ma), see Figure 1B for legend.
Figure 66. Average river length (black line) and the number of rivers (red line) during the Cretaceous.
234
Figure 67. Early Cretaceous Drainage System and Rivers, Jurassic-Cretaceous Boundary (145 Ma
Figure 68. Early Cretaceous (latest Berriasian, 140 Ma)
Figure 69. Early Cretaceous (Valanginian, 135 Ma)
Figure 70. Early Cretaceous (early Hauterivian, 130 Ma)
Figure 71. Early Cretaceous (earliest Barremian, 125 Ma)
Figure 72. Early Cretaceous (early Aptian, 120 Ma)
Figure 73. Early Cretaceous (late Aptian, 115 Ma)
Figure 74. Early Cretaceous (middle Albian, 105 Ma)
Figure 75. Early Cretaceous (latest Albian, 100 Ma)
Figure 76. Mid-Cretaceous (Cenomanian, 95 Ma)
Figure 77. Mid-Cretaceous (latest Turonian , 90 Ma)
Figure 78. Late Cretaceous (Santonian - Coniacian, 85 Ma)
Figure 79. Late Cretaceous (Early Campanian, 80 Ma)
Figure 80. Late Cretaceous (Late Campanian, 75 Ma),
Figure 81. Late Cretaceous (Maastrichtian, 70 Ma)
Figure 82. Cretaceous/Paleogene Boundary (65 Ma)
Figure 83. Major Modern River Systems
Figure 84. Modern Surface Winds
Figure 85. Modern Surface Oceanic Circulation (source: https://en.wikipedia.org/wiki/Ocean_current)
Figure 86. Oceanic Circulation, 145 Ma, 130 Ma, 120 Ma, Valdes et al. (2021).
Figure 87. Oceanic Circulation 105 Ma, 95 Ma, 90 Ma, Valdes et al. (2021)
Figure 88. Oceanic Circulation 85 Ma, 80 Ma, 70 Ma, Valdes et al. (2021).
Figure 89. Paleo-Köppen Map for earliest Cretaceous (145 Ma; from Burgener et al., 2023)
Figure 90. Paleo-Köppen Map for Early Cretaceous (Hauterivian, 130 Ma; from Burgener et al., 2023)
Figure 91. Paleo-Köppen Map for Early Cretaceous (Aptian, 120 Ma; from Burgener et al., 2023).
Figure 92. Paleo-Köppen Map for Early Cretaceous (Albian, 105 Ma; from Burgener et al., 2023).
Figure 93. Paleo-Köppen Map for mid-Cretaceous (Cenomanian, 95 Ma; from Burgener et al., 2023).
Figure 94. Paleo-Köppen Map for mid-Cretaceous (Turonian , 90 Ma; from Burgener et al., 2023).
Figure 95. Paleo-Köppen Map for Late Cretaceous (Santonian - Coniacian, 85 Ma; from Burgener et al., 2023).
Figure 96. Paleo-Köppen Map for Late Cretaceous (Campanian, 80 Ma; from Burgener et al., 2023).
Figure 97. Paleo-Köppen Map for Late Cretaceous (Maastrichtian, 70 Ma; from Burgener et al., 2023).
Figure 98. Global Average Temperature (GAT) during the Mesozoic (Scotese et al., 2021).
Figure 99. Phanerozoic Heat Map.
Figure 100. Tropical, Global Average, Deep Ocean, and Polar Temperatures during the Cretaceous
235
Figure 101. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (Jurassic – Cretaceous
Boundary 145 Ma)
Figure 102. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures. (Hauterivian, 130 Ma).
Figure 103. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (early Aptian, 120 Ma).
Figure 104. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (middle Albian, 105 Ma).
Figure 105. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (late Cenomanian, 95
Ma).
Figure 106. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (latest Turonian, 90 Ma).
Figure 107. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (Coniacian – Santonian,
85 Ma).
Figure 108. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (early Campanian, 80
Ma).
Figure 109. Cretaceous Pole-to-Equator Temperature Gradient and Global Temperatures, (early Maastrichtian, 70
Ma).
Figure 110. Comparison of Modern (top) and mid-Cretaceous Patterns of Precipitation (bottom).
Figure 111. Cretaceous precipitation during the early Berriasian (145 Ma), Hauterivian (130 Ma), and early Aptian
(120 Ma), Valdes et al., 2021.
Figure 112. Cretaceous precipitation during the late Albian (105 Ma), Cenomanian (95 Ma), Turonian-Coniacian (90
Ma), Valdes et al., 2021.
Figure 113. Cretaceous precipitation during the Santonian and Coniacian (85 Ma), Campanian (80 Ma), and
Maastrichtian (70 Ma), Valdes et al., 2021.
Figure 114. Comparison of rainfall prediction, Burgener et al. (2023) (top cell) versus the HadCM3 model (bottom
cell).
Figure 115. Precipitation results with the largest mismatch. (top) Aptian-Albian (120 Ma), (bottom) Berriasian -
Valanginian (145 Ma).
Figure 116. Earliest Cretaceous (145 Ma) Polar Orthographic projection illustrating actual size of south polar icecap
(~ 4 million km2).
Figure 117. Connections within the Earth System
Figure 118. Earth System History Machine
236
J. Appendix
Table 1. Cretaceous Time Intervals
Map
Number
Stratigraphic Stage (Plate Tectonic Reconstruction Age)
1
Cretaceous - Paleogene Boundary (65 Ma)
2
Late Cretaceous (Maastrichtian, 70 Ma)
3
Late Cretaceous (late Campanian, 75 Ma)
4
Late Cretaceous (early Campanian, 80 Ma)
5
Late Cretaceous (Coniacian - Santonian, 85 Ma)
6
Mid-Cretaceous (latest Turonian , 90 Ma)
7
Mid-Cretaceous (Cenomanian, 95 Ma)
8
Early Cretaceous (latest Albian, 100 Ma)
9
Early Cretaceous (middle Albian, 105 Ma)
10
Early Cretaceous (early Albian, 110 Ma)
11
Early Cretaceous (late Aptian, 115 Ma)
12
Early Cretaceous (early Aptian, 120 Ma)
13
Early Cretaceous (earliest Barremian, 125 Ma)
14
Early Cretaceous (Hauterivian, 130 Ma)
15
Early Cretaceous (Valanginian, 135 Ma)
16
Early Cretaceous (latest Berriasian, 140 Ma)
17
Jurassic - Cretaceous Boundary (145 Ma)
237
Table 2. PALEOMAP Tectonic Elements
Plate_ID
North AmericanTectonic Elements
Plate_ID
Chinese and SE Asian Tectonic
Elements
101
North American craton
601
Tarim
102
Greenland
602
Qidam block
103
North Slope-South Annui block
603
Okinawa Trough
104
Mexico
604
N. China Craton
105
Baja California
605
Taiwan
106
Arctic Islands
606
Bai Shan block
107
Grand Banks
607
N. South China Sea
108
West Avalonia
610
Japan
109
Florida-Piedmont
611
Yangtze
110
Alpha Ridge
612
Qiang Tang block
111
Mendeleyev Ridge
613
Lhasa block
112
Chukchi Cap
614
Song Pan Ganzi
113
Northwind Escarpment
615
Indochina
114
Lomonosov Ridge
616
Sibumasu
121
Angayucham
618
E. Andaman Sea
123
E. Yukon Tanana
619
Gulf of Thailand/ E. Malaysia/E.
Sumatra
124
North American Cordillera
620
Borneo & Java
125
N. Wrangellia
621
S. South China Sea
126
S. Wrangellia
622
Palawan & Sulu Sea
127
Stikine
623
Celebes Sea
129
Columbia Embyament
624
C. Burma Accretion
130
Sierra Nevada
625
W. Andaman Sea
131
Western Basin & Range
626
Sumatra-Java Trench & Banda Arc
132
Eastern Basin & Range
628
Amuria
133
Colorado Plateau
631
SE Japan
134
Southern Basin & Range
632
Kurile Is.
135
Gulf Coast
650
N. Parece Vela Basin
137
"northernmost Gulf of Mexico"
651
Bonin Arc, N. Mariana Basin
140
NE Canada Basin
652
S. Mariana Basin
141
Makarov Basin
653
S. Parece Vela Basin
654
W. Parece Vela Basin
South American Tectonic Elements
655
N. Philippine
201
South America Craton
656
S. Philippine
202
Rio de la Plata
659
Philippines
203
NW South America
664
N. Sulawesi
204
Chortis
670
N. Caroline
205
Yucatan
671
Molucca Sea
206
Cuba
672
Halmahera
216
Cayman Ridge
673
S. Caroline
217
West Cayman Trough
675
N. Bismarck Sea
218
East Cayman Trough
676
Bismarck Sea
221
Rosalind Bank
679
S. Sulawesi & Banda Sea
222
Jamaica
680
Timor
238
223
Quinto Sueno
682
Buru & Seram
224
Caribbean plate
684
E. Sulawesi & Molucca Sea
230
Panama arc
237
Puerto Rico
Plate_ID
African Tectonic Elements
239
Lesser Antilles
701
SC Africa
251
"southern GOM oceanic crust"
702
Madagascar
252
Hispaniola
704
Seychelles
257
Yucatan Basin
705
Mascarene Bank
290
Salado sub plate
706
Nazareth Bank
291
Patagonia
707
Morocco, Atlas Mts.
708
Danakil
Plate_ID
European Tectonic Elementrs
709
Somalia block
301
Russian platform
712
Lake Victoria block
303
Northern Highlands
714
NW Africa
304
Iberia
715
NE Africa
305
Variscan Europe
716
Sudan block
306
Corsica & Sardinia
307
Apulia
Australian-Antarctic Tectonic
Elements
308
Balkans
800
New Guinea
309
W. Svalbard
801
Australian craton
310
Barents Sea
802
East Antarctic craton
313
Midland Valley
803
Palmer Peninsula
315
East Avalonia
804
Marie Byrdland
317
E. Rockall Bank
805
Ellsworth Mt.
318
W. Rockall Bank
806
North Island, New Zealand
319
Moesia
807
South Island, New Zealand
320
Balearic Is. Menorca
808
Gulf of Papua
321
Alboran Sea
809
Rennell Ridge
322
Gulf of Gascogne transitional crust
810
N. Coral Sea
323
Malta -Gulf of Gabes
811
S. Shetland Is.
324
Faroes
812
S. Orkney Is.block
325
West Rockall Gap
813
Chatham Rise
326
East Rockall Gap
814
Campbell Plateau (New Zealnd
Plateau)
327
Voring Plateau
815
South Fiji Basin
328
Cantabrian transitional crust
816
Balleny Is.
330
Crete Aschett 324
817
South Solomon Sea
331
Cyprus
818
North Solomon Sea
332
Porcupine Bank
819
Melanesian Border Plateau
350
Yermak Plateau
820
Wharton Basin
821
Lord Howe Rise Stretched Crust
Siberian Tectonic Elements
822
S. Coral Sea
401
Siberian Craton
823
East Tasman Sea
402
Kazakhstania
824
New Hebrides Basin & E. Coral Sea
Basin
403
Omolon
825
W. Fiji Basin
405
Verkhoyansk Mt- Kolyma
826
Coral Sea Plateau
406
Kamchatka / Koni-Murgal Arc
829
Solomon Arc
407
Tien Shan- Junghar
830
Vanuatu Arc
239
408
North Sea of Okhotsk
833
Lord Howe Rise
409
C. Sea of Okhotsk
834
Norfolk Ridge
410
Sea of Okhotsk
840
Marion Reef
413
Kamchatka Trench
846
E. Fiji Basin
414
Kronotsky Terrane
847
W. Havre Trough, Colville Ridge
415
Shirshov Ridge
848
E. Havre Trough, Kermadec Trench
416
Alaskan Aleutians
870
Great Australian Bight
417
Bowers Ridge
871
Tasman Rise
418
Komandorsky - West Aleutian island
arc
872
Naturaliste Plateau
419
Koryak
873
NW Australian Basin & Exmouth
Plateau
420
Bering Sea
875
Timor Sea
876
Cuvier - Perth Abysssal Plains
Plate_ID
Indian Tectonic Elements
877
NE Indian Ocean - Argo Abyssal
Plain
501
India
888
Broken Ridge
502
Sri Lanka
895
Berkner Is.& Ronne Ice Shelf
503
Arabia
896
Sentinel Range
504
Turkey
897
Trans-Antarctic Range
505
Lut Block
506
Sistan block
Plate_ID
Oceanic Tectonic Elements
507
Farah Block
901
Pacific Plate
508
Isreal
902
Nazca Plate
510
Pontides
904
"Aluk"
511
Black Sea
905
Cape Basin
512
Sanandaj-Sirjan Zone
906
Henry Hudson plate
513
South Caspian
907
E. Jan Meyen
514
North Caspian
908
W. Jan Mayen
515
Makran Accretion
909
Cocos Plate
516
Laxmi Ridge
910
Juan de Fuca
550
Erzurum
911
Bauer
912
N. Scotia Sea
913
S. Scotia Sea
916
Kula
930
South Georgia Island
931
W. Scotia Sea
932
E. Scotia Sea
940
Easter Island microplate
944
"a little bit of Aluk or Phoenix"
985
N. Kerguelen Plateau
992
S. Kerguelen Plateau
995
Kula
996
Izanagi
997
Phoenix
998
Farallon
240
Table 3. Finite Rotations Used to Assembly Gondwana
Moving_Plate
Age
Latitude
Longitude
Angle
Fixed Plate
201
0.0
90.0
0.0
0.0
701
201
1.0
60.0
-39.0
0.3
701
201
2.0
60.0
-39.0
0.5
701
201
3.6
60.0
-39.0
0.8
701
201
5.2
60.0
-39.0
1.2
701
201
6.6
60.0
-39.0
1.8
701
201
8.2
60.0
-39.0
2.3
701
201
9.0
60.0
-39.0
2.8
701
201
9.7
60.0
-38.9
3.1
701
201
20.1
58.1
-37.4
7.0
701
201
26.6
57.2
-35.3
10.0
701
201
33.5
56.6
-33.9
13.4
701
201
43.8
57.6
-32.1
17.6
701
201
49.7
59.3
-31.6
20.1
701
201
56.4
61.1
-31.5
22.3
701
201
65.0
63.2
-33.2
24.4
701
201
70.0
63.7
-33.7
26.4
701
201
75.0
63.6
-33.3
29.4
701
201
80.0
62.5
-34.0
31.5
701
201
85.0
60.5
-33.3
34.3
701
201
118.7
51.7
-33.1
52.7
701
201
121.0
50.1
-31.9
53.4
701
201
143.8
46.9
-30.5
57.5
701
201
206.0
46.9
-30.5
57.5
701
202
0.0
90.0
0.0
0.0
201
202
83.0
90.0
0.0
0.0
201
202
105.0
90.0
0.0
0.0
201
202
143.8
-21.7
-63.9
1.9
201
202
206.0
-21.7
-63.9
1.9
201
203
0.0
90.0
0.0
0.0
201
203
56.4
90.0
0.0
0.0
201
203
94.0
-25.6
79.8
2.6
201
203
206.0
-25.6
79.8
2.6
201
290
0.0
90.0
0.0
0.0
202
290
105.0
90.0
0.0
0.0
202
290
143.8
-34.5
-103.1
1.4
202
290
206.0
-34.5
-103.1
1.4
202
291
0.0
90.0
0.0
0.0
290
241
291
105.0
90.0
0.0
0.0
290
291
143.8
43.8
162.7
1.1
290
291
206.0
43.8
162.7
1.1
290
307
0.0
90.0
0.0
0.0
714
307
50.0
90.0
0.0
0.0
714
307
65.0
90.0
0.0
0.0
714
307
70.0
53.9
126.9
-0.8
714
307
75.0
17.9
17.3
5.0
714
307
80.0
32.1
11.6
10.1
714
307
90.0
43.1
17.7
6.4
714
307
100.0
42.3
-3.4
3.8
714
307
120.0
56.8
80.2
-1.5
714
307
130.0
45.0
7.6
-5.5
714
307
140.0
35.9
0.5
-7.3
714
307
150.0
-26.0
176.4
8.1
714
307
160.0
33.7
106.6
2.5
714
307
180.0
-1.0
136.8
4.1
714
307
306.0
-41.1
-177.0
17.3
714
320
0.0
90.0
0.0
0.0
304
320
6.0
90.0
0.0
0.0
304
320
16.0
38.2
-0.8
23.3
304
320
206.0
38.2
-0.8
23.3
304
501
0.0
90.0
0.0
0.0
802
501
9.7
13.1
36.1
-6.6
802
501
20.0
17.1
28.6
-11.8
802
501
30.0
14.9
33.1
-19.5
802
501
40.0
17.3
27.2
-24.4
802
501
50.0
14.0
21.8
-32.1
802
501
60.0
11.7
16.5
-43.7
802
501
65.0
10.0
15.0
-49.0
802
501
70.0
9.9
12.5
-53.3
802
501
83.0
-9.2
-168.0
64.0
802
501
83.0
-17.8
-154.7
55.3
702
501
85.0
-18.6
-155.0
56.2
702
501
90.0
-18.0
-156.1
58.2
702
501
100.0
-19.3
-156.3
58.8
702
501
166.0
-20.2
-156.8
58.1
702
501
206.0
-22.3
-159.2
56.8
702
502
0.0
90.0
0.0
0.0
501
502
112.0
90.0
0.0
0.0
501
502
120.0
9.8
80.3
-29.8
501
502
130.2
-10.2
-98.6
18.4
501
242
502
206.0
-10.2
-98.6
18.4
501
503
0.0
90.0
0.0
0.0
715
503
9.7
36.5
18.0
-3.3
715
503
24.0
36.5
18.0
-5.2
715
503
33.0
-35.3
-158.4
8.5
715
503
206.0
-35.3
-158.4
8.5
715
504
0.0
90.0
0.0
0.0
510
504
13.5
-29.4
-156.7
5.2
510
504
33.5
-20.5
-148.2
10.8
510
504
33.5
-13.4
-118.5
8.0
715
504
47.9
-28.2
-124.8
9.9
715
504
143.9
42.6
17.4
-21.5
715
504
206.0
42.6
17.4
-21.5
715
508
0.0
90.0
0.0
0.0
715
508
9.7
36.8
31.3
-2.6
715
508
24.0
36.8
31.3
-2.6
715
508
33.0
36.8
31.3
-3.6
715
508
206.0
36.8
31.3
-3.6
715
702
0.0
90.0
0.0
0.0
709
702
112.0
90.0
0.0
0.0
709
702
125.1
10.0
-113.1
8.0
709
702
136.0
10.0
-113.1
18.1
709
702
166.0
-0.8
-88.1
20.3
709
702
206.0
-0.9
-92.0
24.4
709
704
0.0
90.0
0.0
0.0
702
704
30.0
44.1
-23.1
-0.8
702
704
40.0
11.3
-29.7
-0.6
702
704
60.0
38.1
-22.1
-3.3
702
704
70.0
24.9
7.8
-7.1
702
704
80.0
25.8
-3.9
-9.8
702
704
206.0
25.8
-3.9
-9.8
702
709
0.0
90.0
0.0
0.0
503
709
2.0
26.5
21.5
0.6
503
709
5.2
26.5
21.5
1.6
503
709
8.2
26.5
21.5
2.8
503
709
24.0
24.6
17.0
6.2
503
709
33.0
-23.4
-154.1
-10.2
503
709
206.0
-23.4
-154.1
-10.2
503
712
0.0
90.0
0.0
0.0
701
243
712
30.0
-17.8
37.1
2.5
701
712
206.0
-17.8
37.1
2.5
701
714
0.0
90.0
0.0
0.0
701
714
60.0
33.2
-114.8
0.4
701
714
84.0
90.0
0.0
0.0
701
714
110.0
90.0
0.0
0.0
701
714
118.7
22.0
11.6
2.2
701
714
130.2
17.9
11.0
5.0
701
714
143.8
22.0
11.6
6.0
701
714
206.0
22.0
11.6
6.0
701
715
0.0
90.0
0.0
0.0
701
715
110.0
90.0
0.0
0.0
701
715
118.7
-18.0
38.4
1.9
701
715
206.0
-18.0
38.4
1.9
701
716
0.0
90.0
0.0
0.0
701
716
110.0
90.0
0.0
0.0
701
716
118.7
-17.8
37.1
1.9
701
716
143.8
-17.8
37.1
2.4
701
716
206.0
-17.8
37.1
2.4
701
800
0.0
90.0
0.0
0.0
801
800
20.0
24.1
-44.1
17.6
801
800
206.0
24.1
-44.1
17.6
801
801
0.0
90.0
0.0
0.0
501
801
34.9
90.0
0.0
0.0
501
801
41.5
90.0
0.0
0.0
501
801
41.5
16.6
29.9
-23.6
802
801
43.8
15.1
31.3
-24.5
802
801
53.3
12.5
31.7
-25.2
802
801
68.7
8.7
33.2
-25.8
802
801
75.0
7.7
34.3
-25.4
802
801
79.1
6.2
35.1
-26.4
802
801
83.0
4.9
35.8
-26.8
802
801
85.0
2.5
37.9
-28.4
802
801
96.0
9.2
-135.9
33.8
802
801
206.0
9.2
-135.9
33.8
802
\
802
0.0
90.0
0.0
0.0
701
802
9.7
6.4
-54.1
1.5
701
802
20.1
7.4
-44.9
2.6
701
802
33.5
9.1
-36.4
5.5
701
802
40.1
10.9
-41.9
7.0
701
244
802
43.8
11.4
-43.7
7.8
701
802
47.9
10.3
-42.9
8.8
701
802
53.3
6.7
-40.6
10.0
701
802
57.9
3.8
-39.7
10.6
701
802
63.6
0.6
-39.2
11.3
701
802
64.7
-0.4
-39.4
11.6
701
802
68.7
1.1
-41.6
11.8
701
802
71.3
-1.8
-41.4
13.5
701
802
79.1
-4.7
-39.7
16.0
701
802
83.0
-2.0
-39.2
17.9
701
802
83.0
9.2
12.0
64.0
501
802
84.0
8.8
11.3
63.9
501
802
90.0
5.9
10.0
70.6
501
802
96.0
5.9
7.1
77.7
501
802
118.7
0.2
8.8
86.4
501
802
136.0
-2.7
13.7
90.4
501
802
166.0
-5.2
17.4
94.2
501
802
1100.0
-5.2
17.4
94.2
501
803
0.0
90.0
0.0
0.0
804
803
90.0
67.7
150.4
3.2
804
803
95.0
72.6
83.6
-6.8
804
803
100.0
47.4
-177.0
0.5
804
803
110.0
67.6
104.1
-7.8
804
803
120.0
73.8
102.7
-17.0
804
803
130.0
74.8
38.4
-15.9
804
803
135.0
73.7
6.2
-14.6
804
803
140.0
75.8
42.5
-20.8
804
803
150.0
74.6
23.0
-17.2
804
803
160.0
74.4
32.7
-19.7
804
803
170.0
69.3
23.9
-19.3
804
803
180.0
73.4
27.8
-21.7
804
803
190.0
77.3
33.2
-24.2
804
803
200.0
75.6
53.1
-27.9
804
803
210.0
70.7
18.4
-25.2
804
804
0.0
90.0
0.0
0.0
802
804
60.0
69.9
71.6
-2.2
802
804
70.0
40.7
-154.6
0.2
802
804
80.0
57.0
24.3
1.0
802
804
90.0
75.2
90.7
1.2
802
804
100.0
77.1
-144.0
-1.5
802
804
110.0
70.1
18.6
2.3
802
804
120.0
56.0
1.6
4.1
802
804
130.0
52.4
15.1
9.9
802
804
200.0
49.2
14.9
8.7
802
245
804
206.0
49.2
14.9
8.7
802
805
0.0
90.0
0.0
0.0
803
805
206.0
90.0
0.0
0.0
803
811
0.0
90.0
0.0
0.0
803
811
3.8
-63.9
-68.0
-9.1
803
811
206.0
-63.9
-68.0
-9.1
803
813
0.0
90.0
0.0
0.0
814
813
84.0
90.0
0.0
0.0
814
813
90.0
47.5
-7.1
20.1
814
813
206.0
47.5
-7.1
20.1
814
814
0.0
90.0
0.0
0.0
804
814
5.2
66.2
-96.5
4.1
804
814
9.7
75.4
-76.5
9.5
804
814
25.2
73.5
-64.9
20.2
804
814
28.7
74.2
-61.4
23.2
804
814
33.5
74.7
-57.0
27.9
804
814
40.1
75.1
-51.3
32.6
804
814
47.9
74.2
-50.9
36.9
804
814
56.4
70.1
-63.6
37.9
804
814
63.6
70.4
-58.0
45.2
804
814
65.0
69.1
-46.4
49.5
804
814
70.0
68.3
-55.3
48.7
804
814
75.0
63.4
-51.3
51.1
804
814
80.0
60.5
-61.2
46.7
804
814
85.0
60.5
-64.5
48.2
804
814
90.0
64.9
-51.0
61.0
804
814
94.0
67.1
-55.9
63.2
804
814
206.0
67.1
-55.9
63.2
804
833
0.0
90.0
0.0
0.0
801
833
40.1
90.0
0.0
0.0
801
833
56.4
5.3
-24.1
0.7
801
833
64.7
9.7
-38.8
7.3
801
833
79.1
9.8
-33.9
16.5
801
833
94.0
9.2
-22.7
23.5
801
833
1100.0
9.2
-22.7
23.5
801
834
0.0
90.0
0.0
0.0
833
834
94.0
90.0
0.0
0.0
833
834
130.0
-46.8
168.1
8.1
833
834
1100.0
-46.8
168.1
8.1
833
246
247
248
Table 5 Apparent Polar Wander Paths for Major Continents
Africa (701)
North America
(101)
South America
(201)
Eurasia (301)
Age
Longitude
Latitude
Longitude
Latitude
Longitude
Latitude
Longitude
Latitude
0
0
90.0
0
90.0
0
90.0
0
90.0
10
-105.9
89.4
-151.4
89.4
37.7
88.9
62.6
89.5
20
-113.4
88.5
-156.9
88.6
42.2
87.7
89.6
89.3
30
-126.0
86.6
-171.9
87.6
57.8
87.0
115.6
88.5
40
-128.0
83.8
-168.3
85.6
77.6
87.1
151.1
87.6
50
-127.5
79.9
-161.6
82.3
129.0
87.4
179.2
85.8
60
-128.5
77.7
-159.5
79.5
153.0
86.4
-174.4
84.2
70
-126.6
74.9
-157.6
77.0
172.8
85.2
-158.9
82.0
80
-125.7
71.1
-164.1
75.8
169.5
83.7
-161.4
79.9
90
-126.5
70.3
-179.4
77.3
127.4
83.7
-178.6
80.0
100
-124.9
66.7
164.9
76.9
117.7
82.2
170.2
78.3
120
-100.0
56.9
177.0
78.0
-72.2
89.3
-167.3
80.0
140
-103.4
48.7
172.8
70.2
-158.1
82.7
-158.6
71.7
160
-110.7
58.4
125.5
71.4
92.8
83.4
159.6
71.8
180
-108.4
64.2
96.6
69.8
57.9
80.0
135.7
70.4
200
-120.5
68.5
81.6
64.6
64.1
73.7
120.5
65.1
220
-129.6
59.1
96.3
58.4
101.1
73.7
135.7
58.9
240
-127.1
48.8
111.4
54.9
137.0
72.4
151.1
55.5
India (501)
Australia (801)
Antarctica (802)
Consensus
APW
Africa
(701)
Age
Longitude
Latitude
Longitude
Latitude
Longitude
Latitude
Longitude
Latitude
A95
0
0
90.0
0
90.0
0
90.0
0
-90.0
-
10
-44.5
83.0
-44.5
83.0
16.8
88.9
-4.5
-86.7
1.6
20
-49.5
78.2
-49.5
78.2
19.9
88.7
3.7
-85.0
2.0
30
-50.5
72.7
-50.5
72.7
48.6
88.7
17.2
-83.2
2.2
40
-54.1
67.3
-54.1
67.3
17.8
89.3
25.6
-80.5
2.5
50
-64.3
59.1
-57.1
64.8
-97.9
88.9
34.1
-77.3
2.2
60
-70.1
48.4
-58.1
64.3
-122.8
88.6
39.9
-74.5
2.0
70
-73.5
35.7
-58.7
62.4
-102.9
87.3
48.3
-72.8
1.4
80
-73.2
26.8
-57.4
61.5
-98.7
87.1
53.2
-70.9
1.0
90
-72.8
24.2
-39.4
60.9
65.8
84.5
59.2
-69.3
1.4
100
-74.7
20.3
-30.6
59.5
71.4
80.0
64.3
-66.2
2.0
120
-65.7
5.7
-28.5
47.5
18.9
74.7
80.3
-57.0
2.1
249
140
-62.8
-6.3
-34.2
42.2
-4.0
72.9
80.9
-50.7
3.1
160
-56.6
8.1
-12.2
47.4
39.9
66.1
78.9
-55.8
3.1
180
-49.8
12.9
-0.5
44.6
45.7
57.9
76.7
-62.5
2.2
200
-49.1
18.5
8.2
46.9
55.8
54.5
62.7
-67.1
2.4
220
-58.7
15.1
-2.5
54.0
60.7
64.0
53.9
-58.7
2.5
240
-65.7
7.3
-20.2
55.5
54.8
74.2
56.0
-47.9
2.8
Table 5. Apparent Polar Wander Paths for Major Continents according to the PALEOMAP
Global Plate Model (North Poles). The Consensus APW is the mean Apparent Polar Wander
path based on multiple estimates (South Pole, see Figure 10).
250
Table 6. Continental
Tectonics
A.Continental rifting
and the formation of
passive margins
Date
Author
Title
1965
E. Bullard,
J.E.Everett,
and A.G.
Smith
The Fit of the Continents Around the Atlantic
1971
A.G. Smith
and A. Hallam
The Fit of the Southern Continents
1974
C.A. Burk and
C.L. Drake
The Geology of Continental Margins
1974
J.F. Dewey
and K.C.A.
Burke
Hot Spots and Continental Break-up: Implications for Collisional Orogeny
1981
P. Barker , I.
Hill, and M.
Osmaston
Back-arc extension in the Scotia Sea
1983
P. Morgan
and B.H.
Baker
Processes of Continental Rifting
1985
S.P. Srivastava
Evolution of the Eurasian Basin and its implications to the motion of Greenland along the Nares Strait
1986
S.P. Srivastava
and C.
Tapscott
Plate kinematics of the North Atlantic
1987
L.A. Lawver
and C.R.
Scotese
A revised reconstruction of Gondwanaland
1988
J.D. Fairhead
Mesozoic plate tectonic reconstructions of the South Atlantic Ocean: the role of the West and Central African Rift System
1988
C. Mc. Powell,
S.R. Roots,
and J.J.
Veevers
Pre-breakup continental extension in East Gondwanaland and the early opening of the eastern Indian Ocean
1988
P. Unternehr,
P. Curie, J.L.
Olivet, J.
Goslin, and P.
Beuzart
South Atlantic fits and intraplate boundaries in Africa and South America
1989
J.D. Edwards
and P.A.
Santogrossi
Divergent/Passive Margin Basins
1989
S.P. Srivastava
and W.R.
Roest
Sea-floor spreading in the Labrador Sea; a new reconstruction
1991
A. Salvador
The Gulf of Mexico
1992
P.A. Ziegler
Geodynamics of Rifting, Volume I. Case History Studies of Rifts: Europe and Asia
1992
J.S. Watkins,
Feng, Z.Q.,
and K.J.
McMillen
Geology and Geophysics of Continental Margins
1998
C. Gaina, R.D.
Müller, J.Y.
Royer, J.
The tectonics history of the Tasman Sea: a puzzle with 13 pieces
251
Stock, J.
Hardebeck,
and P.A.
Symonds
2000
J. Skogseid, S.
Planke, J.I.
Faleide, T.
Pedersen, O.
Eldholm, and
F. Neverdal
NE Atlantic continental rifting and volcanic margin formation
2001
A.M.C. Sengor
and B.A.
Natal'in
Rifts of the world. Mantle plumes: Their identification through time
2001
G.M. Stampfli,
J. Mosar, P.
Favre, A.
Pillevuit, and
J.C. Vanay
Permo-Mesozoic evolution of the western Tethys realm: the Neo-Tethys East Mediterranean Basin connection
2001
R.C.L. Wilson,
R.B.
Whitmarsh, B.
Taylor, and N.
Froitzheim
Non-volcanic Rifting of Continental Margins: A Comparison of Evidence from Land and Sea
2002
M.A. Menzies,
S.L.
Klemperer,
C.J. Ebinger,
and J. Baker
Volcanic Rifted Margins
2002
R.W. Renault
and G.M.
Ashley
Sedimentation in Continental Rifts
2008
D.C. Bradley
Passive margins through Earth history
2008
H. Johnson,
T. Dore, R.W.
Gatliff, R. W.
Holdsworth,
E.R. Lundin.
and J.D.
Ritchie
The Nature and Origin of Compression in Passive Margins
2011
L. Beccaluva,
G. Bianchini,
and M.
Wilson
Volcanism and Evolution of the African Lithosphere
2012
D.G. Roberts
and A.W.
Bally
Regional Geology and Tectonics :Phanerozoic Rift systems and Sedimentary Basins
2012
D.G. Roberts
and A.W.
Bally
Regional Geology and Tectonics: Phanerozoic Passive Margins, Cratonic Basins, and Global Tectonic Maps
B. Subduction along
Andean-type margins
and continental
volcanic arcs
Date
Author
Ttile
1970
J.M. Bird and
J.F. Dewey
Lithosphere plate-continental margin tectonics and the evolution of the Appalachian orogen
252
1975
W.G. Ernst
Subduction Zone Metamorphism
1977
M. Talwani
and W.C.
Pitman III
Island Arcs, Deep Sea Trenches, and Back-Arc Basins
1981
F.J. Vine and
A.G. Smith
Extensional Tectonics Associated with Convergent Boundaries
1982
J.K. Leggett
Trench-Forearc Geology: Sedimentation and Tectonics on Modern and Ancient Active Plate Margins
1988
J. Pindell, S.C.
Cande, W.C.
Pitman III, D.
B. Rowley, J.F.
Dewey, J.F.
LaBreque, and
W. Haxby
A plate kinematic framework for models of Caribbean evolution
1988
M.I. Ross and
C.R. Scotese
A hierarchical tectonic model of the Gulf of Mexico and Caribbean Region
1989
C.S. Hutchison
Geological Evolution of South-east Asia
1990
G. Dengo and
J.E. Case
The Caribbean Region
1990
R.S.
D'Lemos ,R.S.
Strachan, and
C.G. Topley
The Cadomian Orogeny
1990
J. Pindell and
S.F. Barrett
Geological evolution of the Caribbean region: a plate tectonic perspective
1991
K.T. Biddle
Active Margin Basins
1992
B.C. Burchfiel,
P.W. Lipman,
and G.A.
Davis
Tectonic overview of the Cordilleran orogen in the western United States
1992
B.C. Burchfiel,
D.S. Cowan,
and G.A.
Davis
The Cordilleran Orogen: Conterminous U.S.
1995
P. Mann
Geologic and Tectonic Development of the Caribbean Plate Boundary in Southern Central America
1996
G.E. Bebout,
D.W. Scholl,
S.H. Kirby,
and J.P. Platt
Subduction: Top to Bottom
1998
R.J. Pankhurst
and C.W.
Rapela
The Proto-Andean Margin of Gondwana
2000
U.G. Cordani,
E.J. Milani, A.
Thomaz-Filho,
and D.A.
Campos
Tectonic Evolution of South America
2004
S. Lamb
Devil in the Mountain: A Search for the Origin of the Andes
2006
S.M.Kay and
V.A. Ramos
Evolution of an Andean margin: A tectonic and magmatic view from the Andes to the Neuquen Basin (35˚ - 39˚ S lat)
2006
O. Oncken, G.
Chong, G.
Fraunz, P.
Giese, H.J.
Götze, V.A.
Ramos, M.R.
The Andes: Active Subduction Orogeny
253
Strecker, and
P. Wigger
2009
J. Pindell and
L. Kennan
Tectonic evolution of the Gulf of Mexico, Caribbean, and norhern South America in the mantle reference frame: an update
2014
A. Gomez-
Tuena, S.M.
Straub, and
G.F. Zellmer
Orogenic Andesites and Crustal Growth
2015
P.G. DeCelles,
M.N. Ducea,
B. Carrapa,
and P.A. Kapp
Geodynamics of the Cordilleran Orogenic System: The Central Andes of Argentina and Northern Chile
2016
N.R.
McKenzie,
B.K. Horton,
S.E. Loomis,
D.F. Stockli,
N.J.
Planavsky,
and C.T.A. Lee
Continental arc volcanism as the principal driver of icehouse-greenhouse variability
2016
T. Moreno, S.
Wallis, T.
Kojima, and
W. Gibbons
The Geology of Japan
2017
W. Cao, C.T.A.
Lee, and J.S.
Lackey
Episodic nature of continental arc activity since 750 Ma: A global compilation
2019
B. Vaes, D.J.J.
van
Hinsbergen,
L.M.
Boschman
Reconstruction of subduction and back-arc spreading in the NW Pacific and Aleutian Basin: Clues to the causes of Cretaceous and Eocene plate reorganizations
2022
L.M.
Boschman
Andean mountain building since the Late Cretaceous: A paleoelevation reconstruction
C. Obduction of ophiolites onto continents due to the closure of back-arc basins and large ocean basins
1977
R.G. Coleman
Ophiolites: Ancient Oceanic Lithosphere ?,
1980
A. Panyaioyou
Ophiolites: Proceedings of the International Ophiolite Symposium
1983
G.J.H. McCall
Ophiolitic and Related Melanges
1984
I.G.Gass, S.J.
Lippard, and
A.W. Shelton
Ophiolites and Oceanic Lithosphere
2000
Y. Dilek,
Moores, E.M.,
D. Elthon, and
A. Nicolas
Ophiolites and Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program
2002
E.M. Moores
Pre–1 Ga (pre-Rodinian) ophiolites: Their tectonic and environmental implications
2003
Y. Dilek and
P.T. Robinson
Ophiolites in Earth History
2004
W.G. Ernst
Serpentine and Serpentinites: Mineralogy, Petrology, Geochemistry, Ecology, Geophysics, and Tectonics
2006
M. Metzler
The Formation and Obduction of Ophiolites Since the Late Precambrian
2008
J.E. Wright
and J.W.
Shervais
Ophiolites, Arcs, Batholiths: A Tribute to Cliff Hopson
254
2009
Y. Dilek and
P.T. Robinson
Special Issue on Mantle Dynamics and Crust-Mantle Interactions in Collisional Orogens
2014
Y. Dilek and H.
Furnes
Ophiolites
2021
Wakabayashi
and Dilek
D. Continent-
continent collisions
1964
A. Gansser
Geology of the Himalayas
1966
J.T. Wilson
Did the Atlantic Close and then Re-open?
1968
E.E. Oxburgh
The Geology of the Eastern Alps
1970
W. Hamilton
The Uralides and the Motion of the Russian and Siberian Platforms
1970
J. Rodgers
The Tectonics of the Appalachians
1972
T. Strand and
O. Kulling
Scandinavian Caledonides
1973
J.F. Dewey
and K.C.A.
Burke
Tibetan, Variscan, and Precambrian basement reactivation: products of continental collision
1973
J.F. Dewey,
W.C. Pitman
III, W.B.F.
Ryan, and J.
Bonin
Plate Tectonics and the Evolution of the Alpine System
1974
A.M. Spencer
Mesozoic – Cenozoic Orogenic Belts: Data for Orogenic Studies
1975
P. Molnar and
P. Tapponnier
Cenozoic tectonics of Asia: Effects of a continental collision
1975
J.H. Ostrom
and P.M.
Orville
Tectonics and Mountain Ranges, A special volume of the American Journal of Science Dedicated to John Rodgers
1979
A.L. Harris,
C.H. Holland,
and B.E.
Leake
The Caledonides of the British Isles – reviewed
1981
K.R. McClay
and N.J. Price
Thrust and Nappe Tectonics
1982
K.J. Hsü
Mountain Building Processes
1983
R.D. Hatcher,
H. William,
and I. Zietz
Contributions to the Tectonics and Geophysics of Mountain Chains
1985
D.G. Gee and
B.A. Strurt
The Caledonide Orogen – Scandinavia and Related Areas, Parts 1 and 2
1987
J.P. Schaer
and J. Rodgers
The Anatomy of Mountain Ranges
1988
A. L. Harris
and D.J.Fettes
The Caledonian-Appalachian Orogen
1988
P.F. Hoffman
Birth of a Craton: Early Proterozoic Assembly and Growth of Laurentia
1988
R.D. Nance,
T.R. Worsley,
and J.B.
Moody
The Supercontinent Cycle
1989
M.P. Coward,
Dietrich, D.,
and R.G. Park
Alpine Tectonics
255
1989
R.A. Gayer
The Caledonide Geology of Scandinavia
1991
R. Meissner,
T. Wever, and
P. Sadowiak
Continental collisions and seismic signature
1995
H. Williams
Geology of the Appalachian-Caledonian Orogen in Canada and Greenland
1996
A. Yin and
T.M.Harrison
The Tectonic Evolution of Asia
1996
O. Yilmaz, I.O.
Norton, D.
Leary, and R.J.
Chuchla
Tectonic evolution and paleogeography of Europe
1997
J.P. Burg and
M. Ford
Orogeny Through Time
1999
A.
Macfarlane,
R.B. Sorkhabi,
and J. Quade
Himalaya and Tibet: Mountain Roots to Mountain Tops
2000
W. Franke, V,
Haal, O.
Oncken, and
D. Tanner
Orogenic Processes: Quantification and Modelling in the Variscan Belt
2001
M.S. Hendrix
and G.A.
Davis
Paleozoic and Mesozoic tectonic evolution of central Asia: From continental assembly to intracontinental deformation
2002
J.R.M.
Catalan, R.D.
Hatcher, R.
Arenas, and
F.D. Garcia
Variscan-Appalachian dynamics: The building of the late Paleozoic basement
2002
J.A.
Winchester,
T.C. Pharoah,
and J.
Verniers
Palaeozoic Amalgamation of Central Europe
2004
G.M. Stampfli
and G.D.
Borel
The TRANSMED transects in space and time: Constraints on the paleotectonic evolution of the Mediterranean domain
2006
G.M. Stanpfli
and H.W.
Kozur
Europe from the Variscan to the Alpine cycles
2007
R.D. Hatcher,
Carlson, M.P.,
and J.H.
McBride
4-D Framework of Continental Crust
2007
U.
Linnemann,
R.D. Nance, P.
Kraft, and G.
Zulauf
The Evolution of the Rheic Ocean: From Avalonian-Cadomian Active Margin to Alleghenian-Variscan Collision
2008
S.
Siegesmund,
B.
Fügenschuh,
and N.
Froitzheim
Tectonic Aspects of the Alpine-Dinaride-Carpathian System
2009
D.J.J. van
Hinsbergen,
M.A.
Collision and Collapse at the Africa-Arabia-Eurasia Subduction Zone
256
Edwards, and
R. Govers
2011
D.C. Bradley
Secular trends in the geologic record and the supercontinent cycle
2011
R. Gloaguen
and L.
Ratschbacjer
Growth and Collapse of the Tibetan Plateau
2013
R.D. Nance
and J.B.
Murphy
Origins of the supercontinent cycle
2014
F.Corfu, D.
Gasser, and
D.M. Chew
New Perspectives on the Caledonides of Scandinavia and Related Areas
2014
R.D. Nance,
J.B. Murphy,
and M.
Santosh
The supercontinent cycle: A retrospective essay
2014
O.A. Pfiffner
Geology of the Alps
2014
A. Schettino
and E. Turco
Tectonic history of the western Tethys since the late Triassic
2015
J.P. Hibbard,
J.C. Pollock,
J.B. Murphy,
C.R. van Staal,
and J.D.
Greenough
Reeltime Geological Syntheses: Remembering Harold “Hank” Williams
2017
C.J.
Hawkesworth,
P.A. Cawood,
B. Dhuime,
and K.I.S.
Kemp
Earth’s continental lithosphere through time
2019
A.S. Merdith,
S.E. Williams,
S. Brune, A.S.
Collins, and
R.D. Müller
Rift and plate boundary evolution across two supercontinent cycles
2020
D.J.J. van
Hinsbergen,
T.H. Torsvik,
S.M. Schmid,
L.C. Matenco,
M. Maffione,
R.L.M.
Vissers, D.
Gürer, and W.
Spakman
Orogenic architecture of the Mediterranean region and kinematic reconstruction of its tectonic evolution since the Triassic
2021
A.S. Collins,
M.L. Blades,
A.S. Merdith,
and J.D.
Foden
Closure of the Proterozoic Mozambique Ocean was instigated by a late Tonian plate reorganization event
2023
Z.X. Li, Y. Liu,
and R. Ernst
A dynamic 2000–540 Ma earth history: from cratonic amalgamation to the age of supercontinent cycle
E.Exotic terranes and
accretion along active
continental margins
257
1974
C. Burrett
Plate tectonics and the fusion of Asia
1983
M. Audley-
Charles
Reconstruction of eastern Gondwanaland
1983
R.M.
Friedman
Accretionary History of Western North America during the Mesozoic and Cenozoic Eras
1984
M. Audley-
Charles
Cold Gondwana, warm Tethys, and Tibetan Lhasa Block
1984
I. Metcalfe
Stratigraphy, palaeontology and palaeogeography of the Carboniferous of Southeast Asia
1985
D.C.
Engebretson,
A. Cox, and
R.G. Gordon
Relative Motions between Oceanic and Continental Plates in the Pacific Basin
1988
M. Audley-
Charles
Evolution of the southern margin of Tethys (North Australian region) from early Permian to late Cretaceous
1988
A.M.C.
Şengör, D.
Altiner, A. Cin,
T. Ustaomer,
and K. Hsü
Origin and assembly of the Tethyside orogenic collage at the expense of Gondwana Land
1989
J.W. Hillhouse
Deep Structure and the Past Kinematics of Accreted Terranes
1989
D.G. Howell
Tectonics of Suspect Terranes: Mountain Building and Continental Growth
1990
Y. Isozaki, S.
Maruyama,
and F.
Furuoka
Accreted oceanic materials in Japan
1991
J.F. Dewey,
I.G. Gass, G.B.
Curry, N.B.W.
Harris, and
A.M.C. Sengor
Allochthonous Terranes
1992
B.C. Burchfiel,
P.W. Lipman,
and G.A.
Davis
Tectonic overview of the Cordilleran orogen in the western United States
1992
H. Gabrielse
and C.J.
Yorath
Geology of the Cordilleran Orogen in Canada
1993
I. Metcalfe
Southeast Asian terranes: Gondwanaland origins and evolution
1994
M.G.
Mihalynuk, J.
Nelson, and
L.J. Diakow
Cache Creek terrane entrapment: Oroclinal paradox with the Canadian Cordillera
1996
J.W.H.
Monger and
W.J.
Nokleberg
Evolution of the northern North American Cordillera: generation, fragmentation, displacement and accretion of successive North American plate margin arcs
1996
R.D. Nance
and M.D.
Thompson
Avalonian and Related Peri-Gondwanan Terranes of the Circum-North Atlantic
1996
A.M.C. Şengör
and B.A.
Nata'lin
Paleotectonics of Asia: fragments of a synthesis
1999
I. Metcalfe
Gondwana dispersion and Asian accretion: An overview
2000
R.A. Price and
J.W.H Monger
A transect of the southern Canadian Cordillera from Calgary to Vancouver
258
2005
A.P.M.
Vaughan,P.T.
Leat, and R.J.
Pankhurst
Terrane Processes at the Margins of Gondwana
2006
J.W. Haggart,
R.J. Enkin, and
J.W.H.
Monger
Paleogeography of the North American Cordillera: Evidence for and Against Large-Scale Displacements
2007
J.W. Sears,
T.A. Harms,
and C.A.
Evenchick
Whence the Mountains? Inquiries into the Evolution of Orogenic Systems: A Volume in Honor of Raymond A. Price
2007
B.F. Windley,
D. Alexeiev,
W.J. Xiao, A.
Kroner, and
G. Bardach
Tectonic models for the accretion of the Central Asian Orogenic Belt
2008
J.W. H.
Monger
The Evolution of Canada’s Western Mountains
2008
I.Y. Safonova,
V. Simonov,
M. Buslov, T.
Ota, and S.
Maruyama
Neoproterozoic basalts of the Paleo-Asian Ocean (Kurai accretionary zone, Gorny Altai, Russia): geochemistry, petrogenesis, and geodynamics
2010
Y. Isozaki, K.
Aoki, T.
Nakam, and S.
Yanai
New insight into a subduction-related orogen: a reappraisal of the geotectonic framework and evolution of the Japanese Islands
2010
L.M.
Parfenov, N.A.
Berzin, G.
Bardach, V.G.
Belichenko,
A.N. Bulgatov,
A.N. Dril, and
others
Metallogenesis and Tectonics of Northeast Asia
2011
I. Metcalfe
Southeast Asian terranes: Gondwanaland origins and evolution
2011
L.M.
Parfenov, W.J.
Nokleberg,
N.A. Berzin, G.
Bardach, and
others
Tectonic and metallogenic model for northeast Asia
2011
J.P. Burg
The Asia–Kohistan–India Collision: Review and Discussion
2013
T. Kusky, B.
Windley, I.
Safonova, K,
Wakita, J.
Wakabayashi,
A. Polat, and
M. Santosh
Recognition of ocean plate stratigraphy in accretionary orogens through Earth history: A record of 3.8 billion years of sea floor spreading, subduction, and accretion.
2014
C. Burrett, K.
Zaw, S.
Meffre, C.K.
Lai, K.
Somboon,
and C. Pol
The configuration of Greater Gondwana – Evidence from LA ICPMS, U-Pb geochronology of detrital zircons from the Palaeozoic and Mesozoic of Southeast Asia and China
259
2014
Y. Isozaki
Memories of pre-Jurassic lost oceans: How to retrieve them from extant lands.
2014
A.M.C.
Şengör, B.A.
Nata'lin, R.
van der Voo,
and G. Sunal
A new look at the Altaids: A superorogenic complex in Northern and Central Asia as a factory of continental crust, Part II
2014
A.M.C.
Şengör, B.A.
Nata'lin, G.
Sunal, and R.
van der Voo
A new look at the Altaids: A superorogenic complex in Northern and Central Asia as a factory of continental crust, Part I
2014
S. Zahirovic,
M. Seton, and
R.D. Müller
The Cretaceous and Cenozoic tectonic evolution of Southeast Asia
2023
E.L. Advokaat
and D.J.J. van
Hinsbergen
Finding Argoland: Reconstructing a microcontinental archipelago from the SE Asian accretionary orogen
2023
A.M.C.
Sengor, D.
Altiner, C.
Zabci, G.
Sunal, N. Lom,
E. Aylan, and
T. Öner
On the nature of the Cimmerian Continent
260
Table 7. Plate Tectonics Rules
Rule 1. Plates move only if they're pushed or pulled, not dragged (mostly).
The pattern of flow in the mantle is largely driven by lithospheric motions. Though intuitively attractive, the
idea of organized convection cells upon which the plates ride is incorrect and misleading. Oceanic plates
move faster than continental plates. Plates carrying large continents move slowly (e.g., Eurasia; Pangea).
Plates that are not driven by ridge push or slab pull do not move (Caribbean plate, Scotia plate). Plate
motion can be modelled by "balancing the forces" that drive and resist plate motion.
Rule 2. Subduction Rules.
Slab Pull is more important than Ridge Push (80% vs 20% of driving force). The Phanerozoic speed limit is
~20 cm/yr (India, 65 Ma).
Rule 3. Mid-Ocean Ridges are Passive Features.
Mid-ocean ridges are there because the crust, which is weak in tension, breaks when it is pulled.
Continental crust breaks first, because at the same depth it's closer to its melting temperature. Mid-Ocean
Ridges tend to align parallel to trenches.
Rule 4. Subduction is Forever.
Collision is the only way to stop subduction. Subduction is hard to start. Subduction graveyards
exist in the mantle.
Rule 5. Pacific versus Tethyan Subduction Systems
261
Pacific-style subduction systems are characterized by a ring of subduction with a spreading ridge
in the middle can exist for 10’s of millions of years. Tethyan-style subduction systems are asymmetric or
"one-sided" are unstable..
Rule 6. Plates Subduct Normally.
Orthogonal convergence (perpendicular to active margin) is least work.
Rule 7. The Style of Convergent margin depends of the absolute motions of the plates.
Andean margins - net convergence (~10 cm/yr) ; Western Pacific margins - net divergence (roll
back 1-2 cm/yr)
Rule 8. Island Arcs don't ride their trenches across oceans.
Back arc basins never evolve into wide (>30,000 km) ocean basins. 90% of all ophiolites form due to the
collapse of back-arc basins.
Rule 9. Slab Rollback Can Create Odd Intracontinental Ocean Basins
Oceanic lithosphere can become trapped (encircled) by continents (e.g. Mediterranean, Arctic,
Tethys north of Alps). Small, short-lived subduction zones can consume this ocean floor creating
intracontinental extension .
Rule 10. Mantle Plumes (i.e. Hot Spots) are Important (Sort of)
Many Hot Spots are derived from the core/mantle boundary. They provide a "good enough"
reference frame for absolute plate motions. Hot Spots "help" break apart continents.
Rule 11. Continental Collisions are Important (Really Important)
262
Collisions destroys a subduction zones and cause the global balance of plate driving forces to change. A
continent with many sutures will be weak and easily deformed (e.g. Asia following collision with India).
Rule 12. Plate Tectonics is a Catastrophic System (but not chaotic).
Plate motions are generally gradual; but every once and a while "WHAM"! 2 Important Instabilities: 1)
continent-continent collision & 2) ridge subduction. Ridge subduction both breaks supercontinents apart
and brings them back together (Wegener / Wilson Cycle). Supercontinents form and break apart because
of the metastable nature of plate evolution.
263
Table 8. Cretaceous Hot Spots, LIPS, and Hot Spot Tracks
Hot Spot
Name
Location when
Active
Principal
Volcanic
Feature
Location
Old
Age
Young
Age
Area
Associated
Sources
ID
#
bold = still
active
latitude
longitude
bold = LIP
(> .5 million
km2)
my
my
million
km2
Hot Spot
Track
A. Atlantic Ocean
1
Iceland
65 N
20 W
Alpha -
Mendeleev
Ridge
Arctic
150
120
1.3
Alpha-Mendeleev Ridge
"
"
"
Baffin Bay -
Disko Island
N. Atlantic
108
70
0
Baffin Bay-
Disko Ialnd-
Iceland
Johnston
and
Thorkelson,
2000
2
Bermuda
30.5 N
42.4 W
Bermuda
C. Atlantic
72
68
0
none
3
Great
Meteor
30 N
29 W
New England
Seamounts
C. Atlantic
100
75
0.1
New England
seamounts
Condie,
2001
"
"
"
Corner and
Sohm
Seamounts
C. Atlantic
75
60
0.1
Corner and Sohm
Seamounts
4
Cabo Verde
15N
24 W
E. Bahamas
Platform
C . Atlantic
144
130
0.2
Bahamas
Platform
"
"
"
C. Bahamas
Platform
C. Atlantic
153
144
0.4
Bahamas
Platform
5
Caribbean
11 N
73 W
Caribbean -
Colombian
Caribbean
95
85
1.1
none
Kerr et al.,
1997
6
Sierra Leone
- Ceara
1.2 N
28 W
Sierra Leone
- Ceara
Seamounts
S. Atlantic
70
68
0.7
none
Kumar,
1979
7
Tristan da
Cunha
37 S
12 W
Parana Flood
Basalts
Brazil-
Argentina
140
129
1
Rio Grande
Rise
Peate,
1997
"
"
"
Rio Grande
Rise
S. Atlantic
115
78
0.7
Rio Grande
Rise
"
"
"
Entendeka
Basalts
Namibia-
Angola
140
129
1
Walvis Ridge
Peate,
1997
8
Discovery
50 S
18 W
Discovery
Seamount
SW S.
Atlantic
66
58
0.1
none
9
Orcadas -
Meteor
55 S
14 W
Islas Orcadas
Rise
SW S.
Atalntic
66
58
0.1
none
264
"
"
"
Meteor Rise
SE S.
Atlantic
66
58
0.1
none
10
Agulhas -
Georgia
58 S
1 W
NE Georgia
Rise
SW S.
Atalntic
103
92
0.1
none
"
"
"
Agulhas
Plateau
SE S.
Atlantic
103
92
0.5
none
B. Indian Ocean
1
Reunion
21 S
56 E
Deccan
India
66
65
1.8
Mascarene-Chagos-
Laccadive
2
Roo
40 S
81 E
Roo Rise
E Indian
Ocean
130
120
0.1
none
3
Wallaby
44.8 S
69.8 E
Wallaby
Plateau
E Indian
Ocean
124
120
0.1
none
Direeen et
al., 2017
4
Mozambique
- Astrid
46 S
3 E
Astrid Ridge
Antarctica
154
118
0.4
Mozambique
Ridge
"
"
"
N.
Mozambique
Ridge
SW Indian
Ocean
136
130
0.2
Mozambique
Ridge
"
"
"
C.
Mozambique
Ridge
SW Indian
Ocean
130
124
0.2
Mozambique
Ridge
"
"
"
S.
Mozambique
Ridge
SW Indian
Ocean
124
114
0.2
Mozambique
Ridge
5
Marion
47 S
37.8 E
Madagascar
Plateau - del
Cano Rise
SW Indian
Ocean
95
58
1.6
Madagascar
Plateau - del
Cano Rise
Storey et
al., 1997
6
Cuvier
47 S
66 E
Cuvier
Plateau
E Indian
Ocean
138
131
0.1
none
7
Kerguelen
49 S
63 E
S. Kerguelen
SE Indian
Ocean
110
96
0.5
S. Kerguelen
Kent et al.,
1997
"
"
"
C. Kerguelen
- Broken
Ridge
SE Indian
Ocean
96
84
1.5
C. Kerguelen - Broken
Ridge
"
"
"
N. Kerguelen
SE Indian
Ocean
84
0
0.5
Ninetyeast
Ridge
8
Conrad
53 S
35 E
Conrad Rise
SW Indian
Ocean
88
80
0.4
Madagascar Plateau -
Conrad Rise
9
Rajmahal
53.4 S
48.4 E
Rajmahal
Traps
E India
134
120
2.3
none
10
Bovet
54 S
3.5 E
Maud Rise
S. S
Atlantic
120
114
0.1
none
265
11
Bunbury
57.7 S
57.2 E
Bunburry
volcanics
SW
Australia
136
130
0
none
Direeen et
al., 2017
C. Pacific Ocean
1
Hawaii
19.3 N
155.4 W
Hawaiian -
Emperor
N. Pacific
70
0
0.2
Hawaiian-
Emperor
2
Lynn
2 S
137 W
Line Island -
Hess
N.Pacific
115
70
0.1
Line Island -
Hess
Henderson,
1985
3
Hess
5 S
135 W
Hess
N. Pacific
124
112
0.8
none
Vallier et
al., 1981
4
Wilkes
8 S
125 W
Line Island -
Hess
N. Pacific
115
70
0.1
Line Island -
Hess
Henderson,
1985
5
Shatsky
12 S
152 W
Shatsky -
TAMU
N. Pacific
155
140
1
none
6
Somoa
14 S
170 W
Magellan
Seamounts
N. Pacific
70
0
0.1
Magellan
Seamounts
Henderson,
1985
7
Tahiti
18 S
150 W
Gilbert - Mid-
Pacific
Seamounts
N. Pacific
90
0
0.1
Gilbert -
Mid-Pacific
Seamounts
Henderson,
1985
8
Mid-Pacific
21 S
130 W
Mid-Pacific
Mountains
C. Pacific
140
114
1.1
Mid-Pacific
Mountains
Coffin and
Eldholm,
1991
9
Pitcairn -
Tuamotus
26 S
130 W
Austral -
Tuvalu -
Gilbert
C. Pacific
115
0
0.1
Austral -
Tuvalu -
Gilbert
Steinberger
and
O'Connell.
1998
10
Easter
27 S
109 W
Tuamotus-
Line-Mid-
Pacific
Seamounts
C. Pacific
115
0
0.1
Tuamotus-
Line-Mid-
Pacific
Seamounts
Henderson,
1985
11
Macdonald
30 S
140 W
Society-
Austral-Cook
C. Pacific
115
0
0.1
Society-
Austral-Cook
Henderson,
1985
12
Ontong Java
40 S
161 W
Ontong Java
C.Pacific
122
119
4.3
none
Neal et al.,
1997
13
Manihiki
44 S
110 W
Manihiki
C. Pacific
122
119
0.9
none
14
Louisville
51 S
141 W
Louisville
Ridge
S. Pacific
90
0
0.1
Louisville
Ridge -
Gilber -
Marshall
Steinberger
and
O'Connell.
1998;
Sager, 2007
15
Whitsunday
70 S
131 E
Whitsunday
Queensland
120
90
0.2
none
Bryan et
al., 2000
16
Hikurangi
78 S
97 W
Hikurangi
SW Pacific
121
119
0.4
none
266
267
Table 9. Elevation, Environments and Geological Evidence
Code Elevation Environments Geological Evidence
9
10,000 to 4000 m
Collisional mountains
High-T, high-P metamorphics
8
4000 to 2000 m
Andean-type mountains
Andesites/granodiorites in a continental
setting
7
2000 to 1000 m
a. Island arc volcanos
Andesites/granodiorites in a marine setting
b. Intra-continental rift
shoulders
Adjacent fanglomerates
6
1000 to 200m
a. Rift valley
Basalts, lake deposits in grabens
b. Some forearc ridges
Tectonic mélanges
5
200m to Sea Level
a. Coastal plains
Alluvial complexes
b. Lower river systems
Major floodplain complexes
c. Delta tops
Swamps and channel sands
4
Sea Level to -50 m
a. Inner shelves
Heterogeneous marine sediments
b. Reef-dammed shelves
Bahamian-type carbonates
c. Delta fronts
Topset silts and sands
3
-50 to -200 m
a. Outer shelves
Fine sediments, most bioproductites
b. Some epeiric basins
Fine clastics or carbonates
c. Pro-deltas
Foreset silts and proximal turbidites
2
-200 to -4000 m
a. Continental slope/rise
Slump/contourite facies
b. Mid-ocean ridges
Oceanic crust less than 60 m.y. old
c. Pro-delta fans
Bottomset clays and distal turbidites
1
-4000 to -6000 m
Ocean floors
Pelagic sequences on oceanic crust
0
-6000 to -12000 m
Ocean trenches
Turbidites on pelagic sequences
268
Table 10. Bibliography of Sources for Cretaceous Paleogeographic Maps
Explanation: Each bibliographic citation is followed by the region of the world, the total number of Maps, and the
Cretaceous time intervals.
For example: Dercourt, J., Ricou, L.E., and Vrielynick, B., 1993. Atlas Tethys, Palaeoenvironmental Maps, Gauthier-
Villars, Paris, 32 pp., / Tethys region from northern Australia to North America/ 14 Maps / Cretaceous (Early
Aptian, Late Cenomanian, Late Maastrichtian)
Blakey, R. C. (2002). Global Paleogeography, Rectilinear Projection, Colorado Plateau Geosystems, Inc., Flagstaff, AZ,
(DVD)/Global/31 Maps/ Aptian (120 Ma), Albian (105 Ma), Cenomanian-Turonian (90 Ma), Maastrichtian (65 Ma)
Blakey, R. C. (2008). Gondwana paleogeography from assembly to breakup - A 500 m.y. odyssey, Geological Society
of America Special Papers, v. 441, p. 1-28./Gondwana/ 18 Maps/ Early Cretaceous (120 Ma), mid-Cretaceous
(105 Ma), Late Cretaceous (90 Ma), Late Cretaceous (75 Ma)
Blakey, R. C. (2011). Paleogeography of Europe, Colorado Plateau Geosystems, Inc., Flagstaff, AZ, (DVD)/Western
Europe/26 Maps/ Early Cretaceous (125 Ma), Late Cretaceous (100 Ma, 75 Ma)
Blakey, R. C. (2013). Key Time Slices of North American Geologic History, Colorado Plateau Geosystems, Inc.,
Flagstaff, AZ, (DVD)/Western Europe/37 Maps/ Early Cretaceous (130 Ma, 115 Ma, 105 Ma), Late Cretaceous (92
Ma, 85 Ma, 72 Ma)
Boucot, A. J., Chen Xu, Scotese, C. R., & Fan Jun-Xuan (2009). Atlas of Phanerozoic Lithologic Indicators of Climate,
Science Press, Nanjing,204 pp. (in Chinese)./Global/27 Maps/Early Cretaceous (120 Ma), early Late Cretaceous
(100 Ma), late Late Cretaceous (80 Ma)
Bozhko, N. A., & Khain, V. E. (1987). Gondwana Paleotectonic Maps, Ministry of Higher and Secondary Special
education of the U.S.S.R., and Ministry of Geology, U.S.S.R., produced by the Geological Complex of Central
Regions of the U.S.S.R., 30 pp. / Africa, South America, Arabia, Madagascar, India, Antarctica, Australia / 22
Maps/Late Jurassic – Early Cretaceous (160 my), Late Cretaceous (65 my)
Cook, P. J. (1990). Australia: Evolution of a Continent. Bureau of Mineral Resources (BMR), Paleogeographic Group,
Australian Government Publishing Service, Canberra. / Australia / 69 Maps/ Cretaceous (11 Maps)
Cook, T. D., & Bally, A. W. (1975). Stratigraphic Atlas of North America and Central America, Princeton University
Press, Princeton, New Jersey, 272 pp. / North America / 42 Maps/ Cretaceous (top Jurassic-mid Aptian, mid
Aptian- mid Cenomanian, mid Cenomanian – top Turonian, Coniacian-Santonian, Campanian- Maestrichtian)
Cope, J. C. W., Ingham, J. K., & Rawson, P. F. (1992). Atlas of Paleogeography and Lithofacies, Geological Society of
London, Memoir 13, 153 pp. / Great Britain and the North Sea/ ~80 Maps/ Cretaceous (Berriasian, Mid
Hauterivian, Late Aptian, Latest Albian, Early Cenomanian, Late Campanian)
Dercourt, J., Ricou, L. E., & Vrielynick, B. (1993). Atlas Tethys, Palaeoenvironmental Maps, Gauthier-Villars, 307 pp.,
Paris. / Tethys from northern Australia to eastern North America/ 14 Maps/ Cretaceous (Early Aptian, Late
Cenomanian, Late Maastrichtian)
269
Dercourt, J., Gaettani, M., Vrielynck, B., Barrier, E., Biju-Duval, B., Brunet, M. F., Cadet, J. P., Crasquin, S., &
Sandulescu, M. (2000). Atlas Peri-Tethys, Paleogeographical Maps, Commission for the Geologic Map of the
World (CCGM/CGMW, 24 Maps and explanatory notes, 269 pp., Paris. / Western Tethys from Caspian Sea to
Grand Banks/24 Maps/ Cretaceous (Early Hauterivian, Early Aptian, Late Cenomanian, Early Campanian, Late
Maastrichtian)
Evans, D., Graham, C., Armour, A., & Bathurst, P. (2003). The Millenium Atlas: Petroleum Geology of the Central and
Northern North Sea, Geological Society of London, 389 pp. /Central North Sea/>46 Maps/ Early Cretaceous (late
Ryazanian-early Valanginian, late Valanginian-late Barremian, early Aptian-early Albian), Late Cretaeous
(Campanian- Maastrichtian)
Furon, R. (1941). La Paleogeographie, Essai sur l’evolution des continents et des oceans, Payot, Paris, 530 pp. /
Global/15 Maps/ Cretaceous (early-middle, late)
Golonka, J., Ross, M. I., & Scotese, C. R. (1994). Phanerozoic Paleogeographic and Paleoclimatic Modeling Maps, in
A. F. Embry, B. Beauchamp, and D.J. Glass (editors), Pangea, Global Environments and Resources, Canadian
Society of Petroleum Geologists, Memoir 17, p. 1-47. /Global/ 29 Maps/ Cretaceous (Valanginian, Aptian, Albian,
Cenomanian, Coniacian, Maastrichtian)
Golonka, J. (2000). Cambrian-Neogene Plate Tectonic Maps, Rozprawy Habilitacyine No. 350, Wydawnictwo
Uniwersytetu Jagiellonskiego, Krakow, 123 pp./Globa1/ 31 Maps/ latest Jurassic – earliest Cretaceous,
Cretaceous (Early, Early-earliest Late, Late), Late Cretaceous-earliest Paleogene
Hambrey, M. J., & Harland, W. B. (1981). Earth’s pre-Pleistocene glacial record, Cambridge University Press,
Cambridge, 1004 pp. / Global/ >50 Maps/pre-Pleistocene/
Hutchison, C. S. (1989). Geological Evolution of South-East Asia, Oxford University Press, Oxford, 368 pp./ Southeast
Asia/8 Maps/ Late Jurassic – mid Cretaceous, mid Cretaceous – Neogene/
Hulver, M. (1985). Cretaceous Marine Paleogeography of Africa, Master’s Thesis, University of Chicago, Chicago,
/Africa/5 Maps/Cretaceous (Valanginian, Aptian, Ceno Manian, Coniacian, Maastrichtian)/
Kazmin, V. G., & Natapov, L. M. (1998). The Paleogeographic Atlas of Northern Eurasia: Paleogeographic Maps on
the Palinspastic Reconstruction (Orthographic Projection), Institute of Lithospheric Plates, Russian Academy of
Natural Sciences, Moscow. /Northern Eurasia/26 Maps/ 70 Ma ( Maastrichtian), 80 Ma (Santonian & Campanian),
90 Ma (Cenomanian, Turonian, & Coniacian), 100 Ma, (Late Albian), 110 Ma (Early Albian), 120 Ma (Aptian), 130
Ma (Hauterivian & Barremian), 140 Ma (Berriasian & Valanginian)
Khain, V. Ye., Ronov, A. B., & Balukhovsky, A. N. (1976). Cretaceous lithologic associations of the world, International
Geology Review, v. 18, no.11, p. 1269-1295 (English translation from Russian) Sovetskaya Geologiya, 1976, v. 11,
p. 10-39.
Kriest, J. (1991), Plate-Tectonic Atlas, Exploration Bulletin, no. 258 (1995/5), Shell Exploration Company, Den Hague,
8 pp./Global/27 Maps/Hauterivian (135 Ma), Aptian (120 Ma), Albian (105 Ma), Cenomanian/Turonian (90 Ma),
Campanian/ Maastrichtian (75 Ma)
Kiessling, W. (2001). Paleoclimatic significance of Phanerozoic Reefs, Geology, v. 29, no. 8, p. 751-754. / Phanerozoic
Reef Paleolatitudes/
270
Kiessling, W., Flügel, E., & Golonka, J. (2002). Phanerozoic Reef Patterns, SEPM (Society for Sedimentary Geology)
Special Publication Number 72, 775 pp. /Global/ >40 Maps / Cretaceous (Berriasian,late Valangian-early
Aptian,late Apian-middle Ceno Manian, late Ceno Manian-Santonian), Campanian-Danian
Mallory, W. W. (1972). (Editor), Geological Atlas of the Rocky Mountain Region, Rocky Mountain Association of
Geologists, Denver, 331 pp. (available as pdf from AAPG)/Rocky Mountain States/ >62 Maps/ Cretaceous
(Neocomian-Aptian, early Albian, middle-late Albian, late Skull Creek, latest Albian, early Belle Fourche, middle
Greenhorn, middle Carlisle, early Niobrara, middle Niobrara, Telegraph Creek, latest Eagle, early Claggett, middle
Judith River, middle Bearpaw, early Fox Hills, latest Cretaceous)
McCrossan, R. G., Glaister, R. P., Austin, G. H., & Nelson, S. J. (1964). Geological History of Western Canada, Alberta
Society of Petrleum Geologists, Calgary, Alberta, 232 pp./western Canada/ >45 Maps/ Cretaceous (Aptian-lower
Albian, lowewr-middle Albian, upper Albian, Ceno Manian-lower Campanian, middle Campanian, upper
Campanian-early Maastrichtian
Moore, T. L., & Scotese, C. R. (2012). Ancient Earth: Breakup of Pangea, Vers. 1.0, iOS Mobile Application, retrieved
from http://itunes.apple.com/ Global/24 Maps/ KT Boundary (65.5 Ma), Maastrichtian (68 Ma), Early Campanian
(80.3 Ma), Turonian (91.1 Ma), late Albian (101.8 Ma), early Albian (110.0 Ma), early Aptian (121.8 Ma),
Hauterivian (132.0 Ma), Berriasian (143.0 Ma)
Mossop, G., & Shetson, I. (1994). Geological Atlas of Western Canada Sedimentary Basins, Canadian Society of
Petroleum Geologists, Calgary, 510 pp. /Western Canada/ >52 Maps/ Cretaceous (latest Jurassic – earliest
Cretaceous, Berriasian, Berriasian-Valanginian, latest Barremian – early late Aptian, late Aptian, earliest Albian,
early Albian, late early Albian, middle Albian, Cenomanian, Turonian, early Campanian, late early Campanian,
middle Campanian, late Campanian, middle Maastrichtian)
Ronov, A. B., Khain, V. Ye., & Balukhovsky, A. (1989). Atlas of Lithological-Paleogeographical Maps of the World,
Mesozoic and Cenozoic of the Continents, USSR Academy of Sciences, USSR State Committee for Public
Education, Ministry of Geology of the USSR, 79 pp., Leningrad./Global/13 Maps/ Early Cretaceous, Late
Cretaceous
Schandelmeier, H., & Reynolds, P. O. (1997). Paleogeographic-Palaeotectonic Atlas of North-Eastern Africa, Arabia,
and Adjacent Areas: Late Proterozoic to Holocene, A.A. Balke Ma, Rotterdam, 160 pp./Northeast Africa and
Arabia/17 Maps/ Cretaceous (Valanginian, Aptian, Campanian- Maastrichtian)
Schuchert, C. (1955). (published posthumously), Atlas of paleogeographic Maps of North America, C. O. Dunbar & C.
M. Levene (Eds.), John Wiley & Sons, New York, and Chap Mann & Hall, Ltd., London, 177 pp. / North America/
84 Maps/ Lower Cretaceous (Lower Comanchean, Middle Comanchean, Upper Comanchean), Lower Upper
Cretaceous (Turonian), Upper Cretaceous (Lower Senonian-Niobrarian, Campanian) High Upper Cretaceous
(Upper Maastrichtian and ? Danian)
Scotese, C. R. (1998). Digital Paleogeographic Map Archive on CD-ROM, PALEO MAP Project, Arlington,
Texas./global/40 Maps/ Berriasian-Barremian (140 Ma), Aptian (120 Ma), Albian (100 Ma), Ceno Manian-
Turonian (90 Ma), early Campanian (80 Ma), Maastrichtian (70 Ma)
Scotese, C. R. (2001). Atlas of Earth History, Volume 1, Paleogeography, PALEOMAP Project, Arlington, Texas, 52
pp./Global/16 Maps/ Late Cretaceous, KT Boundary
Scotese, C. R. (2004). Cenozoic and Mesozoic Paleogeography: Changing Terrestrial Biogeographic Pathways, in
Frontiers of Biogeography: New Directions in the Geography of Nature, M. V. Lomolino & L. R. Heaney (Eds.),
271
Sinauer Associates, Inc., Sunderland, Massachusetts, p. 1-26. /Global/18 Maps/ Cretaceous(Berriasian,
Barremian-Aptian, Albian-Cenomanian, Turonian, Campanian, middle Maastrichtian)
Scotese, C. R. (2014). Atlas of Late Cretaceous Paleogeographic Maps, PALEOMAP Atlas for ArcGIS, volume 2, The
Cretaceous, Maps 16 - 22, Mollweide Projection, PALEOMAP Project, Evanston, IL. ResearchGate Academia,
/Global/ 7 Maps/ Map 16 K/T Boundary (latest Maastrichtian, 65.5 Ma), Map 17 Late Cretaceous (Maastrichtian,
68 Ma), Map 18 Late Cretaceous (Late Campanian, 73.8 Ma), Map 19 Late Cretaceous (Early Campanian, 80.3
Ma), Map 20 Late Cretaceous (Santonian & Coniacian, 86 Ma), Map 21 Mid Cretaceous (Turonian, 91.1 Ma), Map
22 Mid Cretaceous (Ceno Manian, 96.6 Ma)/
Scotese, C. R. (2014). Atlas of Early Cretaceous Paleogeographic Maps, PALEOMAP Atlas for ArcGIS, volume 2, The
Cretaceous, Maps 23-31, Mollweide Projection, PALEOMAP Project, Evanston, IL. ResearchGate Academia,
/Global/ 14 Maps/ Map 23 Early Cretaceous (late Albian, 101.8 Ma), Map 24 Early Cretaceous (middle Albian,
106 Ma), Map 25 Early Cretaceous (early Albian, 110 Ma) Albian Supersequence Boundary and Transgressive
System Tract, Map 26 Early Cretaceous (late Aptian, 115.2 Ma), Map 27 Early Cretaceous (early Aptian, 121.8
Ma), Map 28 Early Cretaceous (Barremian, 127.5 Ma), Map 29 Early Cretaceous (Hauterivian, 132 Ma), Map 30
Early Cretaceous (Valanginian, 137 Ma) Barremian-Hauterivian Supersequence boundary and Transgressive
Systems Tract, Map 31 Early Cretaceous (Berriasian, 143 Ma) Berriasian Supersequence boundary and Maximum
Flooding Surface/
Scotese, C. R., & Golonka J. (1992). Paleogeographic Atlas, PALEOMAP Progress Report 20-0692, Department of
Geology, University of Texas at Arlington, Texas, 34 p. /Global/ 28 Maps/ Cretaceous (Valanginian, Aptian,
Cenomanian, Coniacian, Maastrichtian)
Scotese, C. R., & Winn, K. (1987). Phanerozoic Paleogeographic Maps, Paleoceanographic Mapping Project (POMP)
Progress Report 33-1287, 31 pp. (UTIG Technical Report 84)/ Global / 14 Maps/ Cretaceous (Cenomanian,
Maastrichtian)
Smith, A. G., Smith, D. G., & Funnell, B. M. (1994). Atlas of Mesozoic and Cenozoic Coastlines, Cambridge University
Press, 99 pp., /Global/31 Maps/ Cretaceous (Valanginian-Berriasian, Barremian-Hauerivian, Aptian, Albian,
Cenomanian, Turonian, Coniacian, Santonian, Campanian, Maastrichtian)
Ulmishek, G. F., & Klemme, H. D. (1990). Depositional Controls, Distribution, and Effectiveness of the World’s
Petroleum Source Rocks, U.S.G.S. Bulletin 1931, Denver, 59 pp./Global/ 6 Maps/ Middle Cretaceous
Veevers, J. J. (1984). Phanerozoic Earth History of Australia, Oxford Monographs on Geology and Geophysics, no. 2,
Oxford University Press, New York, 418 pp. / Australia/ 44 Maps/ Cretaceous (earliest, early, Aptian, Aptian-
Albian, Albian, Cenomanian, Turonian, Campanian, Maastrichtian)
Veevers, J. J. (2000). Billion-year history of Australia and neighbours in Gondwanaland, GEMOC Press, Sydney, 388
pp. /Australia, Antarctica, South Africa, S. South America/ 41 Maps/ Cretaceous(Neocomian-Aptian, Aptian,
Aptian-Albian Cenomanian, Turonian-Campanian ,Campanian)
Vinogradov, A. P., Vereshchagin, V. N., Nalivkin, V. D., Ronov, A. B., Khabakov, A. V., & Khain, V. E. (1968b). Atlas of
Lithological-Paleogeographical Maps of the U.S.S.R. (1:7.500,00), Tome III, Triassic, Jurassic, and Cretaceous,
Tome III editors, V.N. Vereshchagin and A.B. Ronov, Ministry of Geology and the U.S.S.R., Academy of Sciences
of the U.S.S.R., Moscow. /USSR /26 Maps/ Valanginian, Hauterivian, Barremian, Aptian, Albian, Cenomanian,
Turonian, Coniacian, Santonian, Campanian, Maastrichtian
Vrielynck, B., & Bouyesse, P. (2001). Le Visage Changeant de la Terre, L’eclatement se la Pangee et la mobilite des
continents au cours des desniers 250 million d’annees en 10 cartes, Commission de la Carte Geologique du
Modne, Paris, 32 pp. (Also published in English, 2003) /Global/ 19 Maps/Cretaceous (Cenomanian, Maastrichtian)
272
Walsh, D. B. (1996). Late Jurassic through Holocene Paleogeographic Evolution of the South Atlantic Borderlands,
Master’s Thesis, University of Texas at Arlington, 136 pp. / South Atlantic /9 Maps/ Cretaceous (Valanginian,
Aptian, Albian, Cenomanian, Coniacian-Turonian-Santonian, Maastrichtian)
Wang Hongzhen (1985). Atlas of the Paleogeography of China, Chinese Academy of Sciences, Wuhan College of
Geology, Cartographic Publishing House, Beijing, 85 pp. /China/41 Maps/Precambrian Cretaceous (early Early,
late Early, Late)
Wilford, G. E. (1979). Phanerozoic Paleogeography, in BMR Earth Science Atlas of Australia, E.K. Carter (editor),
Bureau of Mineral Resources, Geology & Geophysics, Australia, 31 pp./Australia/20 Maps/ middle Jurassic – early
Cretaceous, Cretaceous (early, late)
Willis, K. J., & McElwain, J. C. (2002). The Evolution of Plants, Oxford University Press, 378 pp. /Global Biome Maps/9
Maps/ Maastrichtian
Ziegler, A. M., Scotese, C. R., & Barrett, S. F. (1983). Mesozoic and Cenozoic paleogeographic Maps, in Tidal friction
and the Earth's Rotation II, P. Broche and J. Sunder Mann, eds., Springer-Verlag, Berlin, p.241-252. /Global/7
Maps/ Cretaceous (Cenomanian, Maastrichtian)
Ziegler, P. A. (1982). Geological Atlas of Western and Central Europe, Shell Internationale Petroleum Maatschappij
B.V., Den Haag, 130 pp. / Western and Central Europe/21 Maps/ Cretaceous (Berriasian-Barremian, Aptian-
Albian), Cretaceous-Tertiary Boundary (Cenomanian-Danian)
Ziegler, P. A. (1988). Evolution of Arctic-North Atlantic and western Tethys, American Association of Petroleum
Geologists, Memoir 43, 198 pp./North Atlantic, Arctic, and western Terthys/ 20 Maps/ Cretaceous (Berriasian-
Barremian, Aptian-Albian. Turonian-Campanian)
Ziegler, P. A. (1990). Geological Atlas of Western and Central Europe, Shell Internationale Petroleum Maatschappij
B.V., Den Haag, 239 pp. / Western and Central Europe/28 Maps/ Cretaceous (Berriasian-Valanginian,
Hauterivian-Barremian, Aptian-Albian, Cenomanian-Turonian,), Cretaceous-Tertiary Boundary
Zonenshain, L. P., Kuzmin, M. I., & Natapov, L. M. (1990). Geology of the USSR: A Plate Tectonic Synthesis, American
Geophysical Union, Geodynamics Series no. 21, 242 pp. /Europe and USSR/ 18 Maps/ Cretaceous (Early, Mid,
Late), Cretaceous-Tertiary Boundary
273
Table 11. Cretaceous Rivers
Age Ma
ID
Name
145
140
135
130
125
120
115
105
100
95
90
85
80
75
70
65
1
Nile
√
√
√
√
√
√
√
√
r
√
√
√
√
√
√
√
2
Lena
√?
√
√?
√?
√?
√
s?
s?
√
√
√
√
√
√
√
3
McKenzie
√
√
√
√
√
√
√
√
s
√
√
s
√
√
√
4
Murray-Darling
N
N
√
√
r
√
√
√
√
√
√
√
5
Amur
N
N
√
√
r?
√
√
√
√
√
√
s
√
6
Niger
√
√
f
f
f
√
f
√
√
√
7
Yangtze
√
√
√
√
√
√
√
√
√
8
Huang Ho
√
√
√
√
√
√
√
√
√
√
9
Uruguay
s
s
√
s
√
√
f
√
√
√
10
Zambezi
√
s
√
√
s
√
√
√
√
√
√
11
Congo
r
r
r
r
S
S
√
r
r
r
r
r
√
r
r
12
Amazon
N
r
r
r
r
r
r
r
r
r
r
√
f
√
r
r
13
Mekong
s
s
√
s
s
s
s
s
s
s
s
s
14
Mississippi
√
√
s
s
s
s
s
s
s
E
f
f
f
√
f
f
15
Yenisey
s
f
f
√
√
√
√
√
√
√
16
Ob-Irtysch
s
s
s
f
√?
√?
17
Volga
s
f
f
f
f
f
s
√?
√
18
Rio Grande
s
f
f
f
f
f
f
f
f
s
f
f
19
Syr-Darya
s
f
f
f
f
f
f
f
f
20
Yukon
√
A
Orinoco
√
√
√
f
f
f
f
√
√
√
√
√
f
√
f
√
B
Patagonian
√
√
√
√
√
f
√
C
Senegal
√
√
√
√
√
√
√
√
√
√
√
√
√
D
Arabian
√
√
√
√
√
√
√
√
√
√
√
√
√
√
E
Lamu
√
√
√
√
√
√
√
√
√
√
√
√
√
√
√
√
F
trans-Saharan
√
√
√
√
√
√
√
√
√
√
√
√
√
√
√
G
Australis
√
√
H
Namib
√
√
√
√
√
√
√
√
I
trans-Hudson
N
W
N
E
N
E
N
W
W
W
f
f
f
f
E
E
J
trans-Asian
S
S
S
S
K
trans-Indian
N
N
N
L
trans-Europe
W
S
S
S
S
S
M
trans-Siberian
W
W
W
W
W
W
W
W
W
W
W
W
W
W
W
W
N
trans-Antarctic
W
N
E
E
W
W
W
W
W
E
W
W
O
Austral-Antarctic
N
N
274
P
Greenland
S
S
N
N
N
N
E
E
E
N
N
N
Table 11. Cretaceous Rivers. The numbered rivers have modern counterparts. The lettered rivers only existed in the
Cretaceous. √ = Cretaceous river, r = flowing in reverse direction, f = flooded, s = shorter length, N, E, S, W
indicate river flow direction; The following major modern rivers were not present in the Cretaceous: Amu,
Brahmaputra, Colorado, Columbia, Danube, Ganges, Indus, Loire, Paraguay=Uruguay, Parana, Rhine, Rhone, Syr,
and Tigres-Euphrates.
275
276
Table 13. Permo-Triassic Extinction
and End Cretaceous Extinction
Resources
Events
Year
Author
Title
PTr
2023
Song & Scotese
The end-Paleozoic Great Warming
PTr
2015
Cui and Kump
Global warming and the end-Permian extinction event: Proxy and
modeling perspectives
PTr
2015
Wignall
The Worst of Times: How Life on Earth Survived Eighty Million Years
of Extinctions
KPg
2014
Wilson, Clemens,
Horner, and
Hartman (editors)
Through the End of the Cretaceous in the Type Locality of the Hell
Creek Formation in Montana and Adjacent Areas
KPg,
PTr
2013
MacLeod
The Great Extinctions: What Causes Them and How They Shape Life
KPg
2012
Hart, Yancey,
Leighton, Miller,
Liu, Smart, and
Twitchett
The Cretaceous-Paleogene boundary on the Brazos River, Texas:
New stratigraphic sections and revised interpretations
PTr
2009
Sengör &
Atayman
The Permian Extinction and Tethys: An Exercise in Global Geology
KPg
2008
Nichols & Johnson
Plants and the K-T Boundary
PTr
2006
Erwin
Extinction: How Life on Earth Nearly Ended 250 Million Years Ago
KPg,
PTr
2004
Hallam
Catastrophes and Lesser Calamities
PTr
2004
Ward
Gorgon: Paleontology, Obession, and the Greatest Catastrophe in
Earth's History
PTr
2003
Benton
When Life Nearly Died: The Greatest Mass Extinction of All Time
Future
2003
Ward & Brownlee
The Life and Death of Planet Earth
KPg
2002
Hartman,
Johnson, &
Nichols (editors)
The Hell Creek Formation and the Cretaceous-Tertiary Boundary in
the Northern Great Plains: An Integrated Continental Record of the
End of the Cretaceous
KPg,
PTr
1999
Courtillot
Evolutionary Catastrophes: The Science of Mass Extinction
KPg
1998
Powell
Night Comes to the Cretaceous: Comets, Craters, Controversy, and
the Last Days of the Dinosaurs
KPg,
PTr
1997
Hallam & Wignall
Mass Extinctions and their Aftermath
KPg,
PTr
1996
Hart
Biotic Recovery from Mass Extinctions
KPg
1996
Officer & Page
The Great Dinosaur Extinction
KPg
1996
Archibald
Dinosaur Extinction and the End of an Era
KPg
1995
Carlisle
Dinosaurs, Diamonds, and Things from Outer Space: The Great
277
Extinction
PTr
1993
Erwin
The Great Paleozoic Crisis: Life and Death in the Permian
KPg,
PTr
1991
Raup
Extinction: Bad Genes or Bad Luck?
KPg,
PTr
1990
Kauffman &
Walliser (editors)
Extinction Events in Earth History
KPg,
PTr
1989
Donovan (editor)
Mass Extinctions: Processes and Evidence
KPg,
PTr
1987
Stanley
Extinction
KPg
1986
Raup
The Nemesis Affair: A Story of the Death of Dinosaurs and the
Ways of Science
KPg
1986
Hsü
The Great Dying: Cosmic Catastrophe, Dinosaurs, and the Theory
of Evolution
KPg,
PTr
1984
Berggren and Van
Couvering
(editors)
Catastrophes and Earth History: The New Uniformitariansim
KPg,
PTr
1984
Nitecki (editor)
Extinctions
KPg
1983
Allaby & Lovelock
The Great Extinction: What killed the dinosaurs and devastated
the Earth?
278
Table 14. Contents of Supplementary Materials
Part I. Phanerozoic Content https://doi.org/10.5281/zenodo.10659112
1. Text Items
a. An annotated bibliography with more than 100 sources that feature Phanerozoic paleogeographic maps
b. Brief essay describing how global plate tectonic models developed from the mid 1970’s to the present-day
c. The Rules of Plate Tectonics
d. A series of tutorials that describe how to build digitial paleogeographic maps
e. A pdf of this paper as well as Scotese(2021) and Scotese et al. (2021)`
2. Animations
a. Plate Tectonic: “Plate Tectonics, 1.5 by – Today: Ancient Oceans and Continents”, https://youtu.be/IlnwyAbczog
b. Plate Tectonic: “Plate Tectonics, 1.5 by – Today: Ancient Oceans and Continents”, (with continents and oceans
labelled) https://youtu.be/AsCYZ-k-0uc
c. Paleogeographic: "Plate Tectonics, Paleogeography, and Ice Ages", https://youtu.be/UevnAq1MTVA
d. Paleogeographic: “Plate Tectonics, Paleogeography, and Ice Ages (dual hemispheres)”,
https://youtu.be/bzvOMee9D1o
e. Paleoclimate: “Phanerozoic Global Temperature”, https://youtu.be/FF3Mz8ZFyh8
f. Paleoclimate: “Phanerozoic Rainfall”, https://youtu.be/88cO9ba0DR8
g. Köppen belt evolution during the Phanerozoic (Scotese et al., 2021), https://youtu.be/DGf5pZMkjA0
3. Global Tectonic Model
a. Rotation model (.rot) and static polygons for plate tectonic reconstructions (v.19o_r1d)
b. Rotation model (.rot) and static polygons for paleogeographic and paleoclimatic reconstructions
(v.m17v2d3_81_v18e)
4. Maps (Equirectangular, Mollweide, Robinson, and North & South Orthographic Polar projections)
a. Plate Tectonic Maps, 0 – 1500 Ma (see Figure 2A)
b. Paleogeographic Maps (with 3D relief, modern coastlines and political boundary overlays, .jpg), 0 – 1000 Ma
(see Figure 2B)
c. Paleogeographic Maps (no overlays, .bmp ), 0 – 1000 Ma
d. Paleotemperature Maps, 0 – 540 Ma (Valdes et al., 2021)
e. Paleorainfall Maps, 0 – 540 Ma (Valdes et al., 2021)
f. Köppen Maps, 0 – 540 MA (Boucot et al., 2013; Scotese et al., 2021)
g. Tectonic Map of the World (see Figure 8)
5. PaleoDEMs (i.e. paleo-Digitial Elevation Model)
a. netcdf format (.nc), .1˚ x .1˚ and 1˚ x 1˚ resolution
b. text format (.csv), 1˚ x 1˚ resolution
6. Paleoclimate Digital Data Files
a. Spreadsheet with Mean Annual Temperature, (360 x 180 grid, .csv)
b. Spreadsheet with Mean Annual Rainfall, 360 x180 grid, .csv)
c. Spreadsheet with Boucot et al. (2013) lithological indicators of climate (0 -540 Ma)
7. Miscellaneous Diagrams
a. High resolution versions of all of the figures that appear in the paper.
b. A complete tectonic tree diagram that extends back through the Paleozoic and into the Neoproterozoic,
Mesoproterozoic and Hadean (see Figure 27)
c. Pole to Equator Latitudinal Gradients diagrams (0 – 750 Ma) (see Figure 101)
279
8. Tables and Spreadsheets
a. Copies of all of the tables that appear in this paper.
b. Complete set of Chronotemp tables describing Phanerozoic temperature events (see Table 10; Scotese et al.,
2021).
c. Spreadsheet with Global Average, Tropical, Deep Sea, and Polar temperatures (0 – 540 Ma, see Figure 100;
Scotese et al., 2021)
d. Spreadsheet with Pole-to-Equator Temperatures (0 – 540 Ma; see Figure 99)
e. Multiple spreadsheets with values for land area, sea level versus % land (Figure 16), flooded continental area
(Figure 17), average elevation (Figure 18), Phanerozoic global sea level (Figure 19), comparison of various sea level
curves (Figure 20), Phanerozoic temperature and CO2 levels, (Figure 21), drainage basin size and mean flow length
(Figure 26), latitudinal limit of snow, continental area (Figure 46), ice and glacial deposits (Figure 48)
9. PALEOMAP Paleomagnetic Database
a. Spreadsheet with list of all paleopoles (0 – 1500 Ma) (list of all paleopoles; Elling, 2023)
b. Spreadsheet with list of APW paths in various continental reference frames (see Table 5)
c. Spreadsheet with list of paleopoles used to calculate Cretaceous Global Mean Poles (Elling , 2023)
Part II. Cretaceous Content (145 Ma – 65 Ma) https://doi.org/10.5281/zenodo.10659104
1. Text Items
a. An annotated bibliography with more than 50 sources that feature Cretaceous paleogeographic maps
2. Animations
a. Animation of Cretaceous plate motions (“Plate Tectonics: 200 million years - Today”)
b. Animations of oceanic circulation during the Cretaceous
c. Köppen belt evolution during the Cretaceous
3. Maps (Equirectangular, Mollweide, Robinson, and North & South Orthographic Polar projections)
a. Plate Tectonic Maps (see Figure 2A)
b. Paleogeographic Maps (with 3D relief, modern coastlines and political boundary overlays, .jpg; see Figure 2B)
c. Paleogeographic Maps (no overlays, .bmp)
d. Paleoriver Maps (see Figure 67)
e. Paleotemperature Maps, (See Figure 3A)
f. Köppen Maps, 145 Ma – 65 Ma Ma (See Figure 2C)
g. Rainfall Maps, 145 Ma – 65 Ma Ma (See Figure 3B)
h. Oceanic Circulation Maps, 145 Ma – 65 Ma Ma (See Figure 3C)
4. PaleoDEMs (i.e. paleo-Digitial Elevation Model)
a. netcdf format (.nc), .1˚ x .1˚ and 1˚ x 1˚ resolution
b. text format (.csv), 1˚ x 1˚ resolution
5. Paleoclimate Digital Data Files
a. Spreadsheet with Mean Annual Temperature, (360 x 180 grid, .csv)
b. Spreadsheet with Mean Annual Rainfall, 360 x180 grid, .csv)
6. Miscellaneous Diagrams
a. High resolution versions of all of the figures that appear in the paper.
7. Tables and Spreadsheets
a. Copies of all of the tables that appear in the paper.
b. Spreadsheet with Global Average, Tropical, Deep Sea, and Polar temperatures (145 Ma – 65 Ma; see Figure 100)
280
c. Spreadsheets with values for Cretaceous sea level (Figure 45), number of mid-ocean ridges and subduction zones
during the Cretaceous (Figure 47) average river length and number of rivers during the Cretaceous (Figure 66)
8. PALEOMAP Paleomagnetic Database
a. Spreadsheet with list of paleopoles used to calculate Cretaceous Global Mean Poles