Download as pdf or txt
Download as pdf or txt
You are on page 1of 229

Atlas of the Underworld

Paleo-subduction, -geography, -atmosphere and -sea level reconstructed


from present-day mantle structure

Atlas van de Onderwereld: paleo-subductie, -geografie, -atmosfeer en -zeespiegel gereconstrueerd


met de huidige mantelstructuur

(met een samenvatting in het Nederlands)

Proefschrift

ter verkrijging van de graad van doctor aan de Universiteit Utrecht op gezag van de rector
magnificus, prof. dr. G.J. van der Zwaan, ingevolge het besluit van het college voor promoties in het
openbaar te verdedigen op woensdag 25 oktober 2017 des middags te 12.45 uur

door

Douwe George van der Meer


geboren op 2 augustus 1977 te Terneuzen
Promotor: Prof. Dr. W. Spakman

Copromotor: Dr. D.J.J. van Hinsbergen


Utrecht Studies in Earth Sciences 139

Atlas of the Underworld


Paleo-subduction, -geography, -atmosphere and -sea level reconstructed
from present-day mantle structure

Douwe G. van der Meer

Ruislip 2017
The research for this thesis was carried out over the past 15 years at various home addresses in
Assen, Groningen, Delft (Netherlands), Ruislip (United Kingdom) and during study leave periods
at Norges Geologiske Undersøkelse (Trondheim, Norway) and Utrecht University (Netherlands).
In addition, significant work was accomplished during flights in international airspace. The work
was conducted under the programme of the research Institute of Earth Sciences of the Faculty of
Geosciences.

Promotor:
Prof. Dr. W. Spakman: Department of Earth Sciences, Utrecht University, The Netherlands
Centre of Earth Evolution and Dynamics, University of Oslo, Norway
Copromotor:
Dr. D.J.J. van Hinsbergen: Department of Earth Sciences, Utrecht University, The Netherlands


Dissertation committee:
Prof. Dr. John Suppe: Department of Earth and Atmospheric Sciences, University of
Houston, USA
Department of Geosciences, Princeton University, USA
Department of Geosciences, National Taiwan University, Taipei,
Taiwan
Prof. Dr. Chris Scotese: Department of Earth and Planetary Sciences, Nortwestern University,
Evanston, USA
Field Museum of Natural History, Chicago, USA
Prof. Dr. Henk Brinkhuis: Laboratory of Palaeobotany and Palynology, Utrecht University, The
Netherlands
Koninklijk Nederlands Instituut voor Onderzoek der Zee - NIOZ,
Den Burg, The Netherlands
Prof. Dr. Arwen Deuss: Department of Earth Sciences, Utrecht University, The Netherlands
Dr. Wouter Schellart: Faculty of Science, Vrije Universiteit, Amsterdam, The Netherlands

ISBN/EAN: 978-90-6266-482-5
ISSN: 2211-4335

Copyright © 2017 Douwe G. van der Meer

Niets uit deze uitgave mag worden vermenigvuldigd en/of openbaar gemaakt door middel van
druk, fotokopie of op welke andere wijze dan ook zonder voorafgaande schriftelijke toestemming
van de uitgevers.

All rights reserved. No part of this publication may be reproduced in any form, by print or photo
print, microfilm or any other means, without written permission by the publishers.

Printed in the Netherlands by Ipskamp Printing, Enschede.


I have come to believe that this is a mighty continent which was hitherto unknown. I propose to
construct a new chart for navigating, on which I shall delineate all the sea and lands of the ocean
in their proper positions under their bearings; and further, I propose to prepare a book, and to put
down all as it were in a picture.

Christopher Columbus, 1492

5
6
Contents

Introduction 9

Chapter 1: Towards absolute plate motions constrained by lower-mantle slab remnants 13

Chapter 2: Intra-Panthalassa Ocean subduction zones revealed by fossil arcs and mantle
structure 21

Chapter 3: Plate tectonic controls on atmospheric CO2 levels since the Triassic 31

Chapter 4: Neoproterozoic-Phanerozoic sea level reconstruction from Strontium isotopes 43

Chapter 5: Atlas of the Underworld: slab remnants in the mantle, their sinking history,
and a new outlook on lower mantle viscosity 63

Epilogue 177

References 179

Samenvatting in het Nederlands 214

Dankwoord (Acknowledgements) 217

Curriculum Vitae 218

Publication List 219

Appendix Table; interpreted slabs and their orogens 220

Appendix Figures; paleo-subduction zone interpretation from seismic tomography 222

7
8
Introduction
Plate tectonic and paleogeographic reconstructions simulate the dynamic and kinematic make up
at the Earth’s surface (Blakey, 2003, 2008; Golonka 2007a-c, 2009ab; Müller et al., 2008, 2013;
Scotese, 2014a-f, 2016; Seton et al., 2012; Stampfli et al. 2004; Torsvik et al., 2008, 2014) and are
more and more used as input for geodynamic and paleoclimate studies. In geodynamic studies,
plate tectonic reconstructions constitute surface boundary conditions for plate tectonic and mantle
convection modelling (e.g. Shephard et al. 2012, Williams et al. 2015) while lithosphere subduction
as predicted by a plate reconstruction can be used as guidance for constraining the local downward
mantle flow (e.g. Bower et al. 2015). Paleogeographic reconstructions provide the basic context for
paleoclimate and paleo-environments in terms of land-sea distributions and first-order bathymetry.
Plate reconstructions can, in addition, also be used to quantify tectonically driven geochemical
cycles such as ‘mantle’ inputs into the atmosphere and seawater, of chemical compounds such
as CO2 and strontium (Berner, 1994). By estimating the average ocean floor age from crustal
spreading, paleogeographic reconstructions are also useful to determine the plate tectonic influence
on paleo-sea level (Müller et al., 2008; Spasojevic and Gurnis, 2012; Verard et al., 2015).
Going back in geological time, reconstructions of relative plate positions become more
uncertain due to decreasing data density, in particular due to a lack of preserved oceanic crust. Prior
to the Early Jurassic this lack effectively results in a ‘World Uncertainty’ of 70% (Torsvik et al.,
2010) in paleogeographic reconstructions. Therefore, global oceanic crustal spreading (important
for sea level reconstruction), and plate tectonic activity (important for volcanic degassing of CO2),
are increasingly less-well constrained further back in time. The locations and ages of intra-oceanic
subduction zones, known from the geological record to have existed in the Panthalassa and Tethys
Oceans, are other examples of which investigation suffers from lack of observational constraints.
These subduction zones are therefore difficult to reconstruct and have been largely omitted from
global paleogeographic reconstructions.
When moving back in geological time, additional uncertainties occur when relative plate
reconstructions are being cast in a so-called mantle frame leading to the concepts of ‘absolute’
position and ‘absolute’ plate motions in a reference frame attached to the deep mantle. Plate
reconstructions in a mantle frame are key for linking mantle processes (e.g. mantle flow, plumes,
melting, and subduction) to the plate tectonic evolution and therefore are a prerequisite for any
dynamic global modelling. Classical techniques to reconstruct absolute plate positions and motions
using hotspot tracks cannot be applied prior to the Cretaceous due to the lack of preserved tracks.
In addition, lateral motions, or deflections, of plumes in the mantle are uncertain (Doubrovine et al.
2012). Modern hypotheses, however, regarding to the deep mantle origin of plumes, Large Igneous
Provinces and kimberlites, provide novel ways of reconstructing past plate positions in the mantle
reference frame prior to the Cretaceous (Burke et al. 2008, Torsvik et al. 2010, Torsvik et al. 2014),
but still with limited data sets.
This thesis aims at searching for new ways of constraining plate tectonic, paleo-geographic,
-atmosphere and -sea level reconstructions, by an extensive investigation of paleo-subduction.
To this end, I explore evidence for paleo-subduction in seismic tomography models of present-
day mantle structure and how this may allow for a new avenue towards reconstructing the plate
tectonic history of our planet. These tomography models exploit seismological data for imaging
of 3-D variations of seismic velocity in the mantle. In the upper mantle, seismically fast regions
coincide with current subduction zones, in part also evident from the depth distribution of

9
earthquake hypocenters. In the lower mantle only few positive wave speed anomalies could
be interpreted so far as remnants of past subduction by identification of their paleo-subduction
zones in the geological record (e.g. Grand et al., 1997; van der Voo et al., 1999a,b; Hafkenscheid
et al., 2006; van Hinsbergen et al., 2005). Potentially, the lower mantle can therefore harbour a
more extensive record of past subduction. Hence, to explore this potential the core of this thesis
concerns to systematically identify the global occurrence of slab-like anomalies from seismic
tomography models for which I use the P-wave velocity model UU-P07 of Amaru (2007) and the
S-wave velocity models S40RTS (Ritsema et al., 2011) and SL2013sv (Schaeffer and Lebedev,
2013). These tomography models provide independent images of mantle structure because they
are based on totally different seismological data and inversion methods. I extensively correlate
imaged slab remnants in the deep mantle with the geological record of past subduction episodes
further exploring the research avenue opened by research such as van der Voo et al. (1999a,b), van
Hinsbergen et al. (2005), or Hafkenscheid et al. (2006). This has led to the following chapters,
meanwhile published since 2010, of my thesis.

Outline of thesis

Chapter 1: Defining the methodology


This chapter is concerned with the first extensive interpretation of lower mantle structure based on
tomography model UU-P07. Emphasis has been on identifying subducted lithosphere, ‘slabs’, in
space and time, in a consistent manner by correlating the location, orientation and trend of mantle
anomalies with the surface geological record of subduction such as start and end of mountain
building or regional arc volcanism. Three previously described lower mantle slabs (Farallon,
Aegean Tethys, Mongol-Okhotsk) were selected as anchor slabs to guide the interpretation which
was further based on the essential starting point that in general, deeper slabs subducted earlier
than shallower slabs, yielding an average subduction age-depth relation. Initially 28 slabs were
characterised (Chapter 1), and seven years later this culminated in the Atlas of the Underworld
with ~100 slabs (Chapter 5). Chapter 1 provides the backbone philosophy and methodology on
how to use tomographic images of the remnants of past subduction to constrain paleo-longitude in
relative plate reconstructions: subduction zones in plate tectonic reconstructions are placed above
the locations of corresponding subducted slabs in the mantle – the so-called ‘slab-fitting’ procedure,
leading to a new mantle frame, here called the ‘subduction frame’.

Chapter 2: Application of the method to constrain intra-oceanic subduction in paleo-


geographic reconstruction
The methodology provided in Chapter 1 offers a new way of constraining the absolute position
of existing plate reconstructions relative to the mantle through slab-fitting. Using the time-depth
relationship established in Chapter 1, Chapter 2 explores to what extent intra-oceanic subduction
zones can be reconstructed in the Panthalassa Ocean, by correlating exotic terranes, accreted to the
Asian and American margins (i.e. Nokleberg et al., 2000), to (central) Panthalassa locations on the
basis of the imaged mantle structure and plate tectonic transport estimates.

Chapter 3: Using mantle structure to constrain past atmospheric CO2 levels


With the global identification of slabs of Chapters 1 and 2, an estimate is made of the total slab
length as a function of depth. With the established average age-depth conversion, this is converted
to subduction zone length through time. This is calibrated against a crustal production curve for

10
the last 140 Ma (Coltice et al. 2013). Assuming the carbon content and convergence velocity of
subducting plates did not change significantly, subduction zone length through time may provide
a first-order estimate of volcanic degassing into the atmosphere as well as mantle strontium input
into the hydrosphere. The normalized subduction length through time was used to calculate the
paleo-strontium record and as input for the GEOCARBSULF (Berner, 2006) carbon-cycle model
to estimate plate tectonic controls on atmospheric CO2 levels for the Mesozoic and Cenozoic.

Chapter 4: A paleo-sea level reconstruction based on plate tectonic activity estimates


The strontium record (Chapter 3), corrected for weathering inputs (Goddéris et al., 2014), provides
an estimate of plate tectonic activity that can be correlated with oceanic spreading and subduction.
Variations in the corrected strontium record, represent oceanic spreading variations, which dictate
the crustal age, ocean depth (Parsons and Sclater, 1977; Stein and Stein, 1992) and therefore ocean
water storage volume. With the ocean storage volume history calculated, at times of lower storage
volume, excess water was redistributed over the combined areas of both oceans and flooded shelfs.
The resulting novel absolute sea-level curve is subsequently compared with ‘conventional’ sea level
curves derived from sequence-stratigraphic methods (i.e. Hallam, 1984; Snedden & Liu, 2010; Vail
et al., 1977) and plate motion models (Müller et al. 2008; Verard et al., 2015).

Chapter 5: The Atlas of the underworld, a catalogue of slab remnants in the mantle imaged
by seismic tomography and their geological interpretation
This final chapter is the culmination of 15 years of (part-time) research during which this thesis was
developed: the Atlas of the underworld, documenting the identification and association of ~100
slabs and slab remnants in the upper and lower mantle with their geological provenance. This is
an expansion of the slab catalogue developed in the first chapter. This extended catalogue provides
a foundation for further studies into developing a subduction constrained mantle frame, while
average sinking velocities may provide new constraints on mantle viscosity, as well as enabling a
next assessment of plate tectonic activity through time, volcanic CO2 degassing and paleo-sea level.

11
12
Chapter 1

Towards absolute plate motions constrained by


lower-mantle slab remnants
This chapter has been published as: van der Meer, D.G., Spakman, W., van Hinsbergen, D.J.J., Amaru, M., and
Torsvik, T.H., 2010; Towards absolute plate motions constrained by lower-mantle slab remnants, Nature
Geoscience 3, p. 36-40. Supplementary figures and table belonging to the original published article have been
omitted from this thesis and have been partly incorporated in Chapters 2, 5 and Appendix.

Since the first reconstruction of the supercontinent Pangea, key advances in plate tectonic
reconstructions have been made (Müller et al., 1993; Norton, 1996; Scotese, 2001; Stampfli et al.,
2004; Steinberger and Torsvik, 2008; Torsvik et al., 2008a). Although the movement of tectonic
plates since the start of the mid-Cretaceous period (100 million years (Myr) ago) is relatively well
understood (Müller et al., 1993; Norton, 1996), the longitudinal position of plates before this
period is not constrained at all. Here, we use a global mantle tomography model (Amaru, 2007) to
estimate the longitude of past oceanic subduction zones. We identify 28 remnants of oceanic plates
that were subducted into the lower mantle and link these to the mountain building zones from
which they are likely to have originated. Assuming that these remnants sank vertically through the
mantle, we reconstruct the longitude at which they were subducted. Our estimates for the location
of the subduction zones are offset by up to 18 degrees compared with plate tectonic reconstructions
for the corresponding period. We did not detect oceanic plate remnants from the Carboniferous
period (300-360 Myr ago), or before, suggesting that the tomographic visibility of subduction is
limited to the past 300 Myr.

Introduction

Since the first qualitative plate reconstruction of the supercontinent Pangea was determined by
fitting paleoclimate belts and modern continental margins, key advances in plate reconstructions
have been made with the development and use of paleomagnetic apparent polar wander paths,
ocean floor magnetic anomalies and hotspot reference frames (Müller et al., 1993; Norton, 1996),
leading to global plate tectonic reconstructions (Scotese, 2001; Stampfli et al., 2004; Torsvik et al.,
2008a). Absolute plate motion models have often been based on assumed hotspot fixity and are
well constrained only up to the Cretaceous period owing to the lack of any preserved older oceanic
hotspot tracks (Müller et al., 1993; Norton, 1996). These, and other models, offer no control on
absolute paleolongitude before the Cretaceous.
Seismic tomography studies of the mantle have allowed for increasingly detailed correlations
between deep mantle structure, mostly focused on presumed remnants of subducted plates and
plate tectonic evolution (De Jonge et al., 1993; Grand et al., 1997; Hafkenscheid et al., 2006; Hall
and Spakman, 2002; Schellart et al., 2009; Sigloch et al., 2008; van der Voo et al., 1999a, 1999b;
van Hinsbergen et al., 2005). This, however, has not led to strong constraints on absolute plate
motion. Recently, correlations between deep, presumably hot and dense mantle heterogeneities at

13
Figure 1.1. Age-depth curve of interpreted slabs. The symbols indicate slab limits, with error bars
in age-depth interpretation. The grey box represents average sinking rate (12 mm/yr). The colour-
coded slabs are previously described marker slabs.

the core-mantle boundary and large igneous provinces were obtained from a plate reconstruction
(Steinberger and Torsvik, 2008; Torsvik et al., 2008a), leading to possible predictions of absolute
paleolongitude for the entire Phanerozoic eon (Torsvik et al., 2008b). This reconstruction model,
however, assumes zero longitude motion for Africa before the Cretaceous.

Interpretation of slab remnants

Here, independently of any reconstruction model, we carry out a global interpretation of positive
seismic anomalies in the lower mantle based on the assumptions that these reflect relatively cold
remnants of subducted lithosphere (De Jonge et al., 1993; Grand et al., 1997; Hafkenscheid
et al., 2006; Hall and Spakman, 2002; Schellart et al., 2009; Sigloch et al., 2008; van der Voo et
al., 1999a, 1999b; van Hinsbergen et al., 2005), and that slabs, once detached, sink more or less
vertically in the mantle (Grand et al., 1997; Steinberger, 2000; van der Voo et al., 1999a, 1999b)
marking the location of their former subduction zones. Geological evidence for former subduction
is constituted by orogens created in plate convergence zones, often comprising the remnants of a
volcanic arc. Studies comparing the amount of Tethyan subducted material predicted by geological
reconstructions with volumes of subducted lithosphere imaged in the mantle by seismic tomography
(Hafkenscheid et al., 2006; van Hinsbergen et al., 2005) suggest a first-order correlation between
the onset and end of subduction and the onset and end of orogenesis, respectively. In line with this
inference, we assume that to first order, the age of the base and top of a lower-mantle slab can be
correlated in time with the onset and end, respectively, of the associated orogeny. Timing errors of
the order of 10-25 Myr are inferred from the geological literature used for interpretation (Cawood
and Buchan, 2007; Golonka et al., 2003; Nokleberg et al., 2000; Stampfli et al., 2004) and include
uncertainties in upper-mantle subduction history.
Using this rationale, we interpret lower-mantle slab remnants from the tomographic model
UU-P07 (Amaru, 2007), the successor of BSE98 (Bijwaard et al., 1998), and correlate these with

14
their corresponding orogenies. Identification, depth of top and bottom, timing and images of
slab remnants are documented extensively in the Appendix. The Farallon, Mongol-Okhotsk and
Aegean Tethys slabs have already been studied extensively and correlated to geologic events and
act here as anchor points for the global interpretation of the lower mantle. The still-subducting
Aegean Tethys slab remnant was shown to reach a depth of 2,000±100 km (Bijwaard et al., 1998;
Hafkenscheid et al., 2006; van Hinsbergen et al., 2005). The onset of subduction in the Aegean
region is inferred to approximate 171±5 Myr, as represented by metamorphic soles below the oldest
Aegean ophiolites (Liati et al., 2004). The Mongol-Okhotsk slab remnant is located below northern
Siberia from the base of the mantle up to mid-mantle and subducted in the Middle Triassic/
Middle Jurassic (Golonka et al., 2003; Stampfli et al., 2004; van der Voo et al., 1999b). The Farallon
slab remnant (Grand et al., 1997) is located below eastern Laurentia from the deep lower mantle to
the upper part of the lower mantle. The slab has been interpreted to represent eastward-subducted
Farallon oceanic lithosphere, but estimates of the start of subduction of the slab vary from Late
Cretaceous (Grand et al., 1997) to Late Jurassic (Sigloch et al., 2008). However, the geological
record proves that subduction at the western continental margin of Laurentia initiated earlier, in
the Early Jurassic (DeCelles et al., 2009; Nokleberg et al., 2000). This is adopted here as the start
of subduction and related to the deepest part of the Farallon slab remnant. In the upper mantle
the Farallon slab disintegrates into smaller fragments, which have been associated with subduction
during the Laramide orogeny (80-40 Myr) to present (Sigloch et al., 2008; van der Lee and Nolet,
1997). The total slab data set, covers about half of the imaged positive wave-speed anomalies. For
the remainder, notably in the deep mantle of the Indian and Pacific oceans, we could not date the
slab remnants with the geological literature, or the tomographic model resolution was insufficient
to warrant a useful quantitative interpretation.

Slab sinking velocity

Plotting the depths of the top and bottom of all slab remnants against the corresponding tectonic
ages (Figure 1.1) demonstrates an average slab sinking velocity in the lower mantle of 12±3 mm/
yr. This result is obtained independent of mantle rheology and thus can serve as a new constraint
in the determination of lower-mantle viscosity. The sinking rate is slower than inferred by mantle
flow estimates using Mesozoic-Cenozoic subduction models (Ricard et al., 1993; Steinberger,
2000), which depend on assumed mantle rheology, but in agreement with other tomographic
interpretations of 1-2 cm/yr (Hafkenscheid et al., 2006; Schellart et al., 2009; van der Voo et al.,
1999a). Figure 1.1 shows that within the uncertainty estimates, the timing of paleo-subduction can
vary by ±25 Myr from the average, and slab remnants from paleo-subduction of any given age can
be found ±300 km from the corresponding average depth. Here we are primarily concerned with
first-order inferences on paleolongitude and lower-mantle sinking rates on a global scale.
We find that all slab remnants in the lowermost mantle correspond to subduction systems of
Triassic-Permian age (Figure 1.1). No geological argument is found for correlating the deepest
positive wave-speed anomalies to Carboniferous, or older subduction systems. We infer that the
tomographic visibility of subduction history may be restricted to the past ~300 Myr of Earth
evolution. Using seismic-anomaly survival times, defined as a 50-100°C temperature anomaly, from
a recent study ( Jarvis and Lowman, 2007), a slab maximum survival time of 300 Myr for slabs
may indicate an increase in lower-mantle viscosity by a factor 100-300 relative to the upper mantle.
The numerical mantle flow model ( Jarvis and Lowman, 2007) however, is based on mid-mantle

15
Figure 1.2. Spatial longitude correction. North Pole orthographic projections. A) Tomographic depth slice (Amaru, 2007) at 1,900 km, colour
scale red (-0:4%) to blue (+0:4%) with present-day continents. Interpreted slabs are outlined in purple: Far, Farallon; Aeg, Aegean Tethys; MO,
Mongol-Okhotsk. B) Unmodified reconstruction at 160 Myr. Offsets exist between the three slabs (purple) and their corresponding subduction
complexes/sutures (red outlines). C, To obtain an improved fit between subduction complexes/sutures (green) and slab locations (purple), the
plate tectonic reconstruction was shifted 18° westward.

16
Figure 1.3. Longitude correction with time. The blue triangles note a good global fit between
slabs and continental plate margins. The orange squares note fits where some slabs are offset
to the margins. The blue line represents the applied longitude correction to the plate tectonic
reconstruction, based on the moving average of good global fits. The red dashed lines and arrows
indicate the uncertainty, varying with latitude, to a lateral uncertainty of 500 km in the surface
location of a paleo-subduction zone.

observation and subducting continuous slabs, whereas our survival time estimate pertains to the
lowermost mantle and detached slab remnants.
Tomographic invisibility can be caused by thermal assimilation, and/or mantle mixing, for
example, by recycling in mantle plumes as indicated by geochemical signatures (Chauvel et al.,
2008). Recycling times of subducted oceanic lithosphere can be as short as a few hundred million
years as indicated in the isotope geochemistry of some mantle plumes (Hauff et al., 2000; Schaefer
et al., 2002). These short recycling times suggest that the deepest slab remnants may be the elusive
source of recycled slab material observed in plumes, mixed with components of up to Archaean age
(Schaefer et al., 2002). Our observation of tomographic invisibility provides new input for future
thermochemical dynamic modelling studies of slab survival in the deepest mantle.

Paleo-longitude correction

To illustrate the implications for paleo-longitude shifts of continents back to the Permo-Triassic,
we use a plate tectonic reconstruction (Torsvik et al., 2008a), recently corrected for true polar
wander (Steinberger and Torsvik, 2008) between 100 and 300 Myr. From 12 depth slices in the
lower mantle, corresponding to subduction during the time frame 40-260 Myr, we determine the
shift of paleo-longitudes of continental blocks from the inferred paleo-subduction zone positions
of the 28 slab remnants. As an example, we show how the three anchor slabs (Grand et al., 1997;
van der Voo et al., 1999b; van Hinsbergen et al., 2005) constrain the plate tectonic reconstruction
(Steinberger and Torsvik, 2008; Torsvik et al., 2008a) (Figure 1.2). A depth slice at 1,900 km in the
tomographic model (Amaru, 2007) corresponds to ~160 Myr of the plate tectonic reconstruction
(Figure 1.1). The unmodified reconstruction shows the Mongol-Okhotsk slab (van der Voo et al.,

17
Figure 1.4. Longitude-corrected plate tectonic reconstructions. The colour scale is given in Figure 1.2, A good fit is obtained between the
tomographic depth slices (Amaru, 2007) at 1,325 km (A) and 2,650 km (B) depth and the modified plate tectonic reconstructions (Steinberger
and Torsvik, 2008; Torsvik et al., 2008a) at 120 Myr (17° corrected) (C) and 240 Myr (10° corrected) (D). Tectonic interpretation: lines with
triangles, subduction zone of the slab data set (red) and other slabs with only a qualitative interpretation (orange); green line, transform zone;
yellow double line, spreading ridge.

18
1999b) to be located below central Asia, the Aegean Tethys slab (van Hinsbergen et al., 2005)
below central Europe and the Farallon slab (Grand et al., 1997) to be completely below the
eastern Panthalassa Ocean (Figure 1.2). All slabs are positioned too far west with respect to the
subduction location inferred by the geological record (Grand et al., 1997; van der Voo et al., 1999b;
van Hinsbergen et al., 2005). A 16°-18° westward shift of the reconstruction provides the best fit,
between slab remnants at depth and the surface location of subduction. Clearly, some offsets remain
between our paleo-subduction positions inferred from tomography and the modified plate tectonic
construction (Figure 1.2c). These discrepancies approach the uncertainties in the inferred surface
location of paleo-subduction with an estimated lateral spatial r.m.s. uncertainty of 250-500 km
resulting from the effects of unknown slab dip (300-700 km), lower-mantle thickening of slabs
(200-400 km) and tomographic imaging error (200-400 km) of the interpreted medium to well-
imaged slab remnants. Other uncertainties are due to complexities in the slab subduction history,
for example, the northern Farallon slab is related to subduction at the continental margin, but its
southern extent may have been caused by simultaneous intra-oceanic subduction (Nokleberg et al.,
2000).
An important overall observation is that, at each time slice considered, the relative position
of continental fragments in the plate reconstruction (Steinberger and Torsvik, 2008; Torsvik et
al., 2008a) proves largely consistent with our inferred relative positions of paleo-subduction zones
(Figure 1.2). This gives confidence in the plate reconstruction (Steinberger and Torsvik, 2008;
Torsvik et al., 2008a) used and in our inferred paleo-subduction zone configurations. This also
allows us to concentrate on longitude shifts per time slice of the entire assembly of continental
fragments. These are summarized in Figure 1.3, whereas Figure 1.4 serves to illustrate our
interpretation process. In Figure 1.4, we also speculate on the paleo-position of intra-oceanic
blocks, spreading ridges and transform zones, but this is not essential for our analysis of absolute
paleolongitude.
On the basis of slab remnants between 45°N and 45°S, we find that paleolongitudes gradually
depart between 40 and 120 Myr, amounting to a westward longitude shift of 18°-20° between
120 and 180 Myr, after which a gradual decrease occurs to 8° at 260 Myr (Figure 1.3). Owing
to uncertainties, this departure is significant from 80 to 260 Myr and provides new constraints
on the paleolongitude of the entire continental-plate configuration through time. Figure 1.4 shows
that the continents on the Northern Hemisphere are constrained by slabs on either side of the
paleo-Pacific Ocean (for example, Farallon slab (Grand et al., 1997)) and by Tethys Ocean slabs
(for example, Aegean Tethys slab (van Hinsbergen et al., 2005)). The longitude correction decreases
to 10° in the Late Permian-Triassic (200-260 Myr). Bordered by slabs at its western and eastern
side (for example, Mongol-Okhotsk slab (van der Voo et al., 1999b)), Pangaea is also constrained in
absolute longitude (Figure 1.4).
Our global atlas of subduction remnants agrees through geological time with the relative
positions of continental plates and agrees within 18°-20° with the paleolongitude of the entire
continental assembly as given by the plate reconstruction (Steinberger and Torsvik, 2008; Torsvik
et al., 2008a) used for comparison. The latter reconstruction also proposes absolute paleo-longitudes
of continental plates (Torsvik et al., 2008b). Given the uncertainties underlying both approaches,
we have obtained a rather close match between independently proposed paleolongitudes through
time. The required longitude corrections do not significantly affect the correlation between mantle
heterogeneities at the coremantle boundary and large igneous provinces (Torsvik et al., 2008b).
The longitude correction pattern of Figure 1.3 may reflect Pangea (Africa) wandering in
longitude, but does require further investigation, eventually leading to an integration of both

19
approaches and a further quantification of plate reconstructions. Here, we have focused on new
inferences on absolute plate motion, lower-mantle sinking rates and the tomographic visibility of
slab remnants, potentially useful as constraints on mantle rheology and on geodynamic modelling
of crustmantle evolution.

Methods

We selected only imaged anomalies in the UU-P07 tomographic model (Amaru, 2007) with peak
velocity amplitudes above +0.4%, and overall amplitudes above +0.2% to delineate slab boundaries
(Hafkenscheid et al., 2006). Subduction zones associated with large oceanic basins are expected
to be of considerable lateral extent; therefore, the interpretation is further restricted to laterally
elongated deep-mantle anomalies, leading to the data set of 28 slab remnants (incorporated
in Chapter 5). The Appendix provides our global atlas of slab remnant interpretations, which
may prove useful for future tomographic and plate tectonic studies. The radial error in the depth
interpretation varies from ±100 km in the shallow lower mantle to ±200 km in the deeper lower
mantle.
The set of 28 slabs is subdivided into three confidence categories. Category I slabs have
previously been interpreted in tomographic studies and correlated to their respective orogens
(Grand et al., 1997; Hafkenscheid et al., 2006; Sigloch et al., 2008; van der Voo et al., 1999a, 1999b;
van Hinsbergen et al., 2005). Age uncertainties are derived from geological literature. Vertical
spatial uncertainties are conservative (upper) estimates of image blurring effects.Slabs in category
II are found in the vicinity of a category I slab, providing a paleogeographic link to the latter,
assuming they have been created and shaped in the same boundary region between two converging
plates. Candidates for corresponding orogens are identified from paleo-geographic reconstructions
(Cawood and Buchan, 2007; Golonka et al., 2003; Stampfli et al., 2004), which are primarily
determined from the synthesis of geological and paleo-magnetic data. For slabs of category III,
there is no apparent relation with other slabs, or the interpretation of subduction remnants has led
to relatively larger uncertainties in subduction timing and depth as compared with the first two
categories.
The initial identification of category II and III slabs from paleo-geographic reconstructions is
tied in with the interpretations of category I slabs. The subsequent estimates for top and bottom
depth from the tomographic model and the more precise timing of start and end of subduction
from the geological literature are independent of this initial identification process.
To test the data-set robustness, the time-depth data set was split in various ways: separately
considering the top and base of slab remnants, per slab remnant confidence category and per region.
The longitude correction was estimated by correlating the plate tectonic reconstruction in 20 Myr
steps with the closest available depth slice (tomographic layer), based on the average time-depth
conversion as indicated by the slab data set. Global longitude shifts were applied in 2° steps and
the fit quality was assessed between the active paleo-margins at surface and slab remnants at depth.
Latitudes of continental fragments were not changed, except for the North China block in the
Permo-Triassic. Mesozoic Europe has been simplified by excluding small continental fragments at
its southern margin. Where flat-slab orogenies occurred, the intra-plate edge of crustal deformation
was taken to match the dispersed slabs remnants. Locations of spreading ridges and transform
zones have also been added, but are based on speculation.

20
Chapter 2

Intra-Panthalassa Ocean subduction zones


revealed by fossil arcs and mantle structure
This chapter has been published as: van der Meer, D.G., Torsvik, T.H., Spakman, W., van Hinsbergen, D.J.J., and
Amaru, M.L., 2012, Intra-Panthalassa Ocean subduction zones revealed by fossil arcs and mantle structure,
Nature Geoscience 5, p. 215-219. Supplementary figures belonging to the original published article have been
omitted from this thesis and have been partly incorporated in Chapter 5 and Appendix.

The vast Panthalassa Ocean once surrounded the supercontinent Pangaea. Subduction has since
consumed most of the oceanic plates that formed the ocean floor, so classic plate reconstructions
based on magnetic anomalies can be used only to constrain the ocean’s history since the Cretaceous
period (Engebretson et al., 1985; Müller et al., 2008), and the Triassic-Jurassic plate tectonic
evolution of the Panthalassa Ocean remains largely unresolved (Nokleberg et al., 2000; Nur and
Ben-Avraham, 1977). Geological clues come from extinct intra-oceanic volcanic arcs that formed
above ancient subduction zones, but have since been accreted to the North American and Asian
continental margins (Nokleberg et al., 2000). Here we compile data on the composition, the timing
of formation and accretion, and the present-day locations of these volcanic arcs and show that
intra-oceanic subduction zones must have once been situated in a central Panthalassa location in
our plate tectonic reconstructions (Steinberger and Torsvik, 2008; Torsvik et al., 2008; van der Meer
et al., 2010). To constrain the paleoposition of the extinct arcs, we correlate them with remnants of
subducted slabs that have been identified in the mantle using seismic-wave tomographic models
(Amaru, 2007; Ritsema et al., 2011). We suggest that a series of subduction zones, together called
Telkhinia, may have defined two separate paleo-oceanic plate systems – the Pontus and Thalassa
oceans. Our reconstruction provides constraints on the paleolongitude and tectonic evolution of
the Telkhinia subduction zones and Panthalassa Ocean that are crucial for global plate tectonic
reconstructions and models of mantle dynamics.

Introduction

The spatial uncertainty in reconstructions of oceanic basins increases the further we go back in
time, because oceanic plates subduct. Plate reconstructions for the Triassic-Jurassic history of the
Panthalassa Ocean, once surrounding the Pangaea supercontinent, are based on the premise that
the Pacific plate formed in the Middle Jurassic as a triangle (Müller et al., 2008), originating from
a triple junction between the Farallon, Phoenix and Izanagi plates (Figure 2.1). The preserved area
of Jurassic paleoplates represents only a small portion (<10%) of the oceanic crust of the Jurassic
Panthalassa Ocean. The remaining part, along with pre-Jurassic oceanic plate configurations, is
entirely unconstrained.
However, the geology of the circum-Panthalassa continental margins provides evidence that
early Mesozoic subduction-related volcanism occurred in the Panthalassa Ocean (Isozaki et al.,
1990; Nokleberg et al., 2000). On subduction, continental fragments, volcanic arcs, ophiolites and

21
Figure 2.1. Present understanding of the Panthalassa Ocean. Modified plate tectonic reconstruction
(Steinberger and Torsvik, 2008; Torsvik et al., 2008; van der Meer et al., 2010) centred at the 180°
meridian at A) the present time and B) 200 Myr ago (earliest Jurassic). Continental blocks in
absolute plate reference frames: slab-fitted frame (van der Meer et al., 2010), light grey; hybrid
true-polar-wander-corrected frame (Steinberger and Torsvik, 2008; Torsvik et al., 2008, 2010), dark
grey. Light blue denotes preserved Jurassic (140-175 Myr old) crust of the Pacific plate. Yellow
zigzag shows the presumed spreading ridge. White ellipses represent terranes: AK, Anadyr-Koryak;
KO, Kolyma-Omolon; ON, Oku-Niikappu; S, Stikinia; WR, Wrangellia

22
ocean-floor sediments collided with or accreted to the continental margins of, in particular, western
North America, far-east Asia and Japan. Marine microfossils in these sediments and paleomagnetic
data constrain their age and paleolatitude, but their paleolongitudes remain undetermined.
Paleogeographic reconstructions therefore show uncorrelated intra-Panthalassa exotic terranes
essentially unconstrained in the early Mesozoic Panthalassa Ocean (Golonka, 2007a; Nokleberg
et al., 2000; Shi, 2006). Notably, some of these terranes preserve relics of fossil Triassic and Jurassic
volcanic arc complexes that accreted many tens of millions of years after their extinction to circum-
Pacific continents (Nokleberg et al., 2000; Shi, 2006; Ueda and Miyashita, 2005). This shows that
Triassic and Jurassic subduction zones existed away from continental margins in the Panthalassa
Ocean. As subducted slab remnants (that is, subducted oceanic crust and lithosphere) associated
with these subduction zones might still reside in the lower mantle, the structure of the sub-Pacific
mantle may provide paleolongitudes of intra-oceanic subduction, allowing significant improvement
of plate tectonic reconstructions of the early Mesozoic Panthalassa Ocean.

Exotic accreted terranes indicate paleo-subduction

We first explore the geological evidence for the locations of these intra-oceanic subduction zones.
The best-constrained exotic terranes with intra-Panthalassa volcanic arc remnants of Triassic-
Jurassic age are the Kolyma-Omolon superterrane (northeast Asia) (Nokleberg et al., 2000; Oxman,
2003; Stone et al., 2003), the Anadyr-Koryak terrane (east Asia) (Filatova and Vishnevskaya,
1997; Harbert et al., 2003; Nokleberg et al., 2000), the Oku-Niikappu terrane ( Japan) (Ueda and
Miyashita, 2005) and the Wrangellia and Stikinia terranes (western North America) (Kent and
Irving, 2010; Nokleberg et al., 2000) (Figure 2.1).
Kolyma-Omolon consists of the cratonic Omulevka and Omolon terranes, which rifted
away from Siberia in the Carboniferous period (Nokleberg et al., 2000; Stone et al., 2003), and is
associated with a fossil Triassic-Jurassic subduction complex of arc volcanics, accretionary wedges
and ophiolite belts (Nokleberg et al., 2000; Stone et al., 2003). Paleomagnetic data (Nokleberg et
al., 2000; Stone et al., 2003) indicate that the Omulevka and Omolon terranes remained stationary
at 40°-55°N in Permian-Triassic times, followed by rapid northward motion until the Late Jurassic-
Early Cretaceous collision with Siberia at high polar latitudes (Nokleberg et al., 2000; Oxman,
2003; Stone et al., 2003).
Radiolarian assemblages from Mesozoic exotic terranes in Anadyr-Koryak (Figure 2.1) indicate
that they represent an amalgamation of rocks deposited at different paleolatitudes, spanning a
latitudinal width greater than 3,000 km (Filatova and Vishnevskaya, 1997). Within these terranes,
Middle-Upper Triassic arc volcanics and back-arc and fore-arc basin sediments are separated
from an overlying Middle Jurassic to Upper Cretaceous sequence of similar facies (Filatova and
Vishnevskaya, 1997; Harbert et al., 2003) by a Lower-Middle Jurassic deep-oceanic radiolarian-
bearing sequence without evidence for volcanism (Filatova and Vishnevskaya, 1997). Taxonomic
composition of the radiolarian assemblages indicate that the second-generation island arc complex
was formed ~2,000 km away from the Asian continental margin (Filatova and Vishnevskaya, 1997).
Faunal (Filatova and Vishnevskaya, 1997) and paleomagnetic data (Harbert et al., 2003; Nokleberg
et al., 2000) show that the exotic terranes formed at low latitudes between ~20° and 40°N (Filatova
and Vishnevskaya, 1997; Harbert et al., 2003; Nokleberg et al., 2000) from Triassic to earliest
Cretaceous times, followed by a northward drift culminating Early-Middle Cretaceous accretion

23
Figure 2.2. Tomographic slices. Two cross-sections of the tomographic model (Amaru, 2007)
centred at the 180° meridian of A) vertical slice at 4° N and B) horizontal slice at 1,900km depth.
Fast (blue) P-wave speed anomalies at the centre of each tomography section show the suggested
Triassic-Jurassic slab remnants. Yellow stars represent central Pacific seismic scatterers, caused by
subducted and folded oceanic crust (Kaneshima and Helffrich, 2010). Zones of low-or absent
image resolution occur above the inclined large-dashed line in the lower panel of (A). Numbers
along the horizontal axis in a denote arc-degrees along the section. The horizontal solid black line
in (B) shows the location of section (A).

24
to the east Asian continental margin at high latitudes (Filatova and Vishnevskaya, 1997; Harbert et
al., 2003; Nokleberg et al., 2000).
Japan’s accretionary wedge formed episodically over the past 500 million years (Myr) (Isozaki
et al., 1990) and contains rocks of an intra-oceanic volcanic arc in the Oku-Niikappu region (Ueda
and Miyashita, 2005). The Oku-Niikappu arc is overlain by Upper Jurassic-Lower Cretaceous
(~145-130 Myr old) radiolarites, marking the minimum age of arc extinction, and Middle-
Cretaceous (~105-95 Myr old) foreland basin clastic deposits, marking the arrival at Japan’s
subduction zone. The Oku-Niikappu arc had thus been extinct for at least 40 Myr at the moment
of accretion to Japan, indicating that the intraoceanic subduction zone from which it was derived
was located far offshore east Asia (Ueda and Miyashita, 2005). Paleomagnetic (Oda and Suzuki,
2000; Uno et al., 2011) and biostratigraphic (Matsuoka, 1996) data from Triassic and Jurassic rocks
in the Japanese accretionary prism consistently show derivation from equatorial to low latitudes
(~5°-30°N) (Matsuoka, 1996; Oda and Suzuki, 2000; Uno et al., 2011), indicating that the plate on
which the Oku-Niikappu arc formed had a northward component of motion after extinction. We
thus infer a Middle-Jurassic low northerly paleolatitude for the Oku-Niikappu arc.
Finally, the Stikinia and Wrangellia terranes of western North America form an amalgamation
of Permian-Jurassic intra-oceanic arcs, ophiolites and accretionary wedges that formed somewhere
west of continental North America. Paleomagnetic data constrain the Late-Triassic paleolatitude of
Wrangellia to ~20°N (Kent and Irving, 2010; Nokleberg et al., 2000), south of Stikinia (~30°-40°N;
(Nokleberg et al., 2000)), and a northward drift by oblique and margin-parallel motion relative to
north America in Cretaceous and younger time (Kent and Irving, 2010; Nokleberg et al., 2000).
Together, these exotic terranes document that intra-oceanic subduction zones existed far
offshore from the circum-Pacific continental margins in Triassic and Early Jurassic time, spread
over latitudes from near-equatorial to ~55°N. We now explore plate tectonic reconstructions in
a mantle reference frame and seismic tomographic models to identify where slab-like anomalies
reside that may constrain the paleolongitudes of these subduction zones.
To estimate the position of possible remnants of intra-Panthalassa Ocean subduction relative to
the circum-Panthalassa continents in Triassic-Early Jurassic time, we place our global plate tectonic
reconstruction (Torsvik et al., 2008a) in a mantle reference frame. For the early Mesozoic, there are
no hotspot reference frames (Steinberger and Torsvik, 2008), but we can resort to two alternatives,
a hybrid true-polar-wander-corrected reference frame (Steinberger and Torsvik, 2008; Torsvik et
al., 2008a; Torsvik et al., 2010) or our slab-fitted frame (van der Meer et al., 2010), which is a
modification of the former. The reference frames constrain the longitudes of the early Mesozoic
Panthalassa Ocean (Figure 2.1B), a region largely occupied by the present Pacific Ocean (Figure
2.1A). This provides the required search area at surface, where remnants of intra-Panthalassa
subduction should reside in the mantle.

Tomographic observations below the Pacific Ocean

To predict the depth range at which Triassic-Jurassic slabs reside in the mantle, we can use
estimates of global average rates of slab sinking. Based on previous correlations between
tomographic interpretation of subducted slabs and the geological record (van der Meer et al.,
2010; van der Voo et al., 1999b), the studied lower-mantle slabs indicate sinking rates of ~1 cm/yr.
Mantle-convection modelling (Lithgow-Bertollini and Silver, 1998; Steinberger, 2000) correlated
with global tomographic models obtain the best fit when higher sinking rates (>1.5 cm/yr) are

25
Figure 2.3. Comparison of tomographic models. Tomographic slice A) at 2,300 km depth (Amaru,
2007), showing positive P-wave speed anomalies below the central Pacific, depicted here by black
dashed oval, and B) at 2,400 km depth (Ritsema et al., 2011), showing S-wave speed anomalies
(Ritsema et al., 2011) with zero to positive S-wave speed amplitudes, separating the LLSVP with
strong negative amplitude into a western and eastern region. Black arrow shows the modelled
extent of the western LLSVP high along a northwest-southeast profile (He and Wen, 2009).

used. We assume average slab-sinking rates are at least 1 cm/yr in the mantle, hence early Mesozoic
subducted slabs originating from the Panthalassa Ocean should reside in the lower half of the
Pacific mantle (>1,500 km depth) at present. There, the tomographic P-wave speed model UU-P07
(Amaru, 2007) shows an elongated zone of segmented positive anomalies from 30°S to 60°N below
the central Pacific Ocean, at depths greater than 1,900 km (Figures 2.2 and 2.3). At 50°N this
elongated zone curves northeastwards, towards anomalies that have previously been interpreted
by us as slabs associated with Mesozoic intra-oceanic subduction west of North America (van der
Meer et al., 2010), associated with the Wrangellia and Stikinia terranes (Figure 2.1). Previously,
the presence of seismic scatterers (Figure 2.2) in the deep mantle below the central Pacific was
interpreted to be caused by remnants of subducted and folded former oceanic crust (Kaneshima
and Helffrich, 2010).
The north-south belt of positive anomalies is well corroborated by a recent tomographic S-wave
speed model (S40RTS; Ritsema et al., 2011) that is based on independent data and methodology.
This model shows fairly similar lateral geometry at depths of 1,800-2,400 km (Figure 2.3). Both
models indicate a separation of the top part of the Pacific large low shear wave velocity province
(LLSVP; Garnero and McNamara, 2008) into two parts to a depth of ~2,700 km (Figure 2.3).
Waveform modelling of trans-Pacific S-wave seismic rays (He and Wen, 2009) along a northwest-
southeast profile also corroborates our findings, where at depths greater than ~2,550 km, the

26
geometry of the Pacific LLSVP consists of two modelled highs, separated by a 740-km-wide gap
(He and Wen, 2009). At 2,400 km depth (Figure 2.3B) the modelled southeastern edge of the
western LLSVP high (He and Wen, 2009) terminates at the north-south belt of positive anomalies.
In both mantle reference frames (Torsvik et al., 2010; van der Meer et al., 2010), the early
Mesozoic Laurentian, east Asian and southern Gondwana continental margins are placed above the
corresponding Farallon (Grand et al., 1997; van der Meer et al., 2010), East China (van der Meer
et al., 2010; van der Voo et al., 1999) and Georgia Islands slabs (van der Meer et al., 2010) (Figure
2.4). The anomalies identified in the deep mantle below the Pacific Ocean (Figures 2.2 and 2.3)
do not project below any continental margin in the reconstruction (Figure 2.4), but instead have
an intra-Panthalassa location. This central-Pacific belt of anomalies is therefore interpreted as the
remnant of intra-Panthalassa subduction and constrains the location of the Triassic-Early Jurassic
intra-Panthalassa subduction zones that are recorded in the circum-Pacific exotic terranes.

Reconstruction of subduction in central Panthalassa Ocean

This interpretation is consistent with the paleolatitudes from near-equatorial to ~55°N recorded
in the Japanese and far-east Asian terranes (Filatova and Vishnevskaya, 1997; Harbert et al., 2003;
Nokleberg et al., 2000; Stone et al., 2003; Ueda and Miyashita, 2005). In both tomographic models
(Amaru, 2007; Ritsema et al., 2011) (Figure 2.3), the amplitudes of the anomalies are weaker than
for other slabs identified at equivalent depths associated with circum-Pangaea subduction zones
(van der Meer et al., 2010). This may be explained by the proximity of the slabs to the hotter
LLSVP (Garnero and McNamara, 2008), where enhanced thermal assimilation of the slab may
combine with lower tomographic resolution to reduce seismic velocity anomalies.
The zone of anomalies is not continuous. This may imply strongly segmented plate boundaries
in the Triassic-Jurassic Panthalassa Ocean (Figures 2.2 and 2.4), perhaps owing to their subduction
history (for example, because of the opening of back-arc basins) similar to Mesozoic-Cenozoic
subduction configurations of the northern Tethys Ocean (van der Meer et al., 2010) and the west
Pacific Ocean at present.
Using the paleolatitude constraints and their position with respect to each other, we suggest
that the Wrangellia and Stikinia intra-oceanic terranes were located outboard of western Laurentia,
and are associated with slabs that are found to the west of the Farallon slab (Grand et al., 1997; van
der Meer et al., 2010) (Figure 2.4). We correlate the Kolyma-Omolon, Anadyr-Koryak and Oku-
Niikappu subduction-related exotic terranes to the array of central-Pacific anomalies (Figure 2.4).
To bring these arcs from a central-Pacific origin to the continental margin after their extinction
requires plate speeds not exceeding 10 cm/yr, which is comparable to modern plate velocities.
If we assume that all main positive anomalies represent slab remnants, their overall
configuration (Figure 2.4) implies a major division of the early Mesozoic Panthalassa Ocean into
a western and eastern realm, separated by a north-south trending belt of intra-oceanic subduction
zones. We name the eastern part the Thalassa Ocean(s), hosting the preserved parts of the Izanagi,
Pacific, Phoenix and Farallon plates. The western part we denote Pontus Ocean, of which all its
oceanic lithosphere has been subducted. The dividing array of intra-oceanic subduction zones
we name Telkhinia. We tentatively place the Japanese and Asian exotic terranes on the western
margin of the Thalassa oceanic plate(s). The two major Pontus and Thalassa oceans are surrounded
by smaller peripheral oceans including the Mezcalera Ocean (Dickinson and Lawton, 2001) in the
east, the Slide Mountain Ocean (Nokleberg et al., 2000) in the northeast, the Oimyakon Ocean

27
Figure 2.4. Plate tectonic interpretation of tomography. A) Tomographic depth slice at 2,300km
(Amaru, 2007). Colour scale as in Figure 2.3. Discussed slabs: EC, east China; F, Farallon; GI,
Georgia Islands; Tk, Telkhinia slab group. B) Modified plate tectonic reconstruction (Steinberger
and Torsvik, 2008; Torsvik et al., 2008; van der Meer et al., 2010) 200 Myr ago ( Jurassic-Triassic
boundary). Red lines with triangles denote interpreted subduction zones, polarities are speculative.
Yellow zigzag denotes presumed spreading ridge. Green line denotes presumed transform zone.
White ellipses denote inferred position of the discussed exotic terranes (see also Figure 2.1).

28
(Nokleberg et al., 2000) in the north and Mongol-Okhotsk Ocean (van der Voo et al., 1999) in
the northwest. In the west a transition to the Tethys Ocean seems to occur through a number of
undocumented subduction zones.
Assimilating these constraints in our reconstruction (Steinberger and Torsvik, 2008; Torsvik
et al., 2008; van der Meer et al., 2010) allows for the following, consistent reconstruction. Intra-
oceanic subduction along Telkhinia from the Triassic to Early Jurassic led to the formation of island
arcs, partly preserved at circum-Pacific continental margins (Filatova and Vishnevskaya, 1997;
Nokleberg et al., 2000; Ueda and Miyashita, 2005), and led to slab remnants below the central
Pacific Ocean, identified in tomographic models (Amaru, 2007; Ritsema et al., 2011) from medium
latitudes in the south to high latitudes in the north. Our reconstructed position of the Telkhinia
subduction zones provides a new estimate for the paleolongitude of the associated superterranes
and yields new insights in the geodynamic evolution of the Panthalassa Ocean. This provides first-
order constraints for more detailed plate tectonic reconstructions of oceanic domains of the early
Mesozoic, and allows for the building of more realistic geodynamic models derived from those.

Methods

Geological descriptions and paleomagnetic data of fossil volcanic arcs of the circum-Pacific margins
were tested against criteria (Triassic-Jurassic fossil arc, exotic fauna, accretion time) for a central
Panthalassa Ocean intra-oceanic subduction zone.
Hit counts of the tomographic model (Amaru, 2007) show that in our area of interest
(30°S-60°N), the mantle is sampled with hit counts varying from ~500 raypaths per cell to
~5,000 raypaths per cell. Spike tests as indicators of spatial resolution quality, show that anomaly
sign patterns on scales of 500-1,000 km are detectable in the lower mantle below the central
Pacific mantle where the cell hit count is >500 (outside the low-resolution zones). The spike tests
indicate that anomalies of 500-1,000 km are detectable, but their amplitudes are underestimated.
We selected only imaged anomalies in the UU-P07 tomographic model (Amaru, 2007), with
peak amplitudes above +0.2%, consistent with our previous efforts of global slab identification
for constraining the longitudes of Pangaea in an absolute mantle reference frame (van der
Meer et al., 2010). Paleosubduction zones were added based on the tomographic interpretation
(Steinberger and Torsvik, 2008; Torsvik et al., 2008; van der Meer et al., 2010) to the plate tectonic
reconstruction (Figure 2.4). The error in paleosubduction location was estimated to be ±500 km
for the anomalies in the lower mantle (van der Meer et al., 2010). We tentatively correlate slab
windows between subduction-zone segments to transfer zones. These remain speculative, as the
apparent slab windows could also be the result of ridge subduction or tomographic imaging.
Nomenclature of the Panthalassa Ocean subdivision (Figure 2.4) has been based on Greek
mythology for consistency with the Greek word Panthalassa (‘all sea’). The offspring of the primeval
gods Pontus and Thalassa were the Telkhines, who were later cast by Zeus to the underworld.

29
30
Chapter 3

Plate tectonic controls on atmospheric CO2 levels


since the Triassic
This chapter has been published as: van der Meer, D.G., Zeebe, R., van Hinsbergen, D.J.J., Sluijs, A., Spakman, W.
and Torsvik, T.H., 2014, Plate tectonic controls on atmospheric CO2 levels since the Triassic, Proceedings of the
National Academy of Sciences of the United States of America (PNAS), 111, p. 4380-4385

Climate trends on time-scales of 10’s to 100’s of millions of years are controlled by changes in solar
luminosity, continent distribution and atmosphere composition. Plate tectonics affect geography,
but also atmosphere composition through volcanic degassing of CO2 at subduction zones and
mid-ocean ridges. So far, such degassing estimates were based on reconstructions of ocean floor
production for the last 150 million years (Myr) and indirectly, through sea level inversion before
150 Myr. Here we quantitatively estimate CO2 degassing by reconstructing lithosphere subduction
evolution, using recent advances in combining global plate reconstructions and present-day
structure of the mantle. First, we estimate that since the Triassic (250-200 Myr) until Present, the
total paleo-subduction zone length reached up to ~200% of the present-day value. Comparing our
subduction zone lengths with previously reconstructed ocean-crust production rates over the past
140 Myr suggests average global subduction rates have been constant, ~6 cm/year: Higher ocean-
crust production is associated with longer total subduction length. We compute a Strontium isotope
record based on subduction zone length, which agrees well with geological records supporting the
validity of our approach: The total subduction zone length is proportional to the summed arc-
and ridge volcanic CO2 production and thereby to global volcanic degassing at plate boundaries.
We therefore use our degassing curve as input for the GEOCARBSULF model to estimate
atmospheric CO2 levels since the Triassic. Our calculated CO2 levels for the mid-Mesozoic differ
from previous modeling results and are more consistent with available proxy data.

Introduction

Volcanism forms the most efficient mechanism to transfer carbon (C) from the mantle to the
ocean-atmosphere system, the exogenic reservoir. At Mid-Ocean Ridges (MORs), mantle
upwelling occurs between two diverging plates. At subduction zones, sinking oceanic plates lead to
arc volcanism (Stern, 2002). A third type of volcanism, driven by mantle plumes, leads occasionally
to large igneous provinces (LIPs) that typically last a few Myr (Marty and Tolstikhin, 1998; Svensen
et al., 2007). Long-term (>>5 Myr) changes in volcanic output thus dominantly relate to the long-
term plate tectonic processes. Constraints on volcanism-related degassing of CO2 at subduction
zones and mid-ocean ridges, hence derives from studies on lithosphere production/consumption
(Kerrick, 2001; Marty and Tolstikhin, 1998; Resing et al., 2004; Sano and Williams, 1996). The
quality of these constraints determines the mechanistic underpinning of climatic changes on time
scales of 107 to 108 years. This is because on such time scales, volcanism is the major source of
CO2 to the atmosphere and thereby drives climate and the critical feedbacks in the carbon cycle,

31
Figure 3.1. Plate tectonic interpretations of tomographic model. A) Present. B) 170 Myr ago. Red
lines with triangles denote interpreted subduction zones (van der Meer et al., 2010; 2012) from
which total subduction zone length is estimated, polarities are speculative. Green lines denote
presumed transform zone. Yellow triangles denote present sub-aerial volcanoes, largely coinciding
with subduction zones. Continental blocks in slab-fitted absolute plate reference frame (van der
Meer et al., 2010).

such as weathering. To date, estimates of CO2 degassing through plate tectonic processes for the
last 150 Myr are based on reconstructions of ocean floor production (Engebretson et al., 1992),
while for older periods indirect inferences were made through sea level inversion (Gaffin, 1987). An
increasing body of work reveals major uncertainties in eustatic sea level reconstructions, however,
based on regional data (Müller et al., 2008b; Sluijs et al., 2008; Stocchi et al., 2013).
Over the past two decades, advances have been made in correlation of global plate tectonic
reconstructions with mantle structure (e.g. Lithgow-Bertelloni and Richards, 1998; Richards and
Engebretson, 1992; Steinberger, 2000). Recent work suggest that remnants of the last ~250 Myr
of subduction can still be imaged in seismic tomographic models (van der Meer et al., 2010; 2012).
We apply this to estimate the total imaged subducted lithosphere ‘slab’ length at depth, and from
that, paleo-subduction zone length at the surface for the past 250 Myr. Subsequently we compare
this to ocean crust production rates to assess the associated average subduction rates. In this way,
we may obtain a quantitative estimate of lithosphere consumption, and through that, production

32
since the Triassic. To validate our estimates, we compute a global Strontium isotope record based
on subduction zone length and test that against the geological record. Finally, we use the thus
obtained lithosphere consumption/production curve as input for the widely used, long-term carbon
cycle model GEOCARBSULF (Berner, 2006) to evaluate the difference in model output with our
new degassing parameter for estimating atmospheric CO2 levels and test these against previous
modeling results and available proxy data (Park and Royer, 2011).

Subduction zone length evolution

Recently, major elongate P-wave seismic velocity anomalies in the mantle were correlated with
locations of former subduction zones (van der Meer et al., 2010) in plate tectonic reconstructions
using a true polar wander corrected paleomagnetic reference frame as basis (Steinberger and
Torsvik, 2008; Torsvik et al., 2008). In a combined analysis of these wave speed anomalies (using
the UU-P07 tomographic model (Amaru, 2007), corroborated by S-wave velocity anomalies in the
independent model S40RTS (Ritsema et al., 2011), and geological records of mountain building
of the past ~250 Myr, the position and timing of past subduction episodes was determined
for 28 major slabs mostly along the continental margins of Pangea (van der Meer et al., 2010)
These analyses implied a global average sinking velocity of slabs in the lower mantle of 1.2 ± 0.3
cm/yr, yielding a ‘mantle memory’ of subduction of some 250 Myr (van der Meer et al., 2010).
Additionally, major intra-oceanic subduction within the Panthalassa oceanic realm surrounding
Pangea was identified from the deep mantle structure below, and the geological records around
the modern Pacific Ocean (van der Meer et al., 2012). These analyses combined resulted in a first-
order interpretation of global subduction history and show that all major elongated positive seismic
velocity anomalies in the mantle can be consistently matched with former subduction zones, and
vice versa. We use this logic to estimate the global subduction zone length evolution from seismic
tomography. Assuming a constant 1.2 cm/yr slab sinking rate, the global configuration of slabs
in a depth section represents the paleo-subduction zone configuration of an approximate age. To
calculate the paleo-subduction zone length from the lateral slab length measured in a depth section,
we compensate the length estimates for volume reduction as a result of mantle phase changes and
bulk compression. Our interpretations of the top 200 km of the tomographic model, representing
late Cenozoic slabs, yield a total subduction zone length of ~42,400 km ± 1,300 km (Figure 3.1A),
comparing well to the total modern trench length (43,400 km ( Johnston et al., 2011) to 48,800
km (Schellart et al., 2007)), see Table 3.1. The calculated total subduction zone length increases to
a maximum of ~75,000 km at ~2000 km depth (~170 Myr) (Figure 3.1B) and decreases to ~65,500
km at 2815 km depth (~235 Myr) (Figure 3.2A). Given the error bar of ±0.3 cm/yr on the 1.2
cm/yr slab sinking rate (van der Meer et al., 2012) the subduction zone length curve derived here
represents an inherently smoothed version of the past. We expect that future studies of individual
slabs and tomographic models will be able to estimate the subduction zone length with ever
increasing accuracy.
The resulting subduction zone length through time can be described as a simple parabolic
function. Increasing total subduction zone length since the Triassic relates to break-up of Pangea,
when continuous subduction along the circum-Pacific continental margins was established (Grand
et al., 1997; van der Meer et al., 2010; Van der Voo et al., 1999) and intra-oceanic subduction zones
in the Panthalassa Oceans formed (Nokleberg et al., 2000; Sigloch and Mihalynuk, 2013; van
der Meer et al., 2012). Since the mid-Jurassic, the total subduction zone length decreased during

33
Figure 3.2. Subduction through time A) Calculated total subduction zone length (blue triangles)
with error bars. B) Production rates by mid-oceanic-ridges according to Coltice et al.(2013) (red),
Rowley (2002) (brown). Standard GEOCARBSULF production rate (purple line) of Berner
(2006). Obtained slab flux (blue triangles) converted to geologic time with 1.2 cm/year slab sinking
rate (van der Meer et al., 2010) and polynomials assuming subduction rates of 6 cm/yr (blue line,
best-fit,), 5 cm/yr (orange line), 7 cm/yr (green line).

34
collisions in the Tethys Ocean (Stampfli, 2003; Torsvik et al., 2010; van der Meer et al., 2010), the
Mongol-Okhotsk Ocean (van der Meer et al., 2010; Van der Voo et al., 1999) and demise of intra-
oceanic subduction zones in the Panthalassa Ocean (van der Meer et al., 2012), concomitant with
growth of the Pacific Plate (Coltice et al., 2013; Seton et al., 2012).

Correlation with Mid-Ocean-Ridge Crustal Production

We aim to derive quantitative estimates for plate tectonic volcanic degassing at mid-ocean ridges
and arcs combined. Total area consumption at subduction zones per unit time, defined as ‘slab flux’
(km2/yr) (Silver and Behn, 2008) is equal to total area production at oceanic ridges. To test how
subduction zone length (km) through time is reconciled by area production (km2/yr) through time,
we need to multiply subduction zone length with a global average subduction rate (cm/yr).
Marine-magnetic anomaly reconstructions have led to two schools of thought for area
production rates since the Cretaceous (Figure 3.2B). One school interpreted production rates to
have not varied significantly (Cogné and Humler, 2004; Rowley, 2002), the second interprets a
decreasing rate (Coltice et al., 2013; Gaffin, 1987; Richards and Engebretson, 1992; Seton et al.,
2012). The latter scenario served as input for the GEOCARBSULF model for atmospheric CO2
evolution (Berner, 2006; Park and Royer, 2011).
Matching our subduction zone length curve to the constant area production rate scenario
(Cogné and Humler, 2004; Rowley, 2002) requires that the global average subduction rate increased
during the Cenozoic (65-0 Myr) from ~3-4 cm/yr to 5.5 cm/yr (present-day, Schellart et al., 2007).
Matching the decreasing rate scenario (Coltice et al., 2013), however, requires a constant average
subduction rate of 6±1 cm/yr (Figure 3.2B), i.e. similar to the present-day (Schellart et al., 2007).
We adopt here a more or less constant global average of subduction rates as we see no geodynamic
cause why the process of subduction and the mechanical interaction between slab and mantle
should have significantly changed over the past few hundred million years. This may be further
illustrated by the observation that the global average of slab sinking rates appear to be more or less
constant since the early Mesozoic (Butterworth et al., 2014; van der Meer et al., 2010).
Prior to 140 Myr, global MOR spreading evolution is largely unknown because pre-Cretaceous
oceanic crust has almost entirely been subducted (Seton et al., 2012). We expressed this as “World
Uncertainty”, the fraction of the Earth’s lithosphere which has been subducted since a given time
(Torsvik et al., 2010), and for which plate motion and configuration at that time is essentially
unknown. World Uncertainty is ~60% at 140 Myr, equivalent to 85% uncertainty of the oceanic
part of the crust. Global plate models of Mesozoic oceans (Seton et al., 2012), and calculations
based on these, such as for seawater chemistry (Müller et al., 2013) are based on extrapolation.
For pre-Cretaceous time, we therefore prefer to use our total subduction zone length curve instead,
since it is based on (indirect) observation, mantle structure.
In Figure 3.2B we show, aside from the estimates for crustal production (Müller et al., 2008a)
and consumption (this paper), also the back-calculated crustal production. This is the standard
input (Berner, 1994; Engebretson et al., 1992; Gaffin, 1987) for the widely used, long-term carbon
cycle model GEOCARBSULF (Berner, 2006; Park and Royer, 2011). The standard input into
GEOCARBSULF curve significantly underestimates the crustal production prior to 120 Myr.
Applying our new curve to GEOCARBSULF will thus likely generate major changes in modeled
CO2 budgets of the pre-120 Myr world. To test the robustness of our estimates, we will therefore
test the predictions of our model for the Strontium record.

35
Figure 3.3. Corroboration by the Strontium record Green: measured 87Sr/86Sr (Prokoph et al.,
2008), blue line and triangles: calculated 87Sr/86Sr curve by using our subduction zone length curve
and mixing model (Allègre et al., 2010).

Correlation with Strontium isotope record

Slab flux (km2/yr) has previously been assumed to be proportional to global arc magmatic
production rates (km3/yr) (Silver and Behn, 2008), an assumption that may be further supported
by our finding that the average subduction rate through time is approximately constant. Our
subduction zone length through time curve (Figure 3.2A) is a proxy for slab flux and may thus
constitute a proxy for bulk plate-tectonic geochemical input to the exogenic pool since the Triassic.
We test this by comparing geologically well-documented Sr isotope ratios (Prokoph et al., 2008)
with values predicted from our subduction zone length curve. We use a recent mixing-model
(Allègre et al., 2010) describing how the 87Sr/86Sr value of the worlds’ oceans results from mixing
inputs of rivers (with higher 87Sr/86Sr) and mantle input from ridge spreading and (weathering)
island arcs (with lower 87Sr/86Sr). Scaling the mantle input with our subduction zone length curve,
we predict a 87Sr/86Sr curve (Figure 3.3A) that correlates well with the 87Sr/86Sr curve measured
from the geological record (Prokoph et al., 2008).
This positive test supports our hypothesis that subduction zone length is a reliable first-order
proxy for plate tectonics-induced long-term geochemical changes of the global exogenic pool
since the Triassic. Our computed curve cannot reproduce second-order variations in the measured
87
Sr/86Sr record, which highlights the limitation of our method to variations in fluvial input i.e. as
the result of changing paleo-geography (Gibbs et al., 1999; Goddéris et al., 2014; Otto Bliesner,
1995) and to timescales shorter than some 10’s of Myr.

36
Plate tectonic control on CO2 production

We now explore the use of our subduction zone length curve to develop the first fully mechanistic
proxy of volcanic CO2 degassing over the Mesozoic and Cenozoic. Upon subduction, most CO2
is released at depths shallower than 100 km (Gorman et al., 2006), after which it is released to the
surface within a few hundred years ( John et al., 2012). About 40-70% (Dasgupta and Hirschmann,
2010) of the subducted C returns to the exogenic reservoir via volcanism. This occurs within ~2
Myr after subduction, which is instantaneous on the analyzed time scale (> 10 Myr). The remainder
could be introduced to the deep mantle and could be returned to the exogenic reservoir over longer
timescales. However, the total carbon budget of the mantle and present-day carbon degassing rates
imply a residence time of C in the mantle of >1 Gyr and perhaps as long as 4.6 Gyr (Dasgupta
and Hirschmann, 2010). This suggests that once carbon enters the deep mantle, it will unlikely be
released as CO2 as a direct result of subduction rate variation over the past 250 Myr (the timescale
of our analyses), but will slowly return to the exogenic reservoir, e.g. via partial melting at ridges
(Dasgupta and Hirschmann, 2006).
The main uncertainty in computing arc CO2 degassing volumes is carbon concentration
of subducting slabs and decarbonation efficiency at subduction zones over time ( Johnston et al.,
2011). Slab carbon concentrations likely varied through biological evolution and the dynamics
of climate and the exogenic carbon cycle, including shelf versus deep ocean carbonate burial,
weathering rates and ocean carbonate chemistry but the net effect on Phanerozoic slab composition
is however poorly known. Additional constraints on carbonate production and preservation at
the ocean floor through time would refine our flux estimates. For now we here assume constant
carbon concentration of subducting lithosphere. Given the constant average global subduction
rates, we also assume that decarbonation efficiency, mainly determined by the thermal structure of
subduction zones ( Johnston et al., 2011), is globally constant.
The carbon degassing from MOR’s has been estimated to be of similar magnitude as arc
volcanism (Dasgupta and Hirschmann, 2006; Marty and Tolstikhin, 1998; Sano and Williams,
1996), although MOR’s may be no net contributor as all CO2 may be stored locally in hydrothermal
carbonates (Kerrick, 2001). Crustal CO2 degassing by ridge volcanism occurs by partial melting of
the upper mantle up to depths of 330 km (Dasgupta and Hirschmann, 2006). As the extent of
melted volume depends on the ridge spreading rate (Dasgupta and Hirschmann, 2006), at a global
level, this translates to overall crustal area production, which as demonstrated above is directly
coupled to the here derived subduction length multiplied with the (constant) subduction rate.
We use the long-term carbon cycle model GEOCARBSULF (Berner, 2006; Park and Royer,
2011) to evaluate the effect of our subduction zone length curve on carbon cycling and atmospheric
CO2 concentrations since the Triassic. In brief, the partial pressure of CO2 in the atmosphere
maintains a long-term flux balance between input and output into and from the exogenic carbon
reservoir through the effect of CO2 on silicate weathering. For instance, an increase in the steady-
state degassing rate requires increased weathering and hence higher steady-state pCO2 to balance
the fluxes (everything else being equal).
In GEOCARBSULF, the total modern degassing rate due to plate tectonic processes
(volcanism, metamorphism, and diagenesis from breakdown of carbonates and organic carbon)
is 7.92*1018 mol CO2/Myr. Given the constant subduction rate through time, we calculate the
total plate tectonic CO2 degassing through time by scaling the modern degassing rate with our
subduction zone length curve (Figure 3.4A). This results in a prediction that CO2 degassing rose

37
Figure 3.4. Subduction impacting CO2 levels A) CO2 degassing parameter, used as input for
GEOCARBSULF (Berner, 2006), Dark blue: standard GEOCARBSULF degassing parameter
(Berner, 2006), Light blue; our subduction zone length curve. Dashed lines; error bounds
B) Atmospheric CO2 levels, as result of Dark blue: NV=0.013 in (Berner, 2006). Light blue: Our
study with NV=0. Grey envelope: proxy data (Park and Royer, 2011) 10 Myr bin interval.

38
from the Triassic to the mid-Jurassic, with a peak CO2 degassing at ~160 Myr, followed by a
gradual decline to the Present (Figure 3.4B).
For the last 120 Myr, our degassing curve is similar to that of previous work based on marine
magnetic anomaly reconstructions (Engebretson et al., 1992). For older times, however, where the
previous curves used sea level inversion (Gaffin, 1987), it is distinctly different. We assume that
some of the discrepancy can be explained by inaccuracies in the sea level reconstruction (Gaffin,
1987).
To test the predictive power of both curves, we ran two different configurations of the
GEOCARBSULF model, one with standard degassing as described above, the other with our
newly obtained degassing parameter (Figure 3.4B). In addition, we optimized a tuning parameter
of the model (NV) to obtain the best match between model-derived pCO2 and proxy records
(Park and Royer, 2011). The tuning parameter NV affects the average value for 87Sr/86Sr of non-
volcanic silicate rocks and ultimately the fraction of Ca-Mg silicate weathering of volcanic rocks
relative to non-volcanic rocks and may be varied within certain limits (see Berner, 2006). While
both degassing parameterizations yield similar pCO2 values from ~120 Myr to the Present, our
degassing parameterization predicts a gradually changing pCO2 level curve since the Triassic
instead of a major increase during the Late Jurassic. Our curve appears to be in better agreement
with proxy reconstructions (Park and Royer, 2011) (Figure 3.4B). For example, a mismatch between
the standard GEOCARBSULF results and pCO2 proxies for the mid-Mesozoic was noted earlier
by Park & Royer (Park and Royer, 2011) and it was shown that a decrease in weatherable land area
in the Jurassic by 50%, would improve the data fit. However, these authors concluded that a 50%
reduction in weatherable land area is unreasonably large for this period of time. In contrast, our
degassing approach as presented here reconciles model results and data during the mid-Mesozoic
without additional assumptions. We note, however, that substantial uncertainties are associated with
both the proxy records and the carbon cycle modeling (Berner, 2006; Park and Royer, 2011; Sano
and Williams, 1996), leading to large discrepancies between models and data during several time
periods. Notably, a major early Cenozoic (50-30 Myr) peak in the CO2 curve cannot be explained
by plate-tectonic processes of either long-term nature: changes in slab flux occur on longer periods,
and the effects of LIPS are of shorter term influence. Rather, this peak may relate to effects of
orogenesis, e.g., in the Neotethyan or Andean realm. This could imply that CO2 variations at that
time were primarily caused by feedbacks in either continental organic carbon burial (Kurtz et al.,
2003) or the methane cycle (Dickens, 2011), for which carbon and sulfur isotopic evidence exists.
Nevertheless, the root mean square error between model-predicted CO2 and proxy CO2 values are
~380 ppmv for the standard degassing parameter and ~250 ppmv for our new degassing parameter,
confirming the better match between model-derived pCO2 and proxy records for the degassing
curve derived here.

Conclusion

In this paper, we quantify global subduction zone length through time since the Triassic using
seismic tomographic images of deep mantle relics of subduction zones. Our curve correlates well
with ocean production rates if the average global subduction zone rate was constant, at 6±1 cm/
yr, which is similar to the modern average subduction rate. We use our results to predict an ocean
strontium isotope curve. This curve tests positively against data, showing that our subduction length
curve provides a first-order estimate of plate tectonics-related mantle degassing since the Triassic

39
for time scales of 10-20 Myr. By applying our subduction length curve to the GEOCARBSULF
(Berner, 2006) climate model, we obtain an unprecedented fit between predicted and reconstructed
atmospheric CO2 levels than previous studies, particularly for pre-Cretaceous time. At present,
global correlation of seismic tomography models to geological evidence of past subduction is still
in its early stages. Increasing resolution in future tomographic models, in combination with more
detailed plate reconstructions will constrain the evolution of subduction zones and their individual
volcanic degassing outputs more accurately. Moreover, improved constraints on subducting slab
carbon content will refine degassing estimates. We aspire that unlocking the deep mantle record
as a novel data source to quantify global volcanic degassing will further our understanding of the
fundamental plate tectonic controls on Earth’s climate system.

Materials & Methods

Lower mantle slab remnants were previously (van der Meer et al., 2010; 2012) inferred from a
tomographic model (Amaru, 2007), where slab extents were determined based on seismic wave-
speed anomaly patterns and where a global paleogeographic context was provided through the
reconstructed position of continents in the geologic past. The total lateral slab length was estimated
from the tomographic model (Amaru, 2007) as a function of depth. These estimates are derived
by the comprehensive analysis of one particular P-wave tomographic model (Amaru, 2007).
Other global tomographic models, based on different data and methods (Houser et al., 2008; Li
et al., 2008; Ritsema et al., 2011; Simmons et al., 2012), are available which generally agree on
the presence of large scale features, but considerable differences exist on smaller scale. We used
conservative errors of 300-400 km at both endpoints of each interpreted slab anomaly in the lower
mantle. Higher degree features are on the scale of this estimated spatial error. Most important issue
is the actual existence of the paleo-subduction zones in the geological record and the associated
slabs in the mantle (van der Meer et al., 2010; 2012). To test the presence of slabs in a different
tomographic model, we compared the mantle anomalies we interpreted in model UU-P07 with the
completely independent (in data, model parameterization, and inversion approach) S-wave seismic
velocity model S40RTS (Ritsema et al., 2011). As this is a first attempt of quantifying subduction
zone length from tomographic models, we hope that other groups will follow our approach and
analyze their tomographic models in similar ways in order to provide more estimates of subduction
zone length through time.
The errors in the paleo-subduction zone length estimates are primarily dependent on
tomographic resolution. Based on previous analyses (van der Meer et al., 2010; 2012), our the
tomographic model has average spatial uncertainties of ~100-200 km in the upper mantle and
~300-400 km in the lower mantle, both in relatively well imaged regions. In summing slab lengths
this propagates into errors of 3-6% of total subduction zone length in the well-imaged parts of
the mantle. However, a more important uncertainty on our estimates of global subduction zone
length stems from parts of the mantle that are insufficiently imaged by seismic tomography. In
part, this uncertainty is reduced by knowledge of absence of subduction in the geological record.
In plate tectonic reconstructions, continents and oceanic crust without any geological evidence for
subduction cover part of the mantle area without sufficient tomographic resolution. The remaining
poorly imaged areas, notably sectors in the former Tethys and Panthalassa oceanic realms, now
mostly below the Indian and Pacific Oceans, may have had intra-oceanic subduction zones that
remain undetected in our tomographic model. These areas are divided by the ratio of the present

40
subduction zone length (~43000 km) over Earth’s ocean surface area (~360 million km2) in order to
arrive at a reasonable estimate maximum error bounds for potentially undetected lengths of paleo-
subduction zones. The combined errors lead to a potential underestimation of global subduction
zone length of up to 17% and may increase the amplitude of the global subduction zone length
curve, but will not change its overall shape.
Because slab material will shrink in volume (3D) with increasing depth as a result of self-
compression and phase changes in the mantle, a correction is needed to obtain the initial
subduction zone length at surface. This (1D) correction is computed from mantle density increase
with depth of the ak135 Earth model (Kennett et al., 1995).
Thickening of slab by pure shear deformation (i.e. without volume change) by factors between
2 and 3 when entering the lower mantle will primarily (Christensen, 1996; Gaherty and Hager,
1994; Hafkenscheid et al., 2006) affect the thickness of the anomaly perpendicular to the strike
of long subduction zones and does not play a significant role in our estimates of along strike
paleo-subduction length. Cooling of the ambient mantle by thermal diffusion may possibly lead
to a broader appearance of lower mantle slabs in tomography images. For thickened slab this is
largely dependent on mantle residence time as estimated from thermal modeling of the cooling of
instantaneously placed (thickened) slab (Hafkenscheid et al., 2006). The first order effect on our
total length estimates would be that we overestimate by ~6% at 100 Myr and ~10% at 200 Myr,
basically increasing the curvature of our curve in Figure 3.2B. The thermal modeling, however, leads
to upper estimates (Hafkenscheid et al., 2006) and does not account for the counter-acting effects
during slab sinking of local advection and heating by the dynamic shear flow around the slab. We
do not expect thermal diffusion to be critically affecting our paleo-subduction length estimates.
Atmospheric CO2 concentrations were calculated using the long-term carbon cycle model
GEOCARBSULF, including weathering of volcanic rocks (Berner, 2006; Park and Royer, 2011). A
FORTRAN version of the code is available at http://earth.geology.yale.edu/~jjpark/Code/gcsv10_
export.f. For all model runs, the following parameter values were used: FERT = 0.4, ACT = 0.09,
LIFE = 0.25, GYM = 0.875 (see Berner, 2006; Park and Royer, 2011; Sano and Williams, 1996
for details). However, for the two different degassing parameterizations, the model parameter NV
was varied. NV affects the average 87Sr/86Sr of non-volcanic silicate rocks (Rnv) and ultimately the
fraction of Ca-Mg silicate weathering of volcanic rocks (Xvolc) relative to non-volcanic rocks via the
relationship (Berner, 2006):

Rnv = 0.717-NV (1 – fR (t)) (3.1)

where fR (t) represents physical erosion over time; Xvolc is a function of Rnv (see (Berner, 2006)).
NV is an arbitrary parameter which may be varied between reasonable bounds of 0.0 and 0.015 and
was optimized here to obtain the best fit between model-derived pCO2 values and proxy records
(Park and Royer, 2011) (Figure 3.4B). We found that NV = 0.013 and NV = 0.0 yields the best fit
to CO2 proxy data for the standard and our new degassing parameter, respectively (smallest root
mean square error). A value of NV = 0 could suggest that physical erosion has little effect on the
volcanic rock weathering fraction. However, we note that NV = 0 only applies to our specific setup
(in addition to NV, other parameters could be varied). Moreover, our model runs are restricted
to the past 235 Myr, while the full scope of the GEOCARBSULF model includes the past 570
Myr. The CO2 proxy data (Park and Royer, 2011) (average, low, and high) were binned in 10 Myr
intervals. The dotted gray line in Figure 3.4B represents average values, while the upper and lower
bounds of the light gray envelope indicate low and high values.

41
Table 3.1 Subduction-zone length data

42
Chapter 4

Reconstructing first-order changes in sea Level


during the Phanerozoic and Neoproterozoic
using strontium isotopes
This chapter has been published as: van der Meer, D.G., van den Berg van Saparoea, A.P., van Hinsbergen, D.J.J., van
de Weg, R.M.B., Godderis, Y., Le Hir, G. and Donnadieu, Y., Reconstructing first-order changes in sea level
during the Phanerozoic and Neoproterozoic using strontium Isotopes, Gondwana Research, v44, pp 22-34, 2017
http://dx.doi.org/10.1016/j.gr.2016.11.002

The eustatic sea-level curves published in the seventies and eighties have supported scientific
advances in the Earth Sciences and the emergence of sequence-stratigraphy as an important
hydrocarbon exploration tool. However, validity of reconstructions of eustatic sea level based on
sequence stratigraphic correlations has remained controversial. Proposed sea level curves differ
because of site-to-site changes in local tectonics, depositional rates, and long-wavelength dynamic
topography resulting from mantle convection. In particular, the overall amplitude of global
Phanerozoic long-term sea level is poorly constrained and has been estimated to vary between ~400
m above present-day sea level to ~50 m below present-day sea level. To improve estimates of past
sea level, we explore an alternative methodology to estimate global sea level change. We utilise the
Phanerozoic-Neoproterozoic 87Sr/86Sr record, which at first order represents the mix of inputs from
continental weathering and from mantle input by volcanism. By compensating for weathering with
estimates of runoff from a 3D climate model (GEOCLIMtec), a corrected 87Sr/86Sr record can be
obtained that solely reflects the contribution of strontium from mantle sources. At first order, the
flux of strontium from the mantle through time is due to increases and decreases in the production
of oceanic crust through time. Therefore, the changing levels of mantle-derived strontium can be
used as a proxy for the production of oceanic lithosphere. By applying linear oceanic plate age
distributions, we compute sea level and continental flooded area curves. We find that our curves
are generally within the range of previous curves built on classical approaches. A Phanerozoic
first order cyclicity of ~250 Myr is observed that may extend into the Neoproterozoic. The low
frequency (i.e., on the order of 10 to 100 Myr) sea level curve that we propose, while open for
improvement, may be used as baseline for refined sequence-stratigraphic studies at a global and
basin scale.

Introduction

Knowledge of past sea-level fluctuations is of fundamental importance to many disciplines in the


geosciences, particularly sequence stratigraphy, which is widely used in petroleum exploration.
Sequence stratigraphy plays a central role in modern sedimentology. It can be regarded as the
third and most recent major revolution in sedimentary geology, following the incorporation of
process sedimentology and plate tectonics into sedimentological research (Miall, 1995). Sequence

43
Figure 4.1. A) Sea level curves from stratigraphic methods (Vail et al., 1977; Hallam, 1984; Algeo and Seslavinsky, 1995; Snedden and Liu, 2010)
and plate tectonic methods (Müller et al., 2008; Verard et al., 2015) B) Continental flooded area curves (Ronov, 1994; Walker et al., 2002; Smith
et al., 2004) and constructed on the basis of the paleogeographic maps of Blakey (Blakey, 2003, 2008), Golonka (Golonka, 2007a, b, 2009a, b,
2012) and Scotese (Scotese, 2014a, b, c, d, e, f, 2016).

44
stratigraphy combines several disciplines into a unified model to explain the development and
stratigraphic architecture of basin fills. The publication of AAPG Memoir 26 (Payton, 1977) is
commonly regarded as the birth of sequence stratigraphy. Since then the concept has developed
into a widely-used method for interpreting the geological record and predicting stratigraphic
architecture, and has proven to be very successful in its application to hydrocarbon exploration. One
way in which sequence stratigraphy has been applied is the reconstruction of past eustatic sea-level
fluctuations. Probably one of the best known sea-level curves was published by Vail et al. (1977).
Their lowest frequency curve related to plate tectonics is shown in Figure 4.1A. However, as will be
discussed below, using sequence stratigraphy for reconstructing eustatic sea level does not seem to
be a valid approach.
As a result, global, long-term sea-level changes remain poorly constrained. Most recently,
incorporating the long-term trends of stratigraphically derived curves (Hardenbol et al., 1998; Haq
and Schutter, 2008), sea-level fluctuations have been estimated to vary during the Phanerozoic from
as much as ~400 m above present-day sea level, to ~50 m below present-day sea level (Snedden and
Liu, 2010, 2011). However, Snedden and Liu (2010, 2011) are careful in interpreting long-term
curves, indicating that there is little consensus on the range of sea level changes, and that most
authors believe that range of sea level during most of the Phanerozoic was within ±100 meters of
the present-day level (Snedden and Liu, 2010).
Alternative methods to reconstruct first-order or lowest frequency (on the order of 10-100
Myr) sea level have used plate tectonic modelling. Changes in eustatic sea level result from
processes that change either the volume of seawater in the oceans or the volume of the ocean
basins (Worsley et al., 1984; Conrad, 2013). Glacio-eustasy, reflecting ice-volume changes on a
timescale of 10 kyr to 1 Myr, is one way to change the volume of seawater (Miller et al., 2005).
Ocean spreading rates dictate the average age of the ocean floor and therefore the average depth
of the ocean floor (Parsons and Sclater, 1977; Stein and Stein, 1992). Conrad (2013) studied six
long-term factors that have modified sea level during the past 140 Myr and concluded that volume
of spreading ridges is the dominant driver of eustatic changes (~250 m). Other factors considered
include plume-related seafloor volcanism (~140 m), dynamic topography (~70 m), climate (~60
m), sediment thickness (~60m) and the area of the oceans (~ 10 m). Plate tectonic reconstructions
based on seafloor isochrons have been used to estimate the changing age distribution of ocean
floor during the Phanerozoic and the resulting change in sea level (Worsley et al., 1984; Verard
et al., 2015). Using mantle tomography and mantle convection models, Müller et al. (2008) and
Spasojevic and Gurnis (2012) incorporated dynamic topographic effects and estimated global sea
level back into the Cretaceous. However, estimates derived from plate reconstruction are hampered
by diminishing record back in time of preserved oceanic crust which leads to large uncertainties
in the age distribution of the ocean floor (Rowley, 2002; Müller et al., 2016) Reconstructions of
times prior to 180 million years (Early Jurassic) suffer from having approximately 70% of the Earth
surface unconstrained (Torsvik et al., 2010; Müller et al., 2016).
In this paper we explore an alternative method of reconstructing first-order sea level, using
the changing values of 87Sr/86Sr isotope ratios recorded in the marine fossil record (Spooner, 1976;
Hallam, 1984; Worsley et al., 1985). Spooner (1976) demonstrated a positive correlation between
increasing land area during the last 70 Myr and an increase in the 87Sr/86Sr isotope ratio. However,
in a review of 87Sr/86Sr isotope ratio values for the entire Phanerozoic (Hallam, 1984) indicates that
this simple correlation between land area and higher 87Sr/86Sr ratios does not hold.
Worsley et al. (1984) created a Phanerozoic sea level model based on variations in the volume of
the world ocean basin caused by changes in ocean area and depth, which produced eustatic sea

45
Figure 4.2. A) 87Sr/86Sr record of (McArthur et al., 2012; Cox et al., 2016), compared with the curves on the basis of crustal production (Verard
et al., 2015) and subduction zone lengths (van der Meer et al., 2014). B) Strontium Flux ratio relative to present. Weathering component is
output from GEOCLIMtec (Goddéris et al., 2014). Mantle component is derived by correcting the 87Sr/86Sr record (McArthur et al., 2012; Cox
et al., 2016) with the weathering component.

46
level changes of several 100 meters. Their model matched low frequency (on the order of 100 Myr)
cycles of eustatic sea level change (Vail et al., 1977; Pitman, 1978). The Worsley et al. (1985) model
also correlated well with the changing 87Sr/86Sr isotope ratio.
Recently, plate tectonic models have become better at qualitatively matching the 87Sr/86Sr
record. Estimates of the production of oceanic crust based on a Phanerozoic plate model (Verard
et al., 2015) and an independent estimate of ocean floor production derived from the pattern of
Mesozoic-Cenozoic subduction deduced from seismic tomography (van der Meer et al., 2014) both
support the notion that variation and amplitude of the 87Sr/86Sr record at first order result from
ridge spreading and subduction zone-length variations (Figure 4.2A). However, in both studies, the
confidence of the estimates decreases back in time due to a lack of preserved oceanic crust (Torsvik
et al., 2010; Müller et al., 2016) and a correlation between imaged slabs in the deep mantle and past
subduction still needs to be quantitatively corroborated (Domeier et al., 2016).
In this paper, we revisit the intriguing correlations between 87Sr/86Sr, plate tectonics, and continental
runoff, and derive a novel method that allows us to reconstruct low-frequency sea level changes
using the 87Sr/86Sr record. This method is summarized in Figure 4.3. We will test the results of our
method with estimates of changing sea level based on sequence stratigraphy, plate tectonic models
and continental flooding.

Sequence stratigraphy and eustatic sea level

Construction of the low frequency curves of Vail et al. (1977) and other estimates derived from
sequence stratigraphy (Figure 4.1) were based on the following general principles. A core concept
of sequence stratigraphy is that changes in the accommodation space of a basin combined with
the relative rates of sediment supply into that basin, ultimately determine the progradation and
retrogradation of the sedimentary systems in that basin (Curray, 1964; Pitman, 1978; Burton et al.,
1987; Jervey, 1988; Posamentier and Vail, 1988; Christie-Blick, 1991; Schlager, 1993). If sediment
supply outpaces the growth of accommodation, then the basin is filled with sediments and the
shoreline retreats seaward. Conversely, if the growth of accommodation outpaces sediment supply,
then the basin deepens and the shoreline transgresses landward. Accommodation is primarily
the result of changes in eustatic sea level, tectonics and to a lesser degree inherited physiography
( Jervey, 1988).
Using these principles, Vail et al. (1977) proposed that it is possible to reconstruct eustatic sea
level by analysing the stratigraphic record. The Vail et al. (1977) curve has since been refined and
expanded in various studies (Haq et al., 1987, 1988; Hardenbol et al., 1998; Haq and Schutter,
2008; Snedden and Liu, 2010, 2011; Haq, 2014; Figure 4.1A).
This model has been widely challenged, however (Hallam, 1984; Hubbard, 1988; Jordan and
Flemings, 1991; Posamentier and Allen, 1993; Schlager, 1993; Miall, 1995). The focus of the
criticism is on the validity of using this method for reconstruction of high frequency (<1 Myr)
fluctuations. While this is certainly a legitimate objection, the fundamental issue is that the model
seems to assume implicitly that changes in accommodation are the key driver in determining
stratigraphic architecture, whilst variations in sediment supply play an insignificant role. This
assumption is unfounded and seems to be incorrect. Although climate change is thought to be able
to cause only transient changes in sediment supply (e.g. Van den Berg van Saparoea and Postma,
2008), it is clear that the response of sedimentary systems to external perturbations is complex (e.g.

47
Paola et al., 1992; Jerolmack and Paola, 2010; Romans et al., 2016). Assuming that climate change
on any time scale has a negligible impact on stratigraphic architecture appears to be unjustified.
Furthermore, it has been argued that large-scale tectonic processes (Cloetingh et al., 1985) and
changes in sediment supply in response to plate tectonic reorganisations (Schlager, 1993; Bonnet
and Crave, 2003) are capable of producing (near-)synchronous responses of similar magnitude.
Even synchronous changes in (global) sea level may not produce synchronous or even similar
responses in different basins (Parkinson and Summerhayes, 1985; Jordan and Flemings, 1991).
Besides these refinements to the original sequence-stratigraphic concepts, additional
processes have come to light that play a significant role in determining the architecture of basin
fills. Dynamic topography (vertical motion of the crust in response to mantle convection on
low frequency timescales of 10-100 Myr) can generate vertical changes in elevation similar in
magnitude to eustatic changes (Gurnis, 1988, 1990, 1992; Burgess and Gurnis, 1995; Moucha et
al., 2008; Spasojevic and Gurnis, 2012; Conrad, 2013). On higher frequency time scales of 1-10
Myr autogenic (internally generated) behaviour of sedimentary systems may be responsible for a
substantial part of the stratigraphic stacking patterns found in the sedimentary record (Muto, 2001;
Muto et al., 2007; Muto and Steel, 2014, 1997; Postma, 2014). Attempts to reconstruct eustatic sea
level from the stratigraphic record alone should be handled with great caution; see Hallam (1984,
2001) and others (Schlager, 1993; Miall, 1995, 2010; Posamentier and James, 2009; Cloetingh and
Haq, 2015).
Finally, the problem of equifinality, i.e. the notion that in open systems a given end state can
be reached by many potential means (Von Bertalanffy, 1968), poses perhaps the most fundamental
challenge to reconstructing past eustatic sea level from the rock record. As Burgess and Prince
(2015) demonstrated, it is likely that different forcing processes or combinations of forcing
processes can produce very similar stratigraphic architectures. Burton et al. (1987) argued that the
impact of a forcing process can only be quantified reliably if the impacts of all the other forcing
processes are known. In other words, an unknown can only be determined reliably if all other
parameters are known. If this is not the case, as in practically every natural system, assumptions have
to be made. This leaves room for substantial uncertainty and a high likelihood of equally plausible
alternative scenarios that are consistent with the observations. The importance of quantifying, or at
least constraining, all of the forcing parameters as much as possible is clear. Until this can be done
more precisely and reliably, it seems prudent to use reconstructions of eustatic sea level based on
sequence stratigraphic interpretations as a reference to check the consistency of results rather than
the primary source of information.
A goal in the analysis of sedimentary basins is thus the development of a eustatic sea level
model that is derived independently from sequence stratigraphic interpretations. To this end,
we here explore whether the 87Sr/86Sr curve may provide an alternative basis to reconstruct low-
frequency eustatic sea level fluctuations.

Methodology:

Plate tectonics and Strontium isotopes


87
Sr/86Sr ratios are measured in marine biogenic and abiogenic carbonates and are assumed to
represent the pristine ratio of 87Sr/86Sr in paleo-oceanic seawater. In our study we use the recent

48
87
Sr/86Sr records of McArthur et al. (2012) and Cox et al. (2016), which are well-constrained for
the Phanerozoic, and data, although less well constrained, extending as far back as the Archean
(Prokoph et al., 2008; Prokoph and Puetz, 2015). In the modern world ocean, the strontium
residence time is 2.5 Myr (Hodell et al., 1990), but this residence time has varied between 1 to
20 Myr since the Late Cambrian (Vollstaedt et al., 2014). Therefore, we will focus on timescales
greater than 20 Myr.
The 87Sr/86Sr ratio of the Earth’s ocean waters results from a mixing of river input (with high
continental 87Sr/86Sr ratios, ~.7136) and input from mid-ocean ridges and volcanic arcs (with low
mantle 87Sr/86Sr ratios, ~.7030). Using the mixing model of Allegre et al. (2010), and assuming that
river input of 87Sr/86Sr has remained constant, van der Meer et al. (2014) produced a 87Sr/86Sr curve
based on evolving subduction zone length that was a first-order match (at 10’s of millions of years)
to the 87Sr/86Sr curve of Prokoph et al. (2008) (Figure 4.2A).
Van der Meer et al. (2010, 2012) used seismic tomographic techniques to image “fossil”
subduction zones in the Earth’s mantle. By mapping these remnants of past subduction, van der
Meer et al. (2014) were able to estimate the total length of subduction zones for time intervals
extending back to the Triassic. These estimates assume that positive seismic wave speed anomalies
correspond to subducted lithospheric remnants. Applying a sinking rate of lower mantle slabs of
12 ± 3 mm/yr, a value based on correlations of seismic tomographic anomalies in the mantle with
geological records of past subduction zones (van der Meer et al., 2010; 2012), total subduction zone
length through time was calculated from depth slices throughout the mantle. This analysis showed
a gradual variation in subduction zone length from ~40,000 km today, ~75,000 km in the Jurassic,
and ~60,000 km in the Triassic. Van der Meer et al. (2014) showed that the subduction zone length
is directly proportional to rates of ocean floor production based on plate tectonic reconstructions
(Müller et al., 2008) if subduction zone length is scaled to production rate through a globally
constant average subduction rate of ~6 cm/yr, similar to the present-day (Schellart et al., 2007). This
suggested that the changing length of subduction zones through time may be used as a proxy for
the production of ocean floor, and hence is a direct corollary of strontium input from the mantle.
Unfortunately, the ‘mantle memory’ of past subduction runs out at the core-mantle-boundary where
slabs are piled up and recycled (van der Meer et al., 2010).
Recently synthetic 87Sr/86Sr ratio curves were calculated on the basis of a Phanerozoic plate
model (Verard et al., 2015). The dominant parameter was the crustal production within the plate
model, but in addition other scenarios were calculated using variable sediment fluxes and changing
the 87Sr/86Sr values of the continental and mantle reservoirs. In Figure 4.2A, we plot the scenario
of Verard et al. (2015) in which they scaled the weathering flux proportional to the volume of
mountain belts of their model. A good fit is established for the Cenozoic, Cretaceous and for the
time period between 325-550 Myr (Figure 4.2A). For the Jurassic, Triassic and Permian portions
of the curve the fit is poorer, which is not surprising considering the diminished oceanic record and
the resulting large uncertainties in crustal production of their model. Despite these uncertainties
their values stay within the overall range of the Phanerozoic 87Sr/86Sr record.
Because these tectonic studies (van der Meer et al., 2014; Verard et al., 2015) obtain a
reasonable first order fit with the 87Sr/86Sr record back through the Mesozoic, it seems reasonable
to use the measured 87Sr/86Sr ratios as a proxy for oceanic crustal production. This approach has
the advantage that high resolution 87Sr/86Sr records are available further back in time than the
Triassic. This would overcome the limitations of plate reconstructions for times when there is no
preserved oceanic crust and the limitations of seismic tomography studies for times when there is
no subducted material preserved in the mantle.

49
Figure 4.3. Schematic of methodology and assumptions to reconstruct low-frequency sea level from 87Sr/86Sr.

50
Integrating plate production over time gives an average ocean floor age, and thus an average depth
of the ocean (Parsons and Sclater, 1977; Stein and Stein, 1992) from which we may infer changes
in global sea level (eustasy). Constraints on the average age of the ocean floor age may provide a
basis for a quantitative first order, low-frequency estimate of eustatic sea level changes. Here, we
explore the simplest-case scenario and assume a constant volume of ocean water with no significant
contribution from either the long-term exchange of water with the mantle, volume changes due
to changes in global temperature or glacio-eustatic variations. We also assume that the volume
of the ocean basins have not been significantly affected by dynamic topography, oceanic plateau
formation, mantle plumes, or sedimentation on ocean sea-floor (Müller et al., 2008; Conrad, 2013).
We assume that the total area of the ocean basins has not changed significantly through time and
that a reliable estimate of the volume of the ocean basins can be obtained by determining the age
and depth distribution of the ocean floor, which is a function of changes in ocean floor production
through time.
Given these assumptions, if changes in the production of ocean floor are known, and the global
volume of seawater is assumed constant, then one can quantify eustatic sea level change through
time. In addition, by combining global sea level changes with continental topography (hypsometry),
one can estimate the amount of continental flooding (area of continental shelves and shallow
epeiric seas). This enables us to test our 87Sr/86Sr derived sea level curve with sea level curves derived
from independent sources and likewise, test our estimates of continental flooding with published
estimates of the ancient area of continental shelves and epeiric seas. In the next section, we will
discuss the applicability of this methodology and identify challenges to overcome.

Correcting 87Sr/86Sr ratios for weathering variations

According to the mixing model of Allegre et al. (2010), the 87Sr/86Sr ratio of sea water is dependent
on input from the weathering of mantle sources (volcanism) and from the weathering of
continental sources. Increased contributions from less radiogenic volcanic sources (mid-ocean ridge,
arc and plume basalts with 87Sr/86Sr ratios of ~.7030) leads to a decrease in the 87Sr/86Sr signature of
seawater. Conversely, increased weathering from more radiogenic continental crust (with 87Sr/86Sr
ratios of ~.7136) leads to an increase of 87Sr/86Sr ratio values. In order to obtain an estimate of
87
Sr/86Sr ratio through time, we therefore need an estimate of the rate of oceanic crust produced
through time as well as an estimate of continental weathering rates through time.
Weathering is strongly dependent on runoff (Dessert et al., 2003; Oliva et al., 2003). Various
estimates of Phanerozoic runoff through time have been proposed (Berner, 1994; Otto-Bliesner,
1995; Gibbs, 1999; Goddéris et al., 2014). The Berner (1994) continental runoff model was
based on an assumed correlation between global temperature and global runoff, modulated
by a continentality factor that took into account the effect of paleogeography on runoff. The
paleogeographic factor was taken from off-line simulations performed with a 3D-climate model
using constant pre-industrial CO2 concentrations and continents with no topography (flat
continents).
Recently, runoff estimates were obtained from simulations performed with the Fast Ocean and
Atmosphere Climate Model (FOAM) which included 22 time slices for the Phanerozoic (Goddéris
et al., 2014) and 7 additional time slices for the Neoproterozoic (Goddéris et al., 2016). The only
forcing functions were the energy input from the Sun and paleogeography. The paleogeographic
reconstructions used in Goddéris et al. (2014, 2016) originated from various sources. Cenozoic and

51
Figure 4.4. Depth to seabed, showing sampled bathymetry from ETOPO1 (Amante and Eakins,
2009) and ocean crust age (Seton et al., 2012) with our calculated best fit.

Figure 4.5. Oceanic area distribution per age bin at present and at 100 Ma of Coltice et al. (2013)
and our linear approximations used in our model.

52
Mesozoic paleogeographies were taken from Herold et al., (2008), Vrielynck and Bouysse, (2003)
and Sewall et al., (2007). Paleozoic paleogeographies were derived from Blakey (2003, 2008).
Neoproterozoic paleogeography was based on Li et al., (2013). Considering the complete set of
paleogeographic reconstructions, the continental land area fluctuates between about 105 and 160
million km2. Assuming a constant volume of seawater, these fluctuations can be either related to
changes in the sea level, continental growth and destruction, or most likely to assumptions, lack of
data, approximates in the various paleogeographic reconstructions or changes in the hypsometry
of the continents due to erosion and tectonic effects. The low spatial resolution of the climate
simulations (7.5° longitude x 4.5° latitude) may also result in an increase or decrease in apparent
land area during the digitization process. The calculated runoff decreases linearly with increasing
land area (Pearson correlation coefficient of -0.66). Strictly speaking, this means that there is a
risk of partial circular reasoning since one parameter used in the sea level reconstruction might
be partly dependent on the sea level itself. At this stage, it is impossible to estimate the separate
contribution of sea level to a relation between runoff and land area, given the variety of methods
used to produce the paleogeographic reconstructions. However, Goddéris et al. (2014) have shown
that runoff correlates strongly with the latitudinal distribution of the continents, and as one might
expect, global runoff is particularly low when a high percentage of land areas are located within the
arid belt (Pearson correlation coefficient -0.84, Goddéris et al., 2014). The relative proportion of
continents located in a specific latitudinal zone (zonal land area divided by the total land area) is
independent of the sea level. This result suggests that the factor controlling the geological history
of runoff is plate motions, with sea level being a second-order factor; although this assumption
deserves future modelling work with the land area held constant through time.
Because rainfall and the resulting runoff is the first-order controlling factor of continental
weathering (Dessert et al., 2003; Oliva et al., 2003), continental weathering rates and associated
flux of radiogenic strontium into the oceans follow the same temporal pattern. Given the direct
relationship between runoff and weathering rates, we used these continental weathering rates to
calculate the other half of the strontium flux equation, namely the flux of 87Sr/86Sr flux from the
mantle (Figure 4.2B). In summary, from the mantle flux of 87Sr/86Sr, we estimated the required
amount of oceanic crust needed to produce that ratio, and from the changing rate of oceanic crust
production we estimated the global age distribution of the ocean floor through time. Finally, from
the changing age distribution if the ocean floor we determined the ocean depth through time.

Depth of the Oceans Through Time

The depth to oceanic basement is classically calculated by empirically fitting curves based on
a half-space thermal cooling model to bathymetric data from the North Pacific and Atlantic
Oceans (Parsons and Sclater, 1977; Stein and Stein, 1992). These empirical curves use a half
space thermal cooling model with a √t relationship (t = the age of the oceanic crust) combined
with an exponential curve to fit the flattening of the depth curve with increasing age. To estimate
the relationship between the age of oceanic lithosphere and the depth of the oceans we adopted a
similar approach based on an empirical √t relationship to account for the cooling of the lithosphere,
and a linear time-dependent relation to account for sedimentation and plume effects.
We sampled the oceanic age grid of (Seton et al., 2012) as well as the present-day ETOPO1
bathymetry (Amante and Eakins, 2009) using a grid of 100x100km cells in an equal-area
Mollweide projection. As shown in Figure 4.4, there is significant scatter and deviation within

53
Figure 4.6. Derived hypsometric curve of ETOPO1 (Amante and Eakins, 2009) and our linear fit
of 0-200 m above present sea level in the insert diagram.

the bathymetry versus ocean age data due to volcanically modified areas (plume swells, hotspot
tracks, seamounts) and areas close to sediment-loaded, passive margins. We accept this scatter as an
inherent feature of the oceans basins and assume that this scatter was a constant, first-order feature
throughout the Phanerozoic and Neoproterozoic. We partly mitigate this scatter by restricting our
curve-fitting to the 1-120 Myr portion of the age distribution. The depth data from ocean floor
older than 120 Myr are few and scattered. Their inclusion results in irregular averages with large
standard deviations.
The relationship between ocean depth (z) and ocean floor age (t) that we use is

z = 2465 + 360 √(t)-15 t (4.1)

where z is the ocean depth in meters, and t is the ocean crust age in million years.

Rates of Ocean Floor Production and the Resulting Age Distribution of Oceanic
Lithosphere

We may use the mantle component strontium flux curve (Figure 4.2B) to estimate the rate of ocean
floor production, and from the rate of ocean floor production we can calculate the corresponding
area-age distribution of the ocean basins. In other words, if the strontium flux ratio doubles, as it
does going from the present-day back to 100 Myr, then the rate of ocean floor production also
would have doubled over the same time interval.
Coltice et al. (2012, 2013) modelled different area-age distributions for ocean floor as a function of
the number and size of continents. They concluded that on model earths with plate-like behaviour,
the average area-age distribution is nearly a linear function, and that the average area-age of the

54
ocean basins is directly correlated with the average spreading rate (Coltice et al., 2012), Following
previous approaches (Cogné and Humler, 2004; Becker et al., 2009; Coltice et al., 2012, 2013), we
use a least-squares method (Figure 4.5) to fit a linear equation to the area versus age relationship
of the present-day oceanic lithosphere from Seton et al. (2012). In Figure 4.5 we illustrate our
linear area-age distributions for the present-day and for 100 Myr, which was a time with higher
crustal production. In the 100 Myr example our theoretical model would indicate that no oceanic
lithosphere older than 90 Myr would be preserved.

We therefore infer that low 87Sr/86Sr ratios (after correction for weathering flux), correspond to
high mantle input due to rapid sea floor spreading (i.e. high oceanic lithosphere production). Rapid
seafloor spreading leads to a larger portion of the total global ocean floor area consisting of young,
more buoyant crust and therefore a steeper slope in the area-age curve.

Volume of the Paleo-Oceans and Continental Flooding

In the next step, we estimate the volume of the paleo-ocean basins by multiplying the area-age
distribution that was estimated from the mantle component of the strontium curve (Figure 4.2B)
with the appropriate ‘age-depth to ocean floor’ (Equation 1). This volume is the amount of water
the world’s paleo-oceans would be able to store relative to present sea level. Any excess water above
or below present-day sea level needs to be redistributed across oceans and continental shelves. This
requires an assumption regarding the average global topography and bathymetry of continental
margins and shelves. This is given by the global hypsometric curve. We therefore first derived the
present-day hypsometric curve by sampling the ETOPO1 digital elevation model (Amante and
Eakins, 2009) using an equal-area Mollweide projection at 5m elevation bins (Figure 4.6). By least-
squares fitting the cumulative area between 0-200 meters elevation, we obtain the following linear
function

Acontinent= c * h + K (4.2)

where Acontinent is area of the continental shelves (km2) that will be flooded by a rise in sea level
of h kilometers. c is the hypsometric gradient of 202,000,000 km2/km, and K is a constant that
depends on the choice of location of the shelf and shelf break. Because we are primarily interested
in the flooded shelfal area relative to the present-day, constant K is irrelevant, and we set K equal
to 0. Consistent with previous approaches (Forney, 1975; Hallam, 1984), we assume that the slope
of the hypsometric curve in the first few hundred meters, is constant and use the linear relationship
in formula (2) to calculate the area of continental flooding for every meter of sea level rise. Excess
water volume (Vexcess) as calculated above, would need to be redistributed across the marine domain,
representing oceans, continental slope and seas (Vmarine) and across the continental domain by
flooding the continents (Vcontinent).

Vexcess=Vmarine+Vcontinent (4.3)

Excess water volume leading to sea level rise is then distributed across areas of the marine domain
(Amarine), and flooded continents (Acontinent).

55
Vmarine= h*Amarine (4.4)

On the basis of our analyses of ETOPO1 (Amante and Eakins, 2009), Amarine amounts to
361,380,356 km2 at present. The flooded continental area Acontinent accommodates only half of
the volume per area, the other half being occupied by topography, therefore:

Vcontinent=h*0.5*Acontinent (4.5)

Inserting formula’s (2), (4), and (5) into (3) leads to:

h2(0.5*c) +h(Amarine)-Vexcess=0 (4.6)

Solving this polynomial function yields:


h=(-Amarine) ± √((Amarine)2-4*(0.5*c *-Vexcess)))/(2*0.5*c) (4.7)

This provides us with h as the resultant water column, or sea level rise relative to the current sea
level. Because sea level on 10 to 100 Myr scales should be fully isostatically compensated (i.e., full
relaxation of ocean basin loading), a mass of mantle equal to that of the added water depth should
become displaced from beneath oceanic lithosphere (Conrad, 2013). Since seawater density is ~30%
that of mantle rock, isostatic compensation of seawater (if fully completed) should cause observed
sea level change to be damped compared to changes in water-level rise (Spasojevic and Gurnis,
2012). This isostatic correction was calculated at 0.689 (Spasojevic and Gurnis, 2012), which we
adopt to calculate absolute sea level rise (habsolute).

habsolute=0.698*h (4.8)

Results

Using the approach outlined above, we have computed a sea level curve that is derived solely from
the 87Sr/86Sr record of seawater (Figure 4.7A). The 87Sr/86Sr -derived curve has notable sea level
lows (<25 m) during the Cambro-Ordovician (460-520 Myr), Triassic (210-250 Myr), and the
Neogene-present-day (0-20 Myr). Highest sea levels (>150 m) are noted in the late Jurassic-late
Cretaceous (160-85 Myr) and prior to the mid-Tonian (~815 Myr).
The 87Sr/86Sr-derived curve shows a first-order 250 Myr cyclicity during the Phanerozoic. A
long-term low for the late Tonian (780-730 Ma) may indicate that cyclicity associated with the
opening and closing of oceans as described by Wilson (1966) extends into the Neoproterozoic.
This 250 Myr cyclicity is in agreement with earlier documented cyclicity for Phanerozoic orogeny,
magmatism, sea level and climate as summarized by Nance et al. (2014). Umbgrove (1940, 1947)
argued for the existence of a ~250 Myr “pulse” whereas Fischer (1981, 1984) argued for two ~300
Myr cycles. Our cyclicity is considerably shorter than the ~400-500 Myr tectonic-climatic cycles
of Worsley et al. (1984) and Veevers (1990), which essentially encompass two of our cycles. Our
cyclicity is also shorter than irregular ~500 Myr to billion year supercontinent cyclicity (Bradley,
2011; Nance et al., 2014; Condie et al., 2015; Van Kranendonk and Kirkland, 2016), which extends
outside our range of sea level reconstruction. The cause of the ~250 Myr cyclicity is to be studied
further, but may be the result of the effect of subduction on heat flow of the core-mantle-boundary

56
(Biggin et al., 2012) and/or is related to the survival time of deeply subducted slabs (van der Meer
et al., 2010).
We also show an alternate sea level curve (Figure 4.7A) that was calculated without the
application of the weathering correction derived from the GEOCLIMtec model (Goddéris et al.,
2014, 2016). The two curves differ notably in the Permo-Triassic and prior to the Cambrian. This
highlights the effect of the weathering strontium flux ratio being substantially different relative to
the present at those times (Figure 4.2B).

First Test: Comparison with Other Eustatic Sea Level Curves

In order to test the validity of our model, we compare our sea level curve (Figure 4.7A) to other
relative sea level curves based on sequence-stratigraphic analyses (Vail et al., 1977; Hallam, 1984;
Algeo and Seslavinsky, 1995; Snedden and Liu, 2010, 2011) or derived from plate tectonic models
(Müller et al., 2008; Verard et al., 2015). The area shown in grey in Figure 4.7A is the bounding
area of the six other sea level estimates (see Figure 4.1A for reference). Our sea level curve falls in
the range of previous estimates of sea level for most of the Phanerozoic, and successfully matches
the 250 million year cyclicity that is seen in the previous analyses. Our curve lies in the lower range
of most sea level estimates for the middle and late Paleozoic (250-450 Myr), and approximately
100 meters below the ranges of previous sea level estimates for the late Cambrian through late
Ordovician (520-450 Myr) (Algeo and Seslavinsky, 1995; Snedden and Liu, 2010, 2011; Verard
et al., 2015). Comparisons cannot be made prior to the Cambrian because there are no published
sea level curves for the Neoproterozoic. Estimates for the Archean suggest that sea levels were
~500 to ~1800 m higher than present-day (Flament et al., 2013). Due to the lack of weathering
corrections, we are not able to use our model to make sea level estimates for times prior to the
Neoproterozoic. However, using an Archean 87Sr/86Sr value of 0.703 (Prokoph et al., 2008; Prokoph
and Puetz, 2015), we would estimate a sea level 425 m higher than present-day. Due to a hotter
mantle, Archean mid-oceanic ridges are generally assumed to have been shallower than today. In
the extreme case where mid-ocean ridges would rise to sea level (Walker and Lohmann, 1989),
we calculate an Archean sea level of +1295 m, which is consistent with reported sea levels of 500-
1800m for the Archean (Flament et al., 2013)

Second Test: Comparison with Continental Flooding

The broad trends in sea level through time can be tested by the independent technique of
estimating continental flooding at successive time intervals (Hallam, 1984). The hypsometric curve,
which describes the areal extent of continental elevations, can be used to estimate the rise in sea
level required to flood varying amounts of continental area. This assumes that ancient hypsometric
curves were similar to the present-day (Forney, 1975; Hallam, 1984). Based on formula (2), which
describes the hypsometric curve for the continental margin and shallow shelves, we follow the
approach of previous studies (Forney, 1975; Algeo and Seslavinsky, 1995) and assume linearity for
the first few hundred meters on either side of the coastline.
Paleogeographic maps that show the amount of continental flooding (millions of ) provide an
independent measurement of past sea-level and have been used to quantify continental flooding.
On the basis of paleogeographic maps (Ronov et al., 1984, 1989; Scotese and Golonka, 1992;

57
Figure 4.7. A) Here derived sea level curves compared with the reference curves from sequence stratigraphic and plate tectonic methods, as
shown in Figure 1A. B) Here derived flooded area curves compared with the reference curves from sequence stratigraphic methods, as shown in
Figure 1B.

58
Golonka et al., 1994; Scotese, 1998; Smith et al., 2004), several Phanerozoic continental flooding
curves have been published over the past decades (Ronov, 1994; Worsley et al., 1994; Otto-
Bliesner, 1995; Walker et al., 2002; Smith et al., 2004; Heine et al., 2015). Following these efforts,
we have constructed curves showing the area of continental flooding based on the most recent
paleogeographic maps of Blakey (2003, 2008), Golonka (2007a, b, 2009a, b, 2012), and Scotese
(2014a, b, c, d, e, f, 2016). For times back to 410 Ma and between 520 to 600 Ma, our 87Sr/86Sr-
derived continental flooding curve is generally within the range of the other continental flooding
curves (Figure 4.1B, 4.7B). Between 410 to 520 Ma, our estimate of continental flooding is up to
20 million km2 less than the other estimates; however, all curves have a similar second-order trends.
We conclude that our new sea level curve and the derived area of continental flooding are, at first-
order, consistent with estimates of continental flooding and sea level of other authors. The causes of
second-order differences, notably in the Late Neoproterozoic to Early Phanerozoic require further
study.

Discussion

Underlying our approach to derive a new low frequency (on the order of 10 to 100 Myr) sea level
curve from 87Sr/86Sr ratios, there are several fundamental assumptions that may require further
verification or refinement. The three most important effects that would impact the first-order
amplitude of sea level derived here are: 1) the constant 87Sr/86Sr values of the strontium inputs used
in the mixing model, 2) the accuracy of the weathering model, and 3) the assumption that the
oceanic crustal ages have a linear relationship with sea floor spreading rates.

Input 87Sr/86Sr variations

Using the mixing model of Allegre et al. (2010), we assume that value of the 87Sr/86Sr component
from the mantle (0.7030) and the value of continental weathering component (0.7136) have been
constant through time, and that the observed variability is simply due the flux of either component.
It is unlikely that the values of the two 87Sr/86Sr components have remained constant because
these components are the result of the mixing of strontium signals from the mid-ocean-ridges, arc
volcanism and ocean-island (plume) basalts for the mantle component, and the mixing of carbonate
and silicate weathering for the continental component. All these components have their own global
average 87Sr/86Sr ratio, but these ratios are likely to have varied to some degree, e.g., as a result
of Himalayan plateau uplift and weathering (Raymo and Ruddiman, 1992; Richter et al., 1992;
Goddéris and François, 1995; Kent and Muttoni, 2008; Verard et al., 2015), continental versus
intra-oceanic arc volcanism (Lee et al., 2012) or flood basalt emplacement and weathering (Mills et
al., 2014).
Van der Meer et al. (2014) showed that for the past 235 Myr the changing length of subduction
zones, which is proportional to the production of oceanic crust at mid-ocean ridges, can explain the
first order shape of the 87Sr/86Sr curve. Mills et al. (2014) showed that a better fit can be obtained
by taking into account temporal variations in the weathering of flood basalts. For the time period
under consideration, flood basalt weathering seems to be of second-order importance. Mills et al.
(2014) estimated that 87% of the 87Sr/86Sr record is due to crustal production at mid ocean ridges,
which from our point-of-view, is the first-order mechanism that drives eustatic changes.

59
Figure 4.8. Comparison of mean oceanic ages of our model, with plate tectonic models (Müller et al., 2016; Seton et al., 2012; Verard et al., 2015)
and with the subduction record (van der Meer et al., 2014).

60
Additionally, flood basalts that erupt in the ocean basins take up a volume that otherwise would
have been occupied by water. This then leads also to eustatic sea level rise (Müller et al., 2008;
Conrad, 2013). Our model may therefore, have underestimated sea level rise. Our linear time-
dependent function partly mitigates this volumetric effect by incorporating the present-day
submerged flood basalts and our indirect assumption of a constant submerged flood basalt volume
through time.

Weathering corrections

The Phanerozoic-Proterozoic weathering correction requires further study. Other climate models
(Berner, 1994; Otto-Bliesner, 1995; Gibbs, 1999) would probably produce different runoff
estimates (Goddéris et al., 2014, 2016) and therefore different weathering corrections for the
observed 87Sr/86Sr values (McArthur et al., 2012; Cox et al., 2016). However, at first-order, these
climate models (Berner, 1994; Otto-Bliesner, 1995; Gibbs, 1999; Goddéris et al., 2014, 2016) agree
with higher runoff during the Jurassic and Cretaceous greenhouse worlds, and lower runoff for the
present-day and the Permo-Triassic. Therefore, the largest correction to the Phanerozoic 87Sr/86Sr
curve (McArthur et al., 2012) would still be made for the Permo-Triassic, and would solve the main
point of contention with Spooner’s observation (Spooner, 1976) regarding the poor correlation
of the measured 87Sr/86Sr values and the sea level curve for the Permo-Triassic (Vail et al., 1977;
Hallam, 1984; Algeo and Seslavinsky, 1995; Snedden and Liu, 2010, 2011; Figure 4.1A).
Other secondary effects may also need to be incorporated in the GEOCLIMtec model (Goddéris
et al., 2014). Topography does not play a role, nor does the contribution of the different erosional
products, i.e. low radiogenic carbonates, flood basalts, arc/hotspot volcanic products and ophiolites,
versus high radiogenic Archean cratons.

Ocean Age distributions

One of the most important assumptions of our approach is that the first-order linear ocean age
distributions shown in Figure 4.5, are a valid approximation of the present-day. In Figure 4.5,
we compare our linear approximation with the estimation of Coltice et al. (2013), who used the
ocean age grids of Seton et al. (2012) sampled at time intervals of 5 Myr. This figure highlights
the excellent first-order fit between a linear function and the present-day area-age distribution. At
second-order, there are differences of up to 0.2 million km2 per age bin, representing a maximum
difference of ~8% at the 25 Myr age bin. For the first-order relations we are pursuing, we therefore
consider the linear function satisfactory. Differences are more substantial when comparing our
curve and the 100 Ma curve of Coltice et al. (2013), with differences of up to 2.4 million km2 per
age bin. However, as discussed before, the 100 Ma age grids (Seton et al., 2012) are only partly
based on data as only a limited amount of crust is preserved and about 50% of the age-grid is
extrapolation of a presumed plate boundary configuration in the oceans. Although there have
been model experiments with deviations from linear distributions (Coltice et al., 2012), the paleo-
age grids beyond the Cenozoic have such large uncertainties that such subtle improvements are
not meaningful. We therefore consider that our linear age distribution assumption satisfactorily
constrains sea level at the first-order.

61
The differences between age distributions in the various plate models are illustrated in Figure
4.8, which shows the mean age of ocean floor through time. Our average mean age fluctuates
between ~30-75 Myr over the past 837 Myr. We compare this range with the studies of Müller et
al. (2016), Verard et al. (2015) and Seton et al. (2012), as well as with a curve calculated using the
subduction record of van der Meer et al. (2014). There is good agreement between the four curves
in the 0-100 Ma interval, but there is greater variation between 100-235 Ma. Prior to 235 Ma, the
study of Verard et al. (2015) has a considerably lower mean age prior to 350 Ma (a mean age of
17 Myr at 535 Ma). Prior to 600 Ma there is no comparable model. Our method, which extends
insights into plate tectonic age distributions ~200 million years into the Neoproterozoic must await
future plate model comparisons.

Conclusions

Recently, it was shown that the record of 87Sr/86Sr can be correlated, at first-order, to the mantle
subduction record and the sea floor spreading models as far back as the Mesozoic. Sea floor
spreading is the first-order driver for sea level and therefore is of fundamental importance to the
interpretation of sequence stratigraphy at regional and larger scales. We correct the 87Sr/86Sr record
for the effects of weathering using estimates of runoff from a recent climate model. From the
corrected strontium record, we calculate the sea floor spreading rate during the last 835 million
years. By assuming linear age distributions and a depth-to-ocean floor-age curve, a new type of
global, eustatic sea level curve is derived. We test this curve by comparing it to estimates of sea level
derived from sequence stratigraphy and plate tectonic models, as well as estimates of continental
flooding. These comparisons show that the sea level curve that we present here is generally within
the range of the other sea level curves, with a few notable exceptions. Because the 87Sr/86Sr-based
eustasy methodology is consistent, verifiable, and has a clear track record, we propose that it should
be considered to be a new, valid method for reconstructing global eustatic sea level during the
Phanerozoic and Neoproterozoic.

62
Chapter 5

Atlas of the Underworld: slab remnants in the


mantle, their sinking history, and a new outlook
on lower mantle viscosity
This chapter is under review as: van der Meer, D.G., van Hinsbergen, D.J.J., Spakman, W., Atlas of the Underworld:
slab remnants in the mantle, their sinking history, and a new outlook on lower mantle viscosity.
Tectonophysics

Across the entire mantle we interpret 94 positive seismic wave-speed anomalies as subducted
lithosphere and associate these slabs with their geological record. We document this as the Atlas
of the Underworld, also accessible online at www.atlas-of-the-underworld.org, a compilation
comprising subduction systems active in the past ~300 Myr. Deeper slabs are correlated to older
geological records, assuming no relative horizontal motions between adjacent slabs following break-
off, using knowledge of global plate circuits, but without assuming a mantle reference frame. The
longest actively subducting slabs identified reach the depth of ~2500 km and some slabs have
impinged on Large Low Shear Velocity Provinces in the deepest mantle. Anomously fast sinking
of some slabs occurs in regions affected by long-term plume rising. We conclude that slab remnants
eventually sink from the upper mantle to the core-mantle boundary. The range in subduction-age
versus –depth in the lower mantle is largely inherited from the upper mantle history of subduction.
We find a significant depth variation in average sinking speed of slabs. At the top of the lower
mantle average slab sinking speeds are between 10-40 mm/yr, followed by a deceleration to 10-15
mm/yr down to depths around 1600-1700 km. In this interval, in situ time-stationary sinking rates
suggest deceleration from 20-30 mm/yr to 4-8 mm/yr, increasing to 12-15 mm/yr below 2000 km.
This corroborates the existence of a slab deceleration zone but we do not observe long-term (>60
Myr) slab stagnation, excluding long-term stagnation due to compositional effects. Conversion of
slab sinking profiles to viscosity profiles shows the general trend that mantle viscosity increases in
the slab deceleration zone below which viscosity slowly decreases in the deep mantle. This is at
variance with most published viscosity profiles that are derived from different observations, but
agrees qualitatively with recent viscosity profiles suggested from material experiments.

Introduction

Seismic tomography has provided a breakthrough in the analysis of plate tectonic history by
allowing to trace now-subducted ancient lithosphere in the Earth’s mantle, where they appear as
plate-like positive seismic wave-speed anomalies in the upper mantle and often more amorphous
structures in the lower mantle (e.g., Spakman et al., 1988; van der Hilst, 1991; Zhao et al. 1992;
Fukao et al., 1992; 2001; Grand et al., 1997; van der Hilst et al. 1997; Bijwaard et al., 1998;
Widiyantoro et al. 2000; Zhao 2004; Fukao and Obayashi 2013). Since the earliest detection of
positive wavespeed anomalies in the deep mantle (Dziewonski, 1984) their interpretation was

63
almost exclusively in terms of remnants of ancient subduction. Only recently, however, this
presupposed link was corroborated as being statistically significant at the 99% confidence level for
subduction of the past ~120 Myr reaching mantle depths of ~2300 km (Domeier et al. 2016).
Research into linking detailed plate tectonic history to present-day mantle structure started around
1990 (e.g. Spakman et al. 1988; Richards and Engebretson 1992; van der Hilst and Seno 1993; De
Jonge et al. 1994; Carminati et al. 1998a; Lithgow-Bertelloni and Richards 1998). Improved global
mantle tomography models led to interpretations of lower-mantle slab structure which initially
focussed on the large Farallon and Tethys anomalies (e.g. Grand et al. 1997; van der Hilst et al.
1997, Bijwaard et al. 1998; van der Voo et al. 1999a; Hafkenscheid et al, 2006) and the Mongol-
Okhotsk (van der Voo et al. 1999b) as third important anchor slab for linking deep mantle structure
to plate tectonic evolution. A major jump was made by van der Meer et al. (2010) who linked
present-day lower mantle structure to the paleo-location of 28 ancient subduction zones, of which
the geological artefacts have moved away by plate motion. Van der Meer et al. (2010; 2012) used
geological dating for the start and end of subduction (e.g. van Hinsbergen et al. 2005) combined
with a detailed relative plate-motion model (Torsvik et al. 2008) to identify slab remnants in the
lower mantle that effectively comprised ~250 Myr of global subduction history. They also derived
a first-order estimate of the average sinking rate of slabs through the lower mantle (12±3 mm/
yr), a range consistent with later estimates (Butterworth et al. 2013, Sigloch and Mihalynuk 2013,
Domeier et al. 2016), which was also used as a novel constraint on mantle viscosity (Cízková et
al., 2012; Bower et al., 2013). Importantly, the identification of slab remnants that subducted at 28
paleosubduction zones provided the first step towards establishing a mantle reference frame based
on subduction (van der Meer et al., 2010) that proved to be relatively close to absolute plate motion
models that are primarily based on hot spot tracks (e.g. Doubrovine et al. 2012).
Recently major progress occurred on linking geological history to slab remnants imaged in the
mantle at regional scale (e.g. Shephard et al., 2013; Sigloch and Mihalynuk 2013; van Hinsbergen
et al., 2014; Hall and Spakman 2015; Zahirovic et al., 2016). By invoking absolute plate motion
models, relatively isolated slab remnants in the deep mantle, of which the geological evidence is
meanwhile displaced by thousands of kilometers, could be identified (e.g. Schellart et al. 2009;
Schellart and Spakman, 2015; Shephard et al., 2016; Vissers et al., 2016). New approaches use
imaged slabs directly as a novel basis for plate kinematic restorations by “surfacing” slab anomalies
(Lister et al. 2012; Wu et al., 2016). Furthermore, mapped slab remnants were correlated with
the location of seismic scatterers in the lower mantle (e.g., Hutko et al., 2006; Kaneshima 2013,
van der Meer et al., 2012, Ma et al. 2016) and with Pacific LLSVP topography (He and Wen,
2009; van der Meer et al., 2012). Global interpretation of mantle structure has led to estimates
of paleo-subduction zone lengths through time that provided constraints for global plate tectonic
activity impacting atmospheric CO2 (van der Meer et al., 2014, Mills et al., 2014, Kashiwagi, 2016),
strontium isotope ratios (van der Meer et al., 2014, 2017) and sea level (van der Meer et al., 2017).
Encouraged by all these new developments, we have expanded our analysis of mantle structure
to the identification of 94 slabs throughout the upper and lower mantle. This review constitutes our
geological interpretation of these imaged slab remnants by linking them to geological records of
subduction. The compilation is summarized in what we call the Atlas of the Underworld, for which
we also developed an online version at www.atlas-of-the-underworld.org that is fully searchable
and includes discussion forums for each interpreted slab to facilitate post-publication peer review.
This catalogue is intended to provide a new basis for future studies in the fields of orogenesis,
mantle convection and plate tectonic reconstruction, and as a first and extensive global framework

64
for interpretation of present-day global mantle structure and its physical properties and how that
relates to our planet’s dynamic evolution of the past ~300 Myr.

Methods

We aim to provide a succinct documentation of many positive seismic wavespeed anomalies in the
mantle as remnants of subducted lithosphere of which we interpret plate tectonic provenance and
the timing of subduction from the global continental geological record of subduction. We expand
on our previous work (van der Meer et al., 2010) by (1) incorporating upper mantle slab anomalies,
(2) by identifying contributions within major mantle anomalies (particularly below North America)
from individual paleo-subduction events, and (3) by incorporating lower mantle anomalies that
were not yet subject to interpretation in terms of subducted lithosphere, if possible.
Our previous compilation (van der Meer et al., 2010) suggested that the deep mantle contains slab
remnants that may have subducted 200-300 Myr ago. Whilst actively subducting upper mantle
slabs are straightforwardly linked to the active orogenesis and to overlying arcs, correlating such
deep, detached slabs to geological records is not straightforward. Any detached lower mantle slab
may, when viewed in isolation, be associated with a large number of geological records, since post-
detachment plate motions may have displaced these records over thousands of kilometres relative
to the location where its slab remnant sinks in the mantle. One way of assessing this problem is
by investigating the past location of a particular geological subduction record in an ‘absolute’
mantle reference frame which potentially may restore the record above a candidate slab remnant
in the underlying mantle (e.g. Schellart et al 2009; 2012; van der Meer et al. 2012; Sigloch and
Mihalynuk 2013; Schellart and Spakman 2015; Vissers et al. 2016). However, with the exception of
two slab anomalies (Malpelo and Mesopotamia), here we refrain from using absolute plate models
as an identification guide because we ultimately aim to use the imaged slab remnants as a constraint
on such absolute plate motion models. Instead, we took the following approach in our correlation
attempt, which is independent of any absolute plate motion frame:
1) We identified positive P-wave speed anomalies in the mantle that occur in the UU-P07
tomography model (Amaru, 2007, where it was called P06; also described in Hall and Spakman
2015 and available on the Atlas website) as well as in S-wave anomaly models that are independent
of model UU-P07 in terms of both seismological data-type used and inversion methodology. For
the lower mantle we examined S40RTS (Ritsema et al., 2011) while for the upper mantle we used
the surface wave model SL2013sv (Schaeffer and Lebedev, 2013). Occurrence of positive anomalies
in these independent models, albeit with different spatial resolution, was taken as a strong
indication of the existence of corresponding mantle heterogeneity. Although here we compare
only one P-model to one independent S-model in the upper and lower mantle, we note that a
thorough comparison of the structural content of 14 recent P- and S-models, including UU-P07
and S40RTS, through the construction of detailed voting maps (Lekic et al. 2012) is currently
underway (Shephard et al., On the consistency of tomographically imaged lower mantle slabs,
submitted). A preliminary comparison shows an overall match between these voting maps with the
anomalies presented here in the Atlas of the Underworld (G. Shephard, personal communication).
Following the approach of Hafkenscheid et al. (2006) and van der Meer et al (2010), we then
estimated the depth range of tops and bases of slabs for P-wave anomaly amplitudes greater than
0.2% in the UU-P07 model. We label each slab by the name of the geographic region under which
the mantle anomalies occur so as to avoid an a priori paleogeographic interpretation of an anomaly

65
Figure 5.1. Interpretation flowchart Tethyan realm. Blue bars indicate depth extent of slabs.
Red arrows show the used inter-slab correlations to constrain geological interpretations of slab
subduction history.

and also allow for future corrections of our interpretations, which then needs not involve the
anomaly name.
As outlined above, correlation of detached slab anomalies to geological records becomes more
difficult with increasing depth. In the Atlas of the Underworld, we aimed to provide concise
descriptions of our correlations, citing the pertinent literature, while avoiding description of the
geological records to which we did not correlate the slab under discussion. Our interpretation
strategy starts in the upper mantle with generally well-identifiable slabs. We work our way down
in the mantle and side-ways by considering new anomalies in the same, or wider, region occuring
at comparable or larger depth. Viable interpretations are next considered in the geological context
of previously identified slabs for the same region, or in a wider region if necessary. Clearly, such
interpretation process requires stepping backward (questioning previous interpretations at
comparable or shallower depth) and forward (evaluating all possible candidate subduction zones).
In more detail, the geological interpretation of slabs was carried out as follows:
2) We first interpreted slabs that are still associated with active subduction zones, orogens,
and arcs. From these geological records, we estimated the age range of the onset of these subduction
zones using a combination of the following data types: the onset of accretion of thin-skinned
nappes or an accretionary prism, the onset of a period of arc volcanism that lasts until today, the
oldest high-pressure, low-temperature metamorphic rocks found in an accretionary orogen, and, in
the case of subduction below oceanic lithosphere preserved as ophiolites, the ages of metamorphic
soles and supra-subduction zone ophiolites. For each individual anomaly, the Atlas provides
documentation of the geological data that were used for their interpreted age of subduction.

66
Figure 5.2. Interpretation flowchart American realm. Legend same as Figure 1.

3) Subsequently, detached slabs were interpreted, whereby we emphasize that we made no


assumption on slab sinking rates. We do assume, however, that regionally adjacent slabs did not
undergo significant lateral motions relative to each other after slab break-off. This means that we
assume slabs to sink dominantly vertically after detachment, as recently shown viable by Domeier
et al. (2016). Progressively deeper slabs were generally correlated to progressively older geological
records of subduction. In the interpretation process of a particular slab anomaly we used other slabs
in the region at shallower or comparable depth as a guide for locating the corresponding geological
record. Particularly in the lower half of the mantle, identification of several slab remnants depends
in this way on the identification of slab remnants at shallower depth. These interpretation links
are made explicit in several schematic figures: Figure 5.1 for the Tethyan realm, Figure 5.2 for
the American realm, and Figure 5.3 for the Australasian realm. Unavoidably, to give direction to
our search for suitable candidate records we initially assume that shallower slabs have generally
younger subduction ages than deeper slabs, and that slabs residing at similar mantle depths may
have roughly similar subduction ages, although the range in subduction age we strictly adopt from
the literature and may comprise up to 100 Myr for slabs at a particular depth. Generally, slabs are
correlated to the youngest possible record of subduction that was not yet associated with shallower
slab remnants.
4) Age ranges for slab break-off are based on a combination of the following data types: the
end of accretion in an orogen, the end of deformation in a plate boundary zone associated with an
orogen, the arrest of volcanism, or the type of volcanism (e.g., adakites), or the arrest of convergence
reconstructed from published plate tectonic reconstructions.
We illustrate this approach by the following example of slab anomalies in the mid-mantle
below northwest Africa. For literature references to stated ages and descriptions of the associated

67
Figure 5.3. Interpretation flowchart Australasian realm. Legend same as Figure 1.

subduction records, we refer to the sections in which the slabs are described. Below NW Africa, the
Algerian slab (section 5.4) is a N-S trending body at ~1300-2400 km depth located below Algeria
and Mali, and the small Reggane slab (section 5.71) is located between 1500 and 1900 km depth
to the west of the Algerian slab. For correlating these slab anomalies to their geological records
we continue on the regional analysis we made for associating slab anomalies at the same depth
or shallower to their respective orogens. To the north of the Algeria and Reggane slabs lie upper
mantle slabs of Gibraltar (section 5.34), the Kabylides slab below northern Algeria (section 5.46),
and the Calabrian slab in southern Italy (section 5.20). These slabs have a record of subduction
in their associated circum-Mediterranean orogens dating back to ~45-85 Ma, depending on the
literature. This precludes correlation of the Algeria and Reggane lower mantle slabs to the active
circum-Mediterranean orogens, where high-temperature metamorphic rocks in Calabria and the
Betic Cordillera provided Eocene, ~45 Ma ages demonstrating that subduction was underway by
this time.
To the northeast of the Algerian slab, the Aegean slab (section 5.1) is found down to depth
of ~1300-1500 km, i.e. at a depth range corresponding to the top of the Algerian and Reggane
slabs. The Aegean orogen in the overriding plate suggests that the slab has continuously subducted
since 120-100 Ma. Immediately below the Aegean slab lies the NW-SE trending East Vardar slab
(section 5.32) between ~1300 and 2100 km, i.e. in the same depth interval as the Algerian slab
suggesting that the Algerian and East Vardar slabs subducted simultaneously along two adjacent
subduction zones. The geological record of the Eastern Mediterranean region shows a double
Jurassic ophiolite belt – the West and East Vardar ophiolite belts – with ~170 Ma metamorphic
sole and supra-subduction zone ophiolitic crustal ages, one emplaced north and eastwards onto
Eurasian crust, and one emplaced westwards onto continental crust of Adria, a microcontinental

68
Figure 5.4. Midpoints location map. Blue: slab midpoints projected to surface. Red: geological records. Arrows indicate the correlations between
slabs and their corresponding geological record and show the displacement of the geological records over time relative to the relatively fixed
locations of sinking slabs.

69
fragment that existed north of Africa surrounded by Alpine and Neotethyan oceanic basins. The
ages of obduction are 130-120 Ma. The correspondence between two adjacent slabs with similar
depth ranges and two records of subduction with similar ages then leads to the correlation of the
Algerian slab to the formation (onset of subduction) and emplacement (end of subduction) of the
West Vardar ophiolite belt, and the East Vardar slab correlated to the East Vardar ophiolite belt.
The Reggane slab to the west of the Algerian slab is then best correlated to a subduction record to
the west, but in the vicinity of Adria. There, plate kinematic reconstruction of the Central Atlantic
Ocean and Bay of Biscay, and paleomagnetic data from Iberia predict up to ~500 km convergence
across the Pyrenees starting ~ 125 Ma and that ending around ~110 Ma, which is followed by
a phase of high-temperature metamorphism and associated volcanism around 100 Ma, which is
then interpreted to reflect the timing of break-off of the Reggane slab. With these correlations, all
geological records of subduction of the (southern) Mediterranean region are correlated to slabs, and
all slab anomalies in the upper and lower mantle below the southern Mediterranean and North
Africa are then correlated to geological records of subduction.

Atlas of the Underworld

In the Atlas below, we describe the geological evidence on which we base the duration of
subduction for each slab. We provide a location, images from the tomographic models UU-P07 and
S40RTS and SL2013sv tomographic models in map-view and cross-section, and the location of the
correlated geological record. In cross-sections, we plot SL2013sv in the upper mantle and S40RTS
in the lower mantle. For every anomaly, we provide 6 cross-sections centred on the midpoint of the
anomaly in the online version of the Atlas at www.atlas-of-the-underworld.org, as well as spike
resolution tests (Spakman and Nolet 1988; Rawlinson and Spakman 2015) for each anomaly in
model UU-P07. Depths and interpreted subduction ages of all slabs are summarized in Table A1.

5.1. Aegean - Aeg


The Aegean anomaly (Figure 5.5) was previously called ‘Aegean Tethys’ by van der Meer et al.
(2010) and is located below south-east Europe from the deep lower mantle up to the surface at the
Hellenic subduction zone. It was one of the first slabs imaged (Spakman et al., 1986a;1988; 1993)
and is widely interpreted as a slab that represents north-dipping African lithosphere (e.g. Spakman
et al. 1988; De Jonge et al., 1994; Piromallo and Morelli, 1997; 2004; Bijwaard et al., 1998; de
Boorder et al., 1998; Faccenna et al., 2003; van Hinsbergen et al., 2005; 2010; Hafkenscheid et al.,
2006; Chang et al., 2010; Biryol et al., 2011). Van der Voo et al. (1999b) identified this slab as the II
anomaly, and Hafkenscheid et al. (2006) called it the ‘Gr’ anomaly, with both papers identifying that
the slab deeply penetrates the lower mantle. Where Faccenna et al. (2003) initially interpreted the
mantle below the Aegean slab to host a single slab that folded and overturned in the lower mantle,
van Hinsbergen et al. (2005) suggested that there are actually two slabs below the Aegean: A deep
anomaly between ~2000 and 1500 km, and an anomaly from 1500 km to the surface, associated
with continuous northward subduction of oceanic and continental lithosphere, upper crust which
accreted to form the Aegean mountain belt on basis of the geological record of the Aegean region,
where three separate subduction records are found: A Jurassic – Early Cretaceous record emplaced
ophiolites westward and northeastward suggesting two oppositely dipping subduction zones of this
age. The northeastern one we associate with the deeper Aegean slab (see ‘East-Vardar slab, section
5.33) and the western with the Algeria slab to the west (section 5.4). Van der Meer et al. (2010)

70
Figure 5.5. Aegean anomaly. (Horizontal) [vertical] cross sections through (A)[E] the UUP07
P-wave) and (B)[F] the combined SL2013 and S40RTS S-wave tomographic models; C) red
line marks the location of the modern geological record that we interpret to have formed during
the subduction of the slab; D) location map of vertical cross-sections E and F. Relative amplitude
strength, vertical, lateral extent and dip trend are very similar between tomographic models.

included the East Vardar and Aegean slabs both in the Aegean Tethys slab and assigned an age
range of mid-Jurassic to Present. Because the subduction of both slabs can be clearly separated
in the geological record, we separate these into two separate lithospheric bodies. For the Aegean
slab we avoid the addition “Tethys” because a considerable part of the subducted lithosphere is of
continental origin rather than the oceanic nature of the Tethys (van Hinsbergen et al., 2005; 2010).
The onset of northward subduction of the Aegean slab must have started after the emplacement
of the East Vardar ophiolites in the early Cretaceous (~130-120 Ma (Schmid et al., 2008)) (See
East Vardar slab, section 5.33), and before the onset of arc volcanism in the Sredna-Gorie belt of
southern Bulgaria starting around 92 Ma (Stoykov et al., 2004; Quadt et al., 2005; Zimmerman
et al., 2007), and is generally considered to occur around ~110±10 Ma (e.g., Faccenna et al., 2003;
van Hinsbergen et al., 2005; 2010; Schmid et al., 2008; Jolivet and Brun, 2010). Subduction of
the Aegean slab is still continuous at the Hellenic trench, and seems in the upper mantle laterally
disconnected to slabs under western Anatolia (de Boorder et al., 1998; van Hinsbergen et al., 2010;
Biryol et al., 2011) Figure 5.6). Based on a kinematic reconstruction of western Turkey, the timing
of break-off of the Aegean slab below Western Anatolia was estimated at ~15 Ma (van Hinsbergen,
2010; van Hinsbergen et al., 2010).

5.2. Agattu - Agt


The Agattu anomaly (Figure 5.6) is a detached slab located below the northernmost Pacific Ocean,
in the upper part of the upper mantle and uppermost part of the lower mantle, predominantly south
and west of the Aleutian trench. At its deepest it is W-E striking and south dipping, changing

71
Figure 5.6. Agattu anomaly. Legend same as figure 5.5. Relative amplitude strength, and lateral
extent are fairly similar between tomographic models. Vertical extent and dip trend differ.

to a NW-SE orientation in the uppermost lower mantle. At its shallowest it is located south of
the Aleutian slab (section 5.3). Van der Meer et al. (2010) interpreted it as a slab representing
the westernmost part of the North Pacific slab (section 5.68). The tectonic model of Shapiro and
Solov’ev (2009) now allow us to refine this interpretation, and define the Agattu slab separately.
The Agattu slab is detached and its top is located at a similar depth as the base of the actively
subducting Aleutian slab, which started to subduct around 56-46 Ma (section 5.3). We hence
searched for geological records in the NW Pacific realm of subduction that ceased in Eocene time.
At a latitude consistent with this intra-oceanic location, Levashova et al. (2000) and Shapiro and
Solov’ev (2009) used paleomagnetic data to reconstruct the Achaivayam–Valaginsky terrane. This
terrane comprises Upper Cretaceous and Lower Paleocene volcanic and sedimentary rocks of
diverse but mostly submarine facies, with older Mesozoic volcano-sedimentary blocks and ophiolite
fragments. The arc originated at about 90–85 Ma (Coniacian–Santonian) as an island arc built
upon oceanic crust (Bogdanov et al., 1987; Sokolov, 1992, as cited in Shapiro and Solov’ev (2009),
subducting to the southeast and culminating in ophiolite emplacement onto Kamchatka, Sakhalin,
and Hokkaido around 50-45 Ma (e.g., Nokleberg et al., 2000). This southeastward subduction zone
is not portrayed in most Pacific plate models (e.g., Seton et al., 2015; Müller et al., 2016; Torsvik et
al., 2017) who do not discuss and incorporate the geological record of the northwest Pacific region.
Paleomagnetic data from the Cretaceous-Paleocene ophiolites and arc rocks of Kamchatka indicate
about 2000 km northward transport of the arc terrane from its initial paleolatitude to its modern
position at the Eurasian margin between the Campanian (~75 Ma) and the Middle Eocene (50–
45 Ma) (Levashova, 1999; Kovalenko, 2003, as cited in Shapiro and Solov’ev (2009) suggesting
north(west)ward roll-back of the slab upon Pacific plate advance. The slab geometry suggests that it
was partly formed during NW ward transport towards the Eurasian margin. We therefore interpret
the top of the slab to have subducted sometime between 50-45 Ma shortly following ophiolite and
arc obduction onto the continental margin.

72
Figure 5.7. Aleutian anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model, and lateral extent is very similar with the S-wave model. Dip trend, vertical extent are very
different between models.

5.3. Aleutian - Al
The Aleutian anomaly (Figure 5.7) is located in the upper mantle below the southern Bering Sea
and Alaska and connects to the surface in the Aleutian subduction zone. It was previously identified
in the upper mantle (Gorbatov et al., 2000; Qi et al., 2007; Koulakov et al., 2011; Martin-Short et
al., 2016). In the UU-P07 model the Aleutian slab is continuous and penetrates the top of the lower
mantle below the southern Bering Sea. Van der Meer et al. (2010) initially assumed an inferred
onset of magmatism in the Kluane arc at the southern margin of Alaska in the late Cretaceous
to Eocene (Nokleberg et al., 2000), as a base age of the slab. We now correlate the Kluane Arc to
a more easterly, deeper slab (Yukon slab, section 5.93). Instead, we use the onset of subduction at
the Aleutian trench as age for the base of the slab. In the tectonic reconstruction of Shapiro and
Solov’ev (2009), 56-54 Ma was inferred as onset of subduction below the Aleutians, based on a
change in plate motion of the Pacific plate (Creager and Scholl, 1973; Scholl et al., 1989; Geist et
al., 1994). 40Ar/39Ar datings of volcanic and plutonic rocks from several islands along the Aleutian
arc shows that volcanism was underway by 46 Ma ( Jicha et al., 2006, Chekhovich et al. (2014),
providing a minimum age for the onset of subduction. Chekhovich et al. (2014) estimated the time
of origination of the Aleutian subduction zone is estimated at to be 47 ± (2–3) Ma. We therefore
adopt a 50-46 Ma age range for the onset of subduction of the Aleutian slab.

5.4. Algeria - Ag
The Algeria anomaly (Figure 5.8) is a detached slab located below NW Africa and the
Mediterranean from the deep mantle up to mid-mantle and was first described by van der
Meer et al. (2010). The neighbouring base of the East Vardar slab (section 5.33) and the N-S
trend of the Algeria anomaly is used to infer that it is the best candidate to represent eastward

73
Figure 5.8. Algeria anomaly. Legend same as figure 5.5. Lateral location is similar between
tomographic models. Relative amplitude strength is weaker in the used S-wave models. Lateral and
vertical extent differs between tomographic models. Dip trends are not imaged.

subducted lithosphere below the West Vardar ophiolites sensu Schmid et al. (2008), found from
the Pannonian Basin to southern Greece over a reconstructed distance (e.g., Maffione et al.,
2015a) similar to the N-S length of the Algerian slab. The slab is interpreted to represent oceanic
lithosphere that was attached to Adria and that subducted below these ophiolites. This lithosphere
was referred to as the Meliata-Maliac ocean by Stampfli and Borel (2004). The onset of subduction
below the West-Vardar ophiolites likely shortly predated the cooling of metamorphic soles below
the west-Vardar ophiolites of Greece, Albania and the Dinarides, dated at 174-157 Ma, by
several million years (Dimo-Lahitte et al., 2001; Liati et al., 2004; Karamata, 2006; Šoštarić et al.,
2014): we consequently interpret the Algeria anomaly as a slab with a 175±5 Ma age for the base.
Obduction of the ophiolites onto eastern Adria, currently in the Dinarides-Hellenides was shown
by sedimentary overlap assemblages in the Dinarides and Hellenides to have finalized in the early
Cretaceous, around 130±10 Ma (Schmid et al., 2008; Scherreiks et al., 2014; Tremblay et al., 2015;
Maffione et al., 2015a), which we take as the age for the top of the slab.

5.5. Al Jawf - AJ
A NW-SE elongated anomaly is imaged overlying the core-mantle boundary below the Red Sea
and most of Arabia, which we here define for the first time as the Al Jawf anomaly (Figure 5.9).
Resolution tests show that the anomaly is poorly resolved, and our interpretation is therefore
tentative. In the west, the body reached shallower depths, rising up to ~2400 km depth into
the lower mantle. Based on correlation with the shallower East Vardar slab (section 5.33), the
Mesopotamia slab (section 5.64) towards the North, and the Central China slab (section 5.30) and
the Mongol-Okhotsk slab (section 5.67) at similar depths to the NE, we infer the Al Jawf slab
likely formed by subduction to the south of Laurasia in pre-middle Jurassic time. We consequently
interpret this anomaly as Paleotethys lithosphere that subducted during the opening of the

74
Figure 5.9. Al Jawf anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with a similar vertical extent, flat on top of the CMB.
Relative amplitude strength, lateral extent differ between tomographic models. In the S-wave
model this anomaly is much larger and merges with the Balkan anomaly to the NW.

Neotethys between the Gondwana-derived ‘Cimmerian’ continental fragments and Arabia-Africa.


Neotethys opening and Paleotethys subduction is not the same age in the Tethyan realm. Between
the Cimmerian continents of Iran and Arabia is paleomagnetically and stratigraphically constrained
to occur between the Late Permian and the Early Triassic collision of the Iranian Cimmerian
continents with Eurasia, 230±10 Ma (Stampfli and Borel, 2002; Muttoni et al., 2009). Closure of
the Paleotethys in the Eastern Mediterranean region was younger, and may either have occurred
by northward (Okay and Nikishin, 2015), or, as more widely accepted, southward subduction
(Sengör and Yilmaz, 1981; Dokuz et al., 2017) below the Sakarya continental fragment. A record
of subduction of is reflected by the Karakaya mélange complex in northern Turkey. This complex
contains Devonian radiolarian cherts, showing that Paleozoic oceanic crust was consumed (Okay et
al., 2011). 40Ar/39Ar ages of eclogite blocks in the Karakaya subduction mélange of 215-203 Ma
(Okay and Monié, 1997; Okay et al., 2002) show that subduction was active until the latest Triassic.
This mélange is unconformably overlain by Lower Jurassic limestones which may mark the end of
subduction around ~200 Ma (see Sayıt and Göncüoglu, 2013 for a review). In addition, Middle
Jurassic, ~170 Ma old lavas in the Pontides were interpreted to reflect Paleotethys slab break-
off (Dokuz et al., 2017) and we adopt a 200-170 Ma age for the top of the Al Jawf slab. Since
there was minimal convergence between Gondwana and Laurasia in the Triassic, the closure of
the Paleotethys was almost entirely accommodated by contemporaneous opening of the Neotethys
(Gutiérrez-Alonso et al., 2008). The oldest radiolarian cherts obtained from mélanges that formed
during subduction of the Neotethys in Turkey along the Izmir-Ankara suture zone are Ladinian-
Carnian in age (242-228 Ma) (Tekin et al., 2002; Tekin and Göncüoğlu, 2007) provide a minimum
age for opening of the Neotethys and thus for the onset of subduction of the Paleotethys. Because
shallower slabs in the mantle in the region of the Al Jawf slab were all correlated to Jurassic and

75
Figure 5.10. Alps anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical, lateral
extent are similar between tomographic models. The slab lies flat in the transition zone and top of
the lower mantle and does not have a dip trend.

younger geological records, we correlate the Al Jawf slab to the youngest, Triassic-Jurassic record of
Paleotethys subduction in the eastern Mediterranean region and therefore adopt a minimum age of
240 Ma for the base of the western end of the Al Jawf anomaly, and a 200 Ma age for the top. The
S-wave tomographic images suggest that to the east, there may be more lithosphere present that
may be the relics of the Permo-Triassic slabs of the Paleotethys that formed east of the Al Jawf slab,
but in the UU-P07 model, we find no sufficient resolution to independently confirm this.

5.6. Alps - Alp


The Alps anomaly (Figure 5.10) was first detected by Spakman (1986b), Kissling (1993) and
Kissling and Spakman (1996) and was interpreted by Spakman and Wortel (2004) as the remnant
of the Piemonte-Ligurian ocean that subducted in the Alps. The Alps anomaly is interpreted as
a curved body of detached lithosphere that follows the oroclinal shape of the western Alps, and
lies on the 660 km discontinuity, with its base penetrating the upper part of the lower mantle,
reaching ~800 km depth (Figure 5.10). The Alps slab is disconnected from shallower anomalies
below the western and eastern Alps that are imaged to less than 200 km depth (Lippitsch et al.,
2003; Kissling, 2008) and that are interpreted as European continental lithosphere in the west,
and Adriatic continental lithosphere in the east that subducted in the last 20 Ma (e.g., Schmid
et al., 2004; Handy et al., 2010; 2014). The Alps slab likely represents oceanic lithosphere of the
Piemonte-Ligurian and Valais oceans and continental lithosphere of the intervening Briançonnais
microcontinent (Handy et al., 2014) that subducted in the Alps. The oldest definitive record of
subduction, with a southward vergence, in the geological record of the western Alps is the HP-LT
metamorphism of the Sesia zone, with an age around 75-65 Ma (Handy et al., 2010) and references
therein). Subduction must have started prior to this time, and may have initiated due to a change
in relative Africa-Europe convergence from E-W to N-S around 85 Ma (Handy et al., 2014),

76
Figure 5.11. Anhui anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical, lateral
extent are similar between tomographic models. The slab lies flat in the transition zone and top of
the lower mantle and does not have a dip trend.

which we adopt as the age of the base of the Alps slab, with a somewhat arbitrary but conservative
uncertainty estimate of ±10 Ma. Break-off of the Alps slap was interpreted to be reflected by a
series of granitoid intrusions of 33±8 Ma (von Blanckenburg and Davies, 1995). In more recent
years also rapid uplift and exhumation in the Alps in the Early Oligocene was interpreted to
possibly reflect a phase of slab break-off ( Jourdan et al., 2013). In any case, breakoff of the Alps slab
in the eastern Alps must have occurred prior to the onset of northward underthrusting of Adriatic
lithosphere, which according to geological interpretations of the southern Alps occurred since ~20
Ma (Ustaszewski et al., 2008). Following these constraints, we conservatively adopt an age for the
top of the Alps slab of 30 ± 10 Ma.

5.7. Anhui - Anh


The Anhui anomaly (Figure 5.11) is located in the upper part of the lower mantle and lower part
of upper mantle below southeastern China. It has previously been imaged by Wei et al. (2012), who
suggested it may represent Pacific or Izanagi lithosphere that detached from the present Pacific slab.
Its shape is irregular at the base and forms a broad anomaly at 810 km depth. At shallower depths
it thins and is not imaged above the 440-km discontinuity. Towards the north, the Manchuria
slab (section 5.54) shares similar flat-slab characteristics, although the Anhui anomaly is located
somewhat deeper in the mantle. We interpret this anomaly as the Anhui slab with the subduction
history below the South China block postdating the subduction of the East China slab (section
5.30).
Following the formation of a slab window and tectonic inversion as a result of ridge subduction
and/or collision with the West Philippine block (Li et al., 2014a), subduction below South
China recommenced with arc magmatism and crustal extension (107–86 Ma). The exact end of
that phase of subduction is unclear. Generally, Cretaceous magmatism migrated progressively

77
Figure 5.12. Antalya anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model, and lateral extent is very similar with the S-wave model. Dip trend, vertical extent are very
different between models.

eastward towards the Taiwan area and is reported to have ceased at 90-86 Ma (Li et al., 2014a; c).
Subsequent transpression is associated with the ESE-ward retreat of the Pacific subduction zone
(Li et al., 2014a), which, however, corresponds to the Izu-Bonin subduction zone that started at
~52 Ma (see section 5.44). This leaves a ~30 Myr subduction gap in the region (see also Hall, 2002;
Li et al., 2012), here interpreted as the waning of the Anhui slab subduction. Future research may
require revising our interpretation of the Anhui slab.

5.8. Antalya - Ant


The Antalya anomaly (Figure 5.12) is interpreted as a N-S striking, eastward dipping slab below
the Bay of Antalya and the western Taurides of southern Turkey. It was first shown by De Boorder
et al., 1998, and later, in more detail, by Biryol et al. (2011). Seismic tomographic images clearly
show that the Antalya slab is disconnected from the Aegean-west-Anatolian slab (section 5.1) (van
Hinsbergen et al., 2010; Biryol et al., 2011) along a NE-SW trending zone that was interpreted
as a subduction transform edge propagator (STEP) fault (Govers and Wortel, 2005). Below 300-
400 km depth, the Antalya slab becomes tomographically indistinguishable from the Cyprus slab
(section 5.30), which plunges down into the top of the lower mantle.
Subduction in southern Anatolia occurred below oceanic lithosphere, preserved in a wide belt of
ophiolites. Metamorphic sole cooling ages of these ophiolites, widely regarded as forming within
several million years after subduction initiation (Hacker, 1990; Wakabayashi and Dilek, 2000;
van Hinsbergen et al., 2015), are ~92±3 Ma (Çelik et al., 2006), suggesting subduction initiation
several million years earlier, ~95-100 Ma, which we adopt for the age of the base of the Antalya and
Cyprus slabs.
Even though the Antalya slab is associated with a Benioff zone (Kalyoncuoğlu et al., 2011), it
is not evident if and where the Antalya slab is connected to a subduction thrust at the surface.

78
Figure 5.13. Arabia anomalies, Legend same as figure 5.5. Relative amplitude strength and dip
trend is similar between tomographic models. Lateral and vertical extent differ.

Biryol et al. (2011) and Schildgen et al. (2012) interpreted the Antalya slab to be a part of the
Cyprus slab that tore off and rotated into a N-S orientation. Recently, however, Koç et al. (2016)
argued that subduction of the Antalya slab accommodated E-W shortening in the heart of the
Isparta triangle until at least Pliocene time, and created E-W overriding plate extension in southern
Turkey since at least middle Miocene time. Based on a paleomagnetic study of a Central Anatolian
volcanic arc, Lefebvre et al. (2013) showed that a N-S trending subduction zone must have already
existed since late Cretaceous time and Advokaat et al. (2014a) and van Hinsbergen et al. (2016a)
argued that African plate subduction in Anatolia was accommodated along discrete E-W and N-S
trending segments since the onset of subduction. The Antalya and Cyprus slabs may thus have
been two independent slabs since the onset of subduction, but their close vicinity makes them
tomographically indistinguishable at depth.

5.9. Arabia - Ar
The Arabia anomaly belt (Figure 5.13) is located in the mid-mantle from the northern part of
the Red Sea to the southeast below the southeastern coast of Saudi Arabia. The Arabia anomaly
belt was previously identified by Hafkenscheid et al. (2006) and identified as the Eg and SA
slabs. They are located south of and at a depth interval overlapping with, and hence probably
subducted in part simultaneously with and south of the Mesopotamia slab (section 5.60) to the
NE that subducted along the Eurasian margin between 150-65 Ma. The Arabia anomalies thus
likely represent Neotethyan lithosphere that subducted intra-oceanically. Intra-oceanic subduction
within the Neotethys between Arabia and Eurasia is a well-known event in the Late Cretaceous,
and culminated in the obduction of ophiolites over the Arabian margin (the Semail ophiolite of
Oman, and the Baer Bassit and Hatay ophiolites of Syria and SE Turkey, respectively, and likely
the Troodos ophiolite of Cyprus, in the Late Cretaceous (70±5 Ma; Koop and Stoneley, 1982;
Al-Riyami et al., 2002; Dilek and Furnes, 2009; Homke et al., 2009; Searle and Cox, 2009; Agard

79
Figure 5.14. Arafura anomaly. Legend same as figure 5.5. Vertical and lateral location and dip trend
are very similar between tomographic models. Relative amplitude strength is weaker in the used
S-wave models.

et al., 2011; Maffione et al., 2017). These foreland basin deposits and overlap assemblages provide
the likely age for break-off of the Arabia slab. The minimum age of onset of subduction below
the Arabian ophiolites is provided by the age of their metamorphic soles (95±5 Ma: Hacker, 1994;
Hacker et al., 1996; Al-Riyami et al., 2002; Warren et al., 2003; Agard et al., 2007).

5.10. Arafura - Af
The Arafura anomaly (Figure 5.14) corresponds to anomaly A6 of (Hall and Spakman, 2002; 2004).
It is interpreted as a NNW-SSE trending slab, flat-lying at the top of the lower mantle and the
base of the upper mantle from north of the Bird’s Head, beneath the Arafura Sea, to the Gulf of
Carpentaria. Towards the north, the Arafura slab is in the vicinity of the base of the Halmahera
slab that started subduction around 15 Ma (section 5.36), suggesting that the Arafura slab should
be correlated to a subduction zone that terminated east of Halmahera in mid-Late Cenozoic
time. Hall and Spakman (2002; 2004) interpreted the slab as the result of northward subduction
underneath the Philippines-Halmahera arc between 45 and 25 Ma. The geological record of this
Paleogene subduction zone is, according to the reconstruction of Hall (2002), now found on the
eastern Philippines. Wu et al. (2016) interpreted the anomaly to be an East Asian Sea south slab,
being the result of southward subduction underneath the Solomon Sea between 50-20 Ma. Both
interpretations are consistent with the modern position of the slab and we therefore adopt 50-45
Ma and 25-20 Ma as age ranges for the base and top of the Arafura slab.

80
Figure 5.15. Atlantis anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with a similar Relative amplitude strength, vertical and
lateral extent, flat on top of the CMB.

5.11. Atlantis - At
The Atlantis anomaly (Figure 5.15) is located below the central Atlantic Ocean, covers the core-
mantle boundary, and was first identified in van der Meer et al. (2010). The Cocos and Hatteras
slabs (section 5.29, 3.37) shallower in the mantle to the west leads us to interpret the Atlantis
anomaly as a slab with a pre-Jurassic subduction age. Several correlations to the geological record
were discussed in van der Meer et al. (2010), with a preferred interpretation that the Atlantis slab
represents Panthalassa oceanic lithosphere that subducted along the western margin of Laurasia
until the middle Triassic. Recently Hadlari et al. (2017) with a a synthesis of U-Pb detrital zircon
data corroborated a Triassic subduction zone and continental magmatic arc system along western
Laurentia. The disconnection of the Atlantis and Cocos and Hatteras slabs results from a pause in
subduction along western Laurentia illustrated by a gap in magmatism at the western continental
margin of Laurentia during the mid-Late Triassic (Ward, 1995; Nokleberg et al., 2000; Barboza-
Gudiño et al., 2008; DeCelles et al., 2009).

81
Figure 5.16. Balkan anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with a similar vertical extent, flat on top of the CMB.
Relative amplitude strength, lateral extent differ between tomographic models. In the S-wave
model this anomaly is much larger and merges with the Al Jawf anomaly to the SE.

5.12. Balkan - Ba
The Balkan anomaly (Figure 5.16) is located below south-eastern Europe and overlies the core-
mantle boundary. It was first identified in van der Meer et al. (2010). The East Vardar slab
(section 5.32), Mesopotamia slab (section 5.60) to the northeast, both higher in the mantle and
correlated to geological records of Jurassic and younger subduction, were used to infer to interpret
the anomaly as the Balkan slab representing lithosphere that subducted prior to the Jurassic age.
The Mongol-Okhotsk subduction in the east constrains the Siberia block to a Triassic to Jurassic
position northeast of the Balkan slab. The Balkan slab therefore is likely the result of subduction
to the west of Siberia. The Al Jawf slab (section 5.5), interpreted as subducted Paleotethyan
lithosphere, suggests a subduction location of the Balkan slab northwest of southern Laurasia. Van
der Meer et al. (2010) considered the Solonker Ocean that subducted between the North China
block and Amuria until Triassic time (Xiao et al., 2010), and the Uralian ocean that intervened
Siberia and Baltica as potential sources of this slab. The position of this slab in the light of global
plate reconstructions makes a correlation to the Uralian Ocean more likely. The collision between
Siberia and Baltica occurred in the Late Permian (~280-250 Ma) (Cocks and Torsvik, 2007; 2011;
Torsvik et al., 2012), which we adopt as the age for the top of the Balkan slab.
In the S40RTS model the Balkan slab is part of a much larger NW-SE trending anomaly of >4000
km, from the North Sea to Anatolia, where it connects to the Al Jawf slab. This large extent may
represent a more complete record of the Permian Uralian Ocean, but in the UU-P07 model only
the Balkan slab (<2000km) is detected.

82
Figure 5.17. Banda anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model, but poorly imaged in the S-wave model. Dip trend, lateral and vertical extent are very
different between models.

5.13. Banda - Bn
The Banda anomaly (Figure 5.17) is interpreted as an amphitheatre-shaped, south-, west-, and
north-dipping slab that underlies the Banda Sea in southeast Indonesia and that is still, connected
to the surface in the north, east, and south. It was first shown in Fukao et al. (1992), Puspito and
Shimazaki (1993), Widyantoro and van der Hilst (1996), and Rangin et al., (1999) and topic of
reconstruction in Hafkenscheid et al. (2001), Wu et al. (2016), and particularly in Spakman and
Hall (2010). The deepest part of the Banda slab reaches the 660 km discontinuity. Spakman and
Hall (2010) showed kinematic reconstructions suggesting that subduction of the proto-Banda
embayment of the Australian continent formed the Banda slab that started around ~15 Ma by
eastward rollback. This was partly associated with delamination of the mantle lithosphere of the
Western Sula Spur directly north of the subducting embayment and with ultra-high temperature
metamorphism in crustal rocks now found on the island of Seram. This metamorphism was recently
dated at 16 Ma (Pownall et al., 2014), suggesting a somewhat earlier initiation of subduction of
the Banda embayment. A recent reconstruction of the fold-thrust belt of Timor, in the south of
the Banda embayment, shows that at least the last 8 Myr of subduction consumed ~350 km of
(continental) lithosphere (Tate et al., 2014; 2015, 2017). The lithosphere now lying between 650
and 500 km depth thus must have subducted prior to ~8 Ma and after 16 Ma. We thus use a 12±4
Ma as estimate for the start of its subduction.

83
Figure 5.18. Beaufort anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with a similar Relative amplitude strength, however
vertical and lateral extent differ.

5.14. Beaufort - Bf
The Beaufort anomaly (Figure 5.18) is located below North Alaska and the Beaufort Sea within
the deep mantle. It was first identified in van der Meer et al. (2010) and subsequently Shephard et
al. (2014) interpreted its dynamic evolution based on modelling. The base of the Mongol-Okhotsk
slab (section 5.63), as well as the base of the Idaho slab (section 5.42) to the south at similar depth,
led us to interpret the anomaly as the Beaufort slab representing lithosphere that likely subducted
during the Early or Middle Mesozoic in the Arctic. Nokleberg et al. (2000) reconstructed
subduction of Arctic/Panthalassa lithosphere at the Alazeya arc and forming the Aluchin
subduction zone terrane, during the Late Triassic-Jurassic, and was chosen by van der Meer
et al. (2010) as most likely correlation for this slab, with the time of subduction of 230-144 Ma.
Shephard et al. (2014) preferred a tectonic model suggested that the source of the Alazeya arc was
subducting Panthalassa lithosphere between 230-193 Ma, followed by a subduction polarity reversal
in the reconstruction of Nokleberg et al. (2000). A more recent geological and geochronological
study of Sokolov et al. (2009) described the Alazeya arc and several of its terranes (Yarakvaam,
Vurguveem, and Aluchin ophiolites) in more detail and found ages of amphibolites associated with
ophiolites (239.1±8 Ma) or in tectonic blocks in sub-ophiolitic melanges (229–257 Ma). Sokolov
et al. (2009) interpreted this to reflect the age range of the entire subduction history below the
Aluchin ophiolite. Amphibolites associated with ophiolites are, however, frequently related to the
early stages of subduction (metamorphic soles), after which temperatures in the subduction zones
decrease and blueschists are formed instead. Sokolov et al. (2009) does not describe the structural
context of these amphibolites within the ophiolite structure in detail, but we conservatively estimate
the 257-229 Ma age range to reflect the age of the base of the Beaufort slab, reasonably in line with
the estimates of Nokleberg et al. (2000), and adopt the 193 Ma age preferred by Shephard et al.
(2014) as the end of subduction.

84
Figure 5.19. Bering Sea anomaly. Legend same as figure 5.5. This slab is well imaged in the
UU-P07 model, and lateral extent is very similar with the S-wave model. Vertical extent is different
between models.

5.15. Bering Sea - BS


The Bering Sea anomaly (Figure 5.19) is located below the Bering Sea, is detached, and is flat-
lying at the 660-km discontinuity. Towards the south it is close to, but separated from the Aleutian
slab (section 5.3). Towards the west, it is close to but separated from the Agattu slab. Towards the
north, it connects to the N-S trending Mayn slab (section 5.59). Gorbatov et al. (2000) suggested
that the separation between the Aleutian slab and the Bering Sea anomaly would have been caused
by the Kula-Pacific ridge subduction and estimated a ~48 Ma age for break-off. There is no clear
reason, however, why the Kula lithosphere would become located farther north than the Pacific
lithosphere to explain the separation between the Bering Sea and Aleutian slabs. The Bering Sea
contains, moreover, evidence for a short-lived Cenozoic intra-oceanic subduction zone to the north
of the Aleutian trench. This subduction formed the volcanic arc of the Bowers Ridge which is
clearly visible in the bathymetry of the western Bering Sea. Nokleberg et al. (2000). The presence
of a trench filled with as much as 12 km of Cenozoic sedimentary rocks at the base of the north
and east slopes of the Bowers Ridge suggests that the unit formed in a Cenozoic arc-trench system
that faced northeast (Nokleberg et al., 2000). We do not see any other slab that could represent
the Bowers Ridge subduction other than the Bering Sea slab. Recently Wanke et al. (2012) dated
volcanic rocks of the Bowers Ridge at Oligocene-Miocene (32.3±2 to 22.7±2.7 Ma), which we
adopt as the age range of the subduction of the Bering Sea slab. The size of the Bering Sea slab
however seems to be bigger than the Bowers basin, which may suggest that the close vicinity of
other slabs may lead to blurring of the tomographic image increasing its apparent size.

85
Figure 5.20. Bitlis anomaly. Legend same as figure 5.5. Positive anomalies are identified in the same
location in both tomographic models with a similar Relative amplitude strength, however vertical
and lateral extent differ.

5.16. Bitlis - Bi
The Bitlis anomaly (Figure 5.20) is a detached and is located below southeast Anatolia in the upper
mantle and uppermost lower mantle. It was identified by Zor (2008), as the western Lh1 anomaly,
which forms in his model a narrow band just south of the modern Bitlis suture. Other tomographic
models published in Faccenna et al. (2006), Hafkenscheid et al. (2006), Lei and Zhao (2007),
Biryol et al. (2011), and Skolbeltsyn et al. (2014), or the UU-P07 model (Figure 5.20) do not show
such detail, but consistently show an anomaly lying on the 660 km discontinuity, interpreted by
these authors as detached lithosphere (Bitlis slab). Our model, as well as Hafkenscheid et al. (2006),
suggests that the Bitlis slab penetrates the 660 km discontinuity. We interpret the Bitlis slab depth
between 440 and 920 km depth. Break-off of the Bitlis slab is widely interpreted to have occurred
at 13-10 Ma ago, which corresponds to the end of deep-marine sediment deposition in the suture
zone, a rapid phase of erosional denudation of the Bitlis massif, and a volcanic flare-up in eastern
Anatolia (Keskin, 2003; Sengör et al., 2003; Faccenna et al., 2006; Hüsing et al., 2009; Okay et al.,
2010). The onset of subduction of the Bitlis slab is more difficult to assess. In late Cretaceous time,
ophiolites were obducted onto northwestern Anatolia in southeastern Turkey, Syria, and further
west in Cyprus and southern Anatolia. These ophiolites formed since ~95 Ma as reflected by crustal
crystallization ages and cooling ages of metamorphic soles, and uppermost Cretaceous, ~70-65
Ma sediments unconformably cover the obduction thrust (Al-Riyami et al., 2002; Kaymakci et al.,
2010; Karaoğlan et al., 2012; 2014). We interpreted this obduction to reflect the end of subduction
of the western part of the Arabia slabs (section 5.10) with a top at ~1100±100 km, to the south
of the Bitlis slab. Moix et al. (2008) and Maffione et al. (2017) suggested that obduction of these
ophiolites followed upon westward radial roll-back of an east-dipping subduction zone into the
eastern Mediterranean region, which would have been accompanied by the opening of a back-arc

86
Figure 5.21. Bitterroot anomaly. Legend same as figure 5.5. Positive anomalies are identified in
similar location in both tomographic models with a similar Relative amplitude strength, however
vertical and lateral extent differ.

until ophiolite obduction. The Bitlis slab probably represents lithosphere of this back-arc, with
subduction starting at or after ~70-65 Ma, which we adopt for the age of the base of the Bitlis slab.

5.17. Bitterroot - Btr


The Bitterroot anomaly (Figure 5.21) is detached and located in the upper mantle below western
North America, and is hook-shaped at a depth of~200km. In tomographic studies it has been
imaged to depths of 230-600 km (Schmandt and Humphreys, 2011). Its distal position relative
to the trench has recently been interpreted as the result of flat-slab subduction, which may be
correlated to the geological record that shows the accretion of the Siletzia micro-continent to
North America starting at ~60 Ma (Schmandt and Humphreys, 2011). Following this accretion,
regular subduction at the Cascadia subduction zone (see Juan de Fuca slab, section 5.45) led to
arc volcanism in Oregon and Washington beginning ca. 45–40 Ma (Schmandt and Humphreys,
2011; Wells and McCaffrey, 2013). The volcanic arc associated with the Bitterroot subduction we
interpret to be the Challis-Absaroka arc, extending nearly 1000km into the present-day Laurentian
continental interior and located above the Bitteroot anomaly. The age of this arc is ~55-45 Ma
(Humphreys 2009). Recently the volcanics were dated by U-Pb geochronology by (Gaschnig et al.
2017) and we interpret the Bitterroot peraluminios suite (66-53 Ma) and Challis intrusives (~50-46
Ma) to have resulted from subduction of the slab. We therefore adopt 66 Ma as start of subduction
and 46 Ma as end of subduction leading to the formation of the Bitterroot slab.

87
Figure 5.22. Brasilia anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical,
lateral extent and dip trend are very similar between tomographic models.

5.18. Brasilia - Br
The Brasilia anomaly (Figure 5.22) is located below central South America from the deep lower
mantle to the Andean trench. It is N-S to NW-SE trending. Above 1175 km it connects to
shallower slabs, including the Venezuela slab (section 5.90) and the Malpelo slab (section 5.53) to
the north. We interpret the anomaly as a slab that resulted from Mesozoic-Present Farallon-Nazca
lithosphere subduction at the South American continental margin. The start of arc magmatism in
Peru has been recently documented to have ages ranging from ~200 Ma (Demouy et al., 2012),
180-145 Ma (Villagomez et al., 2011) to 173 Ma (Boekhout et al., 2012), which we take as the
minimum age range for the start of subduction. These studies find a gap in magmatism between
145-110 Ma (Villagomez et al., 2011; Boekhout et al., 2012; Demouy et al., 2012), followed by
volcanism into the Cenozoic (Demouy et al., 2012; Jones et al., 2014). In the tomographic model
UU-P07 there is not a noticeable gap visible in the slab, however the anomaly considerably weakens
at 2100 km. In S40RTS, the anomaly appears to have a slab window at some locations at 2200 km,
but the anomaly appears to be otherwise contiguous in the south. We do not find evidence from
the tomographic models to spatially separate a deeper Jurassic slab from a Cretaceous-Cenozoic
slab and therefore treat the Brasilia slab as one single slab that subducted since at least 200-173
Ma. The SL2013 model does show a significant positive anomaly on top of the transition zone, up
to 2000km east of the trench (Figure 5.22F). Another flat-lyuing anomaly in the transition zone
and shallow lower mantle is visible in the tomographic model of Li et al. (2008) and modelled
by Facenna et al. 2017. In their interpretation it has from subduction possibly as old as 100 Ma.
These shallower anomalies are not expressed in the UU-P07 model and we therefore suggest that
the tomographic imaging of these anomalies and subduction evolution at depth require further
investigation.

88
Figure 5.23. Burma anomaly. Legend same as figure 5.5. Relative amplitude strength and dip trend
are similar between tomographic models. Vertical and lateral extent differ.

5.19. Burma - Bur


A N-S striking, steeply east-dipping anomaly has been described below the West Burma block
(Figure 5.23) by Huang and Zhao (2006), Li et al. (2008), Zhao and Ohtani (2009), Pesicek
et al. (2010), Replumaz et al. (2010a), and Koulakov (2011). This was interpreted as a slab that
is disconnected in the upper few hundred kilometres of the mantle from the Sunda slab below
Sumatra (section 5.85) through a slab window below the Andaman Islands. The deepest parts of
the Burma slab reach ~710 km depth (Figure 5.24). The eastward dip of the Burma slab led Li et
al. (2008) to interpret the Burma slab as Indian plate lithosphere that subducted below the Burma
block. The Burma block in Neogene time has acted as a forearc sliver that moved northward relative
to the adjacent Indochina block, bounded in the west by a fold-thrust system (along which the
Burma slab subducted), and in the east by a transform fault zone (Morley, 2002; Bertrand and
Rangin, 2003; Maurin and Rangin, 2009). The West Burma block contains Paleozoic and older
metasediments which tectonically overlies an ophiolite with upper Cretaceous oceanic crust that
in turn was obducted onto India (Acharya and Mitra, 1986). The ophiolite overlies a foreland basin
with Cretaceous to Upper Eocene to Lower Oligocene clastic sediments, and is unconformably
covered by shallow marine to fluvial Upper Eocene sediments (Ghose et al., 2010), constraining
the final obduction onto India to the Late Eocene-Early Oligocene, ~35±5 Ma. By that time,
northeastern India was in a position adjacent to, and underwent highly oblique convergence with
the West Burma block. A kinematic restoration of the oblique Burma subduction zone of van
Hinsbergen et al. (2011a) suggests that ~600-700 km of E-W convergence occurred between India
and the Burma Block since ~30-40 Ma, alongside ~1100 km of northward motion of the Burma
block relative to Indochina. This 600-700 km corresponds to the length of the Burma slab and we
consequently interpret the base of this slab to have started subduction around 35±5 Ma.

89
Figure 5.24. Calabria anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model, but poorly imaged in the S-wave model. Vertical extent is similar, but dip trend and lateral
extent are quite different between models.

5.20. Calabria - Cl
The Calabria slab (Figure 5.24) was first documented by Spakman, (1986a; 1993), and has since
been a prominent feature in tomographic models (e.g. Piromallo and Morelli, 1997; 2003; Spakman
and Wortel, 2004; Chang et al., 2010). The Calabria slab is lying flat at the base of the upper
mantle and that has not penetrated the lower mantle yet. It is still subducting along the Calabrian
subduction zone. Estimates for the onset of northward subduction in the western Mediterranean
vary from ~80 to 35 Ma (Rosenbaum et al., 2002b; Faccenna et al., 2004). A recent kinematic
reconstruction of the western Mediterranean region of van Hinsbergen et al. (2014) showed that
Africa-Iberia convergence, as estimated from the Atlantic ocean spreading reconstructions and
corrected for intra-Iberian and intra-African shortening, started around 85 Ma, and was very
slow until 45 Ma (<100 km), after which subduction accelerated. This led to inception of arc
volcanism in Sardinia and in the Provence around 38 Ma (Lustrino et al., 2009), and was followed
by the inception of roll-back and back-arc extension around 30 Ma (Faccenna et al., 2001; 2004;
Rosenbaum et al., 2002b; Rosenbaum and Lister, 2004; van Hinsbergen et al., 2014). Because it
is uncertain when the slow convergence culminated in the inception of Calabrian subduction, we
adopt a wide time range for the base of the Calabrian slab of 65±20 Ma.

90
Figure 5.25. Caribbean anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model, but poorly imaged in the S-wave model. Dip trend, lateral and vertical extent are quite
different between models.

5.21. Caribbean - Ca
The Caribbean anomaly (Figure 5.25) is located in the upper mantle below the eastern Caribbean
Sea rising to the surface at the Lesser Antilles subduction zone where Central Atlantic oceanic
lithosphere subducts westward. The anomaly is interpreted as a slab dipping steeply westwards from
the surface to a depth of 600 km, changing to a sub-horizontal position below this depth (van der
Hilst and Spakman, 1989; van Benthem et al., 2013) . The base of the slab is at least 45 Ma old, as
shown by the ages of onset of the Lesser Antilles arc (Burke, 1988), and the Paleocene Grenada
and Tobago basins may have formed shortly after initiation of subduction along the Lesser Antilles
subduction zone (Boschman et al., 2014). The Lesser Antilles subduction zone likely formed by
inversion of a STEP fault that accommodated the northward retreat of the Hispaniola slab (section
5.41) after that slab decoupled from the Venezuela slab (section 5.90) (see also van Benthem et
al., 2013). This may explain why the Caribbean slab is not physically connected – at least in its
central portion – to these two deeper slabs (Boschman et al., 2014). Based on the reconstruction of
Boschman et al (2014) and the onset age of the Lesser Antilles arc, we adopt a 50±5 Ma age for the
base of the slab

91
Figure 5.26. Carlsberg anomaly. Legend same as figure 5.5. Vertical, lateral extent and dip trend
are very similar between tomographic models. Relative amplitude strength is weaker in the S-wave
model.

5.22. Carlsberg - Cr
The Carlsberg anomaly (Figure 5.26) is detached and NNE-SSW trending below the northwestern
Indian Ocean and was identified by Gaina et al. (2015) in the top of the lower mantle between 800
and 1400 km depth. The location and depth of the top suggest that the slab should be correlated to
a geological record of past subduction in Late Mesozoic or Early Cenozoic time to the southwest
of the Himalaya. Gaina et al. (2015) interpreted the anomaly as the Carlsberg slab that subducted
as a result of highly oblique subduction of Indian lithosphere below oceanic lithosphere of the
African (Arabian) plate in the Late Cretaceous to Paleocene. Initiation and arrest of its subduction
is reflected in the geological record by formation and obduction of the Bela and Muslim Bagh
ophiolites in Pakistan and the Kabul-Altimur ophiolites on the Kabul block in Afghanistan (Gnos
et al., 1997; Gaina et al., 2015). The oldest radiometric age of supra-subduction ophiolites in
western Pakistan that were interpreted to have formed above this subduction zone is an 80.2±1.5
Ma U/Pb age from a plagiogranite dyke (Kakar et al., 2012), whereas metamorphic soles below
these ophiolites have 65-70 Ma ages (Mahmood et al., 1995; Gnos et al., 1998). We follow the
interpretation of the Carlsberg slab of Gaina et al. (2015) and adopt a 73±8 Ma age range for the
subduction of its base. The end of subduction is reflected by Eocene foreland basin deposits on the
Indian and Arabian margin and overlap assemblages sealing the obduction thrust of ~55-45 Ma
(Gnos et al., 1997; Khan and Clyde, 2013; Gaina et al., 2015).

92
Figure 5.27. Caroline Ridge anomaly. Legend same as figure 5.5. Positive anomalies are identified
in the same location in both tomographic models with a weaker Relative amplitude in the S-wave
models. Vertical and lateral extent differ.

5.23. Caroline Ridge - CR


The Caroline Ridge anomaly (Figure 5.27) was identified by Hall and Spakman (2002; 2004) and
is draped on the 660 km discontinuity below the Caroline Ridge, north of Papua New Guinea. It
lies south of the Mariana slab (section 5.57), but according to the reconstruction of Hall (2002)
is not directly related to it. (Hall, 2002) reconstructed an intra-oceanic subduction zone that
accommodated Pacific plate subduction at the location of the Caroline Ridge anomaly initiating
at ~25 Ma along a transform fault that accommodated clockwise rotation of an intra-oceanic
subduction zone that rolled back southward towards Papua New Guinea, opening the Caroline
Sea plate lithosphere in its back-arc position. This subduction zone came to an arrest at ~5 Ma
according to Hall (2002). Wu et al. (2016) named this anomaly the New Guinea Offshore slab
and suggested an age range of 25-10 Ma. We use the largest age range of these interpretations and
adopt 25 and 5 Ma as the ages for onset and end of subduction of the Caroline Sea plate forming
the Caroline Ridge slab.

93
Figure 5.28. Carpathian (north) anomaly. Legend same as figure 5.5. This detached part of the
slab is well imaged in the UU-P07 model, but less well delineated in the S-wave model. Relative
amplitude strength and lateral extent are similar between models. There is no dip trend visible, and
the slab appears to flat-lying on top of the transition zone.

5.24. Carpathians - Cp
The Carpathian anomaly (Figure 5.28) is found draped on the 660 km discontinuity below the
Carpathian oroclinal fold-thrust belt and its back-arc region, the Pannonian basin. It has been
long known and is imaged in numerous tomographic models (Spakman 1991; Spakman et al 1993;
Bijwaard et al., 1998; Wortel and Spakman, 2000; Piromallo and Morelli, 2003; Ren et al., 2012;
Zhu et al., 2012). Apart from the southeastern corner of the Carpathian orocline – the Vrancea
area, where a slab and associated Benioff zone are still imaged to be connected to the European
lithosphere – the Carpathian anomaly is disconnected from the surface.
The anomaly is interpreted as the Carpathian slab that resulted from westward subduction of
Eurasian lithosphere below several terranes (Tisza-Datca and AlCaPa blocks) that became
separated from Eurasia during Jurassic opening of the Piemonte-Ligurian Ocean (Schmid et al.,
2008; Vissers et al., 2013) and that were deformed in a Cretaceous orogenesis unrelated to, and
prior to the westward Carpathian subduction history (Csontos and Voros, 2004; Schmid et al.,
2008). This westward subduction led to the formation of a thin-skinned fold-thrust belt – the outer
Carpathians. The onset of this deformation occurred around 35±5 Ma (Matenco and Bertotti,
2000; Matenco et al., 2003; Gągała et al., 2012; Handy et al., 2014), and became associated with
back-arc extension since ~20 Ma, opening the Pannonian basin (Horvath et al., 2006; Matenco
and Radivojević, 2012), followed by arrest of shortening and unconformable covering of the
frontal thrust system around 10-12 Ma in the northeastern Carpathians (Matenco and Bertotti,
2000; Gągała et al., 2012). We adopt the 11±1 Ma as the age of break-off of the Carpathian
slab. Finally, we note that the northern part of Carpathians hosted by the AlCaPa terrane, north
of the Periadriatic-Ballaton fault, underwent several hundreds of kilometres more subduction
and extension in the Miocene than the area hosted by the Tisza-Datca terranes to the south

94
Figure 5.29. Carpentaria anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with a weaker Relative amplitude in the S-wave
models. Vertical and lateral extent differ.

(Ustaszewski et al., 2008). This suggests that the Carpathian anomaly is in fact composed of two
slabs, separated by a transform fault, whereby the southern slab may still be connected to the surface
in the Vrancea area (Figure 5.30). Our tomographic image, as well as that of Ren et al. (2012)
appears to show such a disconnection between the two bodies below the Ballaton line.

5.25. Carpentaria - Cn
The Carpentaria anomaly (Figure 5.29) corresponds to anomaly A8 of Hall and Spakman, (2002;
2004). It lies in the upper part of the lower mantle below the Gulf of Carpentaria below northern
Australia and Papua New Guinea. It is located deeper in the mantle than the Arafura and Papua
slabs that lie to the west and east respectively, and which likely subducted since ~45 Ma (sections
3.70 and 3.10, respectively). This led Hall and Spakman (2002; 2004) to interpret the Carpentaria
anomaly as a slab with a pre-middle Eocene age of subduction. They consequently suggested that
the Carpentaria slab may represent Australian plate lithosphere that subducted westward prior to
the polarity switch associated with the onset of subduction of below the Melanesian arc around
45 Ma (see also Gurnis et al., 2000). Hall and Spakman (2002; 2004) suggested a Cretaceous age
for the onset of subduction. Based on plate reconstructions of the SW Pacific and southern Pacific
ocean basins, Seton et al. (2012) showed that convergence between the Pacific and SW Pacific
basins must have occurred since at least 83 Ma, which we adopt as the minimum age for the onset
of the Carpentaria slab. Wu et al. (2016) interpreted the anomaly to be an East Asian Sea south
slab, being the result of southward subduction underneath Papua New Guinea and Solomon Sea
between 50-20 Ma. We therefore use an age range of 83-50 and 45-20 Ma for the base and top of
the slabs.

95
Figure 5.30. Caucasus West anomaly. Legend same as figure 5.5. Relative amplitude strength,
vertical and dip trend are similar between tomographic models. Lateral extent differs.

5.26. Caucasus - Cau


Tomographic images and earthquake locations have shown a complex mantle structure below
the Caucasus (Figure 5.30). The Eastern Caucasus is underlain by a northward dipping zone of
seismicity with earthquakes reaching more than 150 km depth, which was interpreted to bound
a 130-280 km long, northward dipping slab (Maggi and Priestley, 2005; Skolbeltsyn et al., 2014;
Mumladze et al., 2015) (Figure 33). The western Caucasus, however, does not display seismicity
deeper than 50 km (Mumladze et al., 2015). Instead, seismic tomographic images show an anomaly
between ~650 and 350 km depth that may represent a body of lithosphere that delaminated from
the base of the western Caucasus. This was first noted in Brunet et al. (2000b), and shown for the
first time in Hafkenscheid et al. (2006). Subsequently, seismic tomographic images of Zor (2008)
and Koulakov et al. (2012) confirmed the location and size of this body.
The onset of shortening and uplift in the Caucasus, as a proxy for the age of subduction of the
base of the Caucasus anomalies, may be estimated by a first phase of denudation in Oligocene
that is detected by low-temperature thermochronology (Vincent et al., 2007; 2010; Cowgill et al.,
2016), at 30±5 Ma, whereby Vincent et al. (2017) showed that the older end of this age range is
probably most realistic. Break-off of the anomaly below the western Caucasus may coincide with
the rapid phase of uplift and crustal shortening that started around 5 Ma (Forte et al., 2010; 2014;
Avdeev and Niemi, 2011). Because the slab is much narrower than the elongated E-W subduction
zone that existed until at least Early Cenozoic time along the southern Pontides of Turkey (e.g.,
Meijers et al., 2010), we consider it unlikely that the narrow western Caucasus slab is related to this
subduction zone. We have not interpreted a slab of this Pontide subduction zone, but suspect that it
is located within the wide anomaly below Turkey of which the southern parts are interpreted as the
Cyprus (section 5.33), Bitlis (section 5.17), and Antalya slabs (section 5.9).

96
Figure 5.31. Central China anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with an overall stronger Relative amplitude in the
S-wave models. Vertical and lateral extent differ.

5.27. Central China - CC


The Central China anomaly (Figure 5.31) is located in the lower mantle from the core-mantle
boundary upwards to a depth of ~1500 km. It was first defined by van der Meer et al. (2010), who
separated the slab from the Mongol-Okhotsk slab (section 5.67) as defined by Van der Voo et al.
(1999a). Van der Meer et al. (2010) interpreted the Central China slab as representing Paleotethyan
lithosphere that subducted between North China, northeast Tibet and Eurasia following
interpretations of Stampfli and Borel (2004), assigning a Permian to Early Cretaceous age to this
slab. Recently, however, Van der Voo et al. (2015) proposed that the Mongol-Okhotsk subduction
zone started as a westward dipping subduction zone below Siberia in the north and Amuria in
the south, whereby the northern, currently WNE-WSW trending margin of Amuria along the
Mongol-Okhotsk suture was oriented N-S until Triassic time and rotated counterclockwise
towards its current orientation throughout the Late Triassic and Jurassic, culminating in an
orocline around the western termination of the Mongol-Okhotsk suture in western Mongolia. In
that interpretation, the Central China slab subducted since at least Triassic times until the latest
Jurassic below Amuria, and form the southward continuation of the Mongol-Okhotsk slab. The
paleomagnetic data summarized in Van der Voo et al. (2015) in addition imply that the North
China block moved 1000’s of km northward relative to Siberia-Kazakhstan until the latest
Jurassic-earliest Cretaceous, requiring that the North China-Kazakhstan plate boundary should
accommodate significant (oblique) convergence. Although the location of such a plate boundary
remains enigmatic, a ‘Paleotethyan’ origin as implied in Stampfli and Borel (2004) is not excluded.
If related to the advance of NE Tibet to Eurasia, the Central China slab subducted since at least
Triassic and probably earlier time, and ended around the Jurassic-Cretaceous boundary 140±10
Ma. An alternative interpretation would be that the Central China slab resulted from subduction
along the Songpan Ganzi suture between the Qiangtang and Kunlun terranes. This subduction also

97
Figure 5.32. Chukchi anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with similar dip trend, lateral and vertical extent and
Relative amplitude strength.

started prior to the Triassic, and ended in Late Triassic or Early Jurassic time, ~200 Ma ago (e.g.,
Pullen and Kapp, 2014). In the recent reconstruction of Van der Voo et al. (2015) the reconstructed
position of the Songpan-Ganzi suture would however be ~15° south of the slab and is therefore
considered an unlikely candidate associated with the subduction of the slab.

5.28. Chukchi - Ch
The W-E trending Chukchi anomaly (Figure 5.32) is located below North Siberia and the Chukchi
Sea (Arctic Ocean) within the mid-mantle. At its base, at equivalent latitudes it is located in
between the top of the Mongol-Okhotsk slab (section 5.63) towards the west and the Hudson slab
(section 5.41) towards the east. It was first identified in van der Meer et al. (2010), who interpreted
it as the Chukchi slab representing paleo-Arctic lithosphere, subducted below the Upper Jurassic
to Lower Cretaceous Koyukuk arc in a (near-) continental margin position. However, in the
subsequent interpretation van der Meer et al (2012), we considered it more likely that the Chukchi
slab resulted from northward Panthalassa Ocean subduction. This subduction may correlate with
the Kony-Murgal arc Nokleberg et al. (2000), in which the age of was poorly constrained between
Triassic and Early Cretaceous. Sokolov (2010) pointed out that the Middle Jurassic of the Kony-
Murgal arc is characterized by structural rearrangements and deformation coinciding in time with
the onset of the formation of the system of Pacific plates controlling the present-day appearance of
the Pacific Ocean. This results in the start of the Late Jurassic–Early Cretaceous Uda–Murgal arc
in the region of the Chukchi slab. Noteworthy is that the slab at its eastern extent has a N-S trend
both in the UU-P07 and S40RTS model, suggesting a change in subduction zone orientation. This
change would fit with the eastward extent of the Uda-Murgal arc into the Pekulney and Oloy arcs
during the Late Jurassic -Early Cretaceous (Nokleberg et al., 2000) and subsequently modelled by
Shephard et al. (2013) between 160-108 Ma. According to Sokolov et al. (2009), the accretionary

98
Figure 5.33. Cocos anomaly. Legend same as figure 5.5. Vertical, lateral extent and dip trend are
similar between tomographic models. Relative amplitude strength is weaker in the S-wave model.
The continuation in the transition zone is less well resolved in the SL2013 model.

prism (Beregovoi terrane) started forming in the Middle Jurassic, and the volcanic arc (central
Taigonos terrane) ended in the Barremian-Albian. We therefore adopt a 174.1-163.5 Ma for the
start of subduction and an end of 129.4-100.5 Ma for the end of subduction.

5.29. Cocos - Co
The Cocos anomaly (Figure 5.33) is part of the set of anomalies commonly, and in our previous
compilation in van der Meer et al. (2010), referred to as the Farallon slab, as originally defined by
Grand et al. (1997) and also imaged by e.g. Bijwaard et al. (1998), Fukao and Obayashi (2013)
and Sigloch and Mihalynuk (2013). The anomaly is interpreted as the Cocos slab and is NW-SE
trending, and dips from the Central American trench, where it connects to the subducting
Cocos plate, down to the lower mantle below the northern Caribbean region and eastern North
America. Rogers et al. (2002) showed tomographic images of the Cocos slab below Central
America and suggested that it may have detached in recent times at a depth of ~200-300 km, but
our tomographic model UU-P07 does not reveal this gap. Recently detached or not, the Cocos
slab is the longest contiguous anomaly in our database. It can be distinguished from the Hatteras
slab (section 5.37), although it may connect with this slab at mid-mantle depths, where it locally
assumes a N-S strike. Above 1100 km depth the NW-SE strike reappears, all the way up towards
the Central American trench.
The location of the deepest part of the anomaly corresponds to the slabs that (Sigloch and
Mihalynuk, 2013) interpreted as resulting from Franciscan intra-oceanic subduction, forming
their SF1 slab, which resides at shallower depth. We adopt their interpretation, and extend the SF1
slab into the deepest mantle. The Cocos slab is the only still-subducting segment of the once vast
Farallon oceanic plate and is currently subducting the Cocos plate below Mexico. Correlating the
Cocos slab to the Franciscan melange predicts that North America moved northward relative to the

99
Figure 5.34. Cyprus anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical and
lateral extent are similar between tomographic models. Dip trend is better visible in the UU-P07
model.

Cocos slab during its subduction, which may be tested in future studies. Although there is debate
when the ‘Franciscan’ phase of subduction initiated, most agree that subduction was underway by
170-160 Ma, the time of formation of high temperature – high-pressure metamorphic rocks in the
structurally higher levels of the Franciscan complex (Anczkiewicz et al., 2004; Wakabayashi and
Dumitru, 2007; Hoisch et al., 2014; Wakabayashi, 2015), which we adopt as the minimum age of
the start of subduction of the Cocos slab.

5.30. Cyprus - Cy
The Cyprus anomaly (Figure 5.34) was imaged by Faccenna et al. (2006) and Biryol et al. (2011)
and is interpreted as a north-dipping slab that may or may not be still attached to the African
lithosphere to the south of the island of Cyprus. In the upper several hundred kilometres of the
mantle, it is clearly disconnected from the Antalya slab (section 5.8), but in the lower part of the
upper mantle, the two slabs become indistinguishable (Biryol et al., 2011). We do therefore not add
a separate estimate for the age of the base of the Cyprus slab, but apply the estimate for the base
of the Antalya slab of 95±5 Ma. The Cyprus slab consequently likely contains oceanic lithosphere
of the Izmir-Ankara ocean, continental lithosphere of the Taurides, and oceanic lithosphere that
subducted after the Late Cretaceous-Eocene accretion of the Taurides. This oceanic lithosphere
must have connected to the Troodos ophiolite on Cyprus that emplaced in the late Cretaceous onto
continental rocks exposed in the Kyrenia range following roll-back of an original East-dipping
subduction zone into the SE Mediterranean region (Moix et al., 2008; Maffione et al., 2017),
whereby the Cyprus (Troodos) and Syria (Baer Bassit) ophiolites underwent large counterclockwise
rotations (Morris et al., 2002). The age of obduction is best constrained by Maastrichtian sediments
that unconformably cover the contact of the ophiolite with its underlying melange that contains
Triassic passive margin carbonates (Swarbrick and Naylor, 1980; Bailey et al., 2000). Similar to the

100
Figure 5.35. East China anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with similar, lateral and vertical extent and Relative
amplitude strength. Dip trend is not visible. The transition to other anomalies further west is quite
different between tomographic models, and these anomalies are stronger in the S40RTS model.

Bitlis slab (section 5.16), we suggest that the much of the Cyprus slab consists of this back-arc that
formed during late Cretaceous roll-back.

5.31. East China - EC


The East China anomaly (Figure 5.35) is located below eastern Asia from the core mantle boundary
up into the deep mantle (~2000 km depth) and east of the Mongol-Okhotsk slab (section 5.63). At
its shallowest point (~1700 km), it connects to Mongolia slab (section 5.62) to the NW. Due to the
thickness of the anomaly slab (>1000km) we tentatively suggest it may comprise multiple SW-NE
trending slabs. Van der Voo et al. (1999a) identified this anomaly as a Pacific slab, which formed
due to westward subduction of paleo-Pacific (Panthalassa) lithosphere at the east-Asian margin
during Mongol-Okhotsk subduction. This suggests that we best correlation the East China slab to
a Triassic-Jurassic geological record east of Mongolia. Previously, we correlated the anomaly with
the Triassic to Late Jurassic evolution of the Korean peninsula, the North China block, and Japan
(van der Meer et al., 2010). Li and Li (2007) described how subduction that started in the Triassic
culminated during the Early Jurassic in a flat slab below the South China craton. Foundering of
this slab took place between 180-155 Ma. However, Li et al. (2014a) showed that magmatism in
South China occurred over a zone of ~600 km wide between 145-118 Ma, which they inferred
to be caused by the migration of a subducting ridge. This ridge subduction may be used as timing
of slab break-off which we adopt here, although future study into the kinematic history of the
Panthalassa-North China subduction history may require altering this interpretation.

101
Figure 5.36. East Vardar anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical,
and dip trend are very similar between tomographic models. Lateral extent differs.

5.32. East Vardar - Evd


The East Vardar anomaly (Figure 5.36) is located in the lower mantle slightly south of, and
immediately below the Aegean slab, between 1500 and 2000 km. The name derives from our
preferred paleogeographical interpretation of this slab rather than its modern geographical
position, since it directly underlies the Aegean slab. Previously, van der Meer et al. (2010) followed
the interpretation of van Hinsbergen et al. (2005) that the East Vardar slab represents Jurassic to
Early Cretaceous intra-oceanic (north)eastward dipping subduction within the western part of the
Neotethys culminating in ophiolite obduction of the Hellenic and Albanides ophiolites (e.g. Othris,
Pindos and Mirdita ophiolites) in the Early Cretaceous (~130 Ma ago) – now known as the ‘west-
Vardar ophiolites’ of Schmid et al. (2008). Stampfli and Borel (2004) and Schmid et al. (2008),
and more recently Jahn-Awe et al. (2010) and Natal'in et al. (2012), however, pointed out that two
Jurassic-Early Cretaceous intra-oceanic subduction zones existed in the Vardar ocean, not only
emplacing the west-Vardar, but also the East-Vardar ophiolites. The latter were thrusted onto the
fringes of the Moesian platform (i.e. onto the circum-Rhodope units and Biga peninsula) It is more
likely that the slab present below, and slightly south of the Aegean slab is associated with southward
subduction below the east Vardar ophiolites instead – hence we deem it the ‘East-Vardar slab’. The
age of onset of subduction is similar to that previously inferred, and is occurred ~170 Ma ago as
shown by the oldest ages of the supra-subduction zone ophiolitic crust of the northeast Aegean
ophiolites (e.g., Guevgueli, Samothraki, and Sithonia ophiolites (Magganas et al., 1991; Robertson,
2002; Koglin et al., 2009). The age of the top of the slab is likely somewhat older than the age of
the base of the Aegean slab, for which we inferred an age of 110±10 Ma (section 5.1). The west-
Vardar slab should be found farther to the west (here interpreted as the Algeria slab, section 5.4).

102
Figure 5.37. Georgia Islands anomaly. Legend same as figure 5.5. Relative amplitude strength,
vertical, and lateral extent are very similar between tomographic models. Lateral extent differs. The
slab appears to be a pile lying on top of the CMB.

5.33. Georgia Islands - GI


The Georgia Island anomaly (Figure 5.37) is located below the southern Atlantic and Antarctic
Oceans from the core-mantle boundary up into the mid-mantle, separated from the shallower
Scotia slab (section 5.79). It is interpreted as a slab that appears to be flattened on the core-
mantle boundary and has a W-E trend at shallower depths. It was first identified by van der
Meer et al. (2010), who inferred an Early-Middle Mesozoic age of subduction. On trend with
other deep slabs (i.e. Cocos (section 5.29) and Brasilia (section 5.18) that subducted below the
Americas in the Mesozoic, the Georgia Islands slab probably subducted farther south along the
southwestern Gondwana margin. The magmatic arc history of southern South America shows a
gap in magmatism between the Early Triassic and the Jurassic, after which subduction re-initiated
(Martin, 2007). The Jurassic and younger subduction history we associate with the more westerly
located South Orkney Island slab in the mid-mantle (section 5.84).
Pre-Jurassic subduction in southern South America started at least in Carboniferous time
(Pankhurst et al., 2006), although the anomaly that is still visible at the core-mantle boundary may
represent lithosphere that subducted well thereafter at this trench. This subduction is associated
with a continuous volcanic arc peaking in activity during the Late Permian (280-270 Ma) and
remaining active until the Early Triassic (~245 Ma) (Pankhurst et al., 2006; Ramos, 2008). This
episode of subduction culminated in the Gondwanide orogeny of the Triassic (~250-215 Ma),
followed by regional extension (Pankhurst et al., 2006; Cawood and Buchan, 2007; Elliot and
Fanning, 2008; Ramos, 2008; Tankard et al., 2009). The timing of slab detachment is highly
uncertain, and may be inferred from the tectonic model of Dalziel et al. (2013) who interpreted
the following sequence of events. Prior to 270 Ma a mantle plume impinged under the subducting
slab, underplating it and causing the subducting slab to flatten. This induced the Gondwanide
deformation and reduced arc volcanism during the 270–240 Ma period. Normal oceanic subduction

103
Figure 5.38. Gibraltar-West anomaly. Legend same as figure 5.5. Relative amplitude strength,
vertical, and dip trend are very similar between tomographic models. Lateral extent differs.

and arc volcanism returned after the period of Gondwanide deformation. At approximately 200
Myr, a new mantle plume impinged beneath the subducted slab below Patagonia and southern
Africa, eventually resulting in the Karoo-Ferrar Large Igneous Province at 182 Ma, and break-up
and detachment of the Gondwana Island slab. We assume this 200-180 Ma age as the age of break-
off, somewhat younger than in our previous interpretation of van der Meer et al. (2010).

5.34. Gibraltar - Gb
The Gibraltar anomaly (Figure 5.38) is located in the upper mantle below the Alboran Sea. Its
base is resting on the 660 km discontinuity, and it is still connected to Atlantic crust in the west,
is dipping eastwards and is interpreted as subducted lithosphere. Below southern Iberia, the slab
curves into an E-W strike, dipping southwards, and is largely detached. The first tomographic
images of subduction remnants in the Gibraltar region imaged the E-W trending, detached
portion below southern Iberia Spakman 1986a; Blanco and Spakman (1993) and interpreted
subduction in the Gibraltar region to have occurred southwards. With the advent of more advanced
tomographic models it became clear, however, that the Gibraltar slab is curved, and mainly dips
eastwards reaching the 660 km discontinuity (Gutscher et al., 2002; Piromallo and Morelli
2003; Spakman and Wortel, 2004; Bezada et al., 2013; van Hinsbergen et al., 2014; Chertova
et al. 2014). Since plate convergence between Iberia and Africa is only 100-250 km (increasing
eastwards) since the Cretaceous, the bulk of subduction of the Gibraltar slab must have resulted
from roll-back. Assuming that the slab did not stretch significantly, subduction of the Gibraltar slab
started below the Balearic islands (Spakman and Wortel, 2004), with plate circuit and geological
constraints suggesting an onset of deep underthrusting around 45 Ma following very slow
convergence between 85 and 45 Ma (van Hinsbergen et al., 2014), culminating in roll-back since
~27 Ma (Lonergan and White, 1997; Rosenbaum et al., 2002b; Faccenna et al., 2004; 2014). The
E-W trending, south-dipping portion of the Gibraltar slab below southern Iberia likely resulted

104
Figure 5.39. Great Basin anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical,
lateral extent and dip trend are very similar between tomographic models.

from >180° counterclockwise rotation during roll-back (Rosenbaum et al., 2002b ; Spakman and
Wortel 2004; Chertova et al., 2014; van Hinsbergen et al., 2014) and its detachment likely occurred
following the arrest of rapid thrusting in the sub-Betic fold-thrust belt, around 7-8 Ma (Platt et al.,
2003; García-Castellanos and Villaseñor, 2011).

5.35. Great Basin - GB


A positive wave speed anomaly below the Great Basin, Nevada (Figure 5.39) was originally
named the Nevada Cylinder (Roth et al., 2008). The Great Basin anomaly is located in the upper
mantle below southwestern North America, and has been imaged at depths of~100-800 km (West
et al., 2009). It has an overall N-S trend and dips to the NE. It has been interpreted as either a
lithospheric drip due to gravitational instability (West et al., 2009), a relic of the Laramide flat
slab (Schmandt and Humphreys, 2011), perhaps fragmented by fracture zones (Sigloch, 2011), or
a putative mantle plume (Obrebski et al., 2010). Based on its elongation and dip, we interpret the
anomaly to result from Cenozoic subduction, and follow the interpretation of Liu and Stegman
(2011) who consider it subducted Farallon oceanic lithosphere. Based on mantle convection
modelling, they argued that its subduction initiated at ~35 Ma, and ended with slab detachment
at ~15 Ma. Its subduction may correspond to phase of ignimbritic volcanism in the Great Basin
between 31-20 Ma (Best, 1991; Dickinson, 2006).

5.36. Halmahera - Hm
The Halmahera anomaly (Figure 5.40) corresponds to the A4 anomaly of Hall and Spakman (2002;
2004) and was previously imaged by Widiyantoro and van der Hilst (1997), , Hall and Spakman
(2015), and Wu et al. (2016). It is interpreted as one of the two actively subducting slabs attached to
the Molucca Sea basin lithosphere, the other being the Sangihe slab (section 5.76). It is subducting
eastwards below the island of Halmahera and reaches the base of the upper mantle. Wu et al. (2016)

105
Figure 5.40. Halmahera anomaly. Legend same as figure 5.5. This slab is well imaged in the
UU-P07 model with clear dip trend, but lateral and vertical extent are less well imaged in the
S-wave models.

separated this slab into three slabs; Molucca Sea East slab in the upper mantle and East Asian Sea
west and Ayu Trough Deep in the upper-lower mantle transition zone. In our tomographic model
we cannot confidently separate these slabs. Subduction below Halmahera started at 15 Ma in both
reconstructions of Hall (2002) and Wu et al. (2016), which we adopt as age for the base of the slab.

5.37. Hatteras - Ha
The Hatteras anomaly (Figure 5.41) is part of the set of anomalies commonly, also in our previous
compilation in van der Meer et al. (2010), interpreted as the Farallon slab, initially defined by Grand
et al. (1997). It is located below eastern North America within the lower mantle. The anomaly is
generally interpreted to represent eastward subducted Farallon oceanic lithosphere (Grand et al.,
1997; Bunge and Grand, 2000; van der Meer et al., 2010) below the North American continental
margin. The Hatteras slab corresponds to the MEZ and in part the SF1 slab of Sigloch and
Mihalynuk (2013), who, based on their interpretation of the geological records of the Wrangellia
and Stikinia composite terranes of the North American Cordillera, interpreted the slab to be
predominantly intra-oceanic, westward subducted Mezcalera Ocean lithosphere (their MEZ slab)
and only a small part Farallon-derived (their SF2 slab). Their pre-100 Ma interpretation, however,
is based on a true-polar wander corrected paleomagnetic reference frame without longitude control
under the assumption that Africa was longitudinally fixed (Torsvik et al., 2008). Van der Meer et al.
(2010) showed that such an assumption leads to major mis-locations of reconstructed subduction
zones relative to their corresponding slabs, particularly for the still-subducting Aegean slab, of
1000-2000 km. A much better match between slabs and subduction zones arises with a westward
longitude shift, which places the location of the Hatteras slab much closer to the west-North
American margin, making a Mezcalera origin unlikely. In addition, Liu (2014) found that westward
subduction of Mezcalera oceanic lithosphere would not be able to generate sufficient volume for

106
Figure 5.41. Hatteras anomaly. Legend same as figure 5.5. Dip trend, Relative amplitude strength,
vertical and lateral extent are very similar between tomographic models.

the slab. One of the arguments supporting that part of the Hatteras slab resulted from intra-oceanic
subduction is the mismatch between the reconstructed location and shape of the North American
continental margin and the outline of the slab (Sigloch and Mihalynuk, 2013). This occurs mostly
towards the south, an area where Sigloch and Mihalynuk (2013) have interpreted the Franciscan
intra-oceanic subduction zone, forming the deeper parts of the Cocos slab (their SF1 slab, see
section 5.29). By extending the Cocos slab and preceding Franciscan intra-oceanic subduction into
the deepest parts of the mantle, this would resolve most of the mismatch. The Hatteras slab would
then represent (predominant) subduction at the continental margin.
Debate exists on the age of onset of continental margin subduction, with estimates varying from
~100 Myr (Grand et al., 1997), prior to 100 Ma (Liu et al., 2008), or since 150 Ma (Sigloch et
al., 2008). In the interpretation of Sigloch and Mihalynuk (2013), the intra-oceanic subduction
zone was overridden by the continental margin, followed by initiation of Rocky Mountain
deformation, recorded by synorogenic clastic wedge (160–155 Ma), which we adopt as minimum
age of start of subduction. Tectonic studies (Ward, 1995; Nokleberg et al., 2000; DeCelles et al.,
2009), however, suggested that subduction at the western margin of North America (then part
of Laurentia) initiated earlier, from the Early Jurassic onwards. Consequently, van der Meer et al.
(2010) interpreted the base of the slab as Early Jurassic. Magmatism in the Sierra Nevada batholith
started at ~200 Ma (DeCelles et al., 2009), which we adopt as the maximum age of the base of the
Hatteras slab.
Towards the top of the lower mantle, the N-S trending Hatteras slab ceases to exist and is replaced
by the NW-SE trending SF3 slab of Sigloch and Mihalynuk (2013). This happened after the
South Farallon trench steps westward following accretion of the Shatsky Rise Conjugate plateau,
and a slab window formed leading to Sonora volcanism including the Tarahumara ignimbrite
province (Sigloch and Mihalnyuk, 2013). The volcanic rocks and granitic plutons were dated by
González-León et al. 2011. The age of the Tarahumara Formation is between ca. 79 and 59 Ma; the

107
Figure 5.42. Himalayas anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical
and lateral extent are very similar between tomographic models. The slab does not have a dip trend
and appears flat-lying just below the transition zone.

monzonitic to granitic plutons have ages of ca. 71–50 Ma (We have adopted 59-50 Ma as the end
of subduction of the Hatteras slab.

5.38. Himalayas - Hi
The Himalayas anomaly (Figure 5.42) is located below the northern part of the Indian continent
from the upper part of the lower mantle up to the upper mantle. In previous tomographic studies
it has been referred to as the IV (Van der Voo et al., 1999b) or Hi anomaly (Hafkenscheid et al.,
2006), and it was also imaged by Li et al. (2008), and shown in Replumaz et al. (2004; 2010b;
c) and van Hinsbergen et al. (2012). Van der Meer et al. (2010) followed the interpretation of
Hafkenscheid et al. (2006), who suggested that this lithosphere represents Neotethyan lithosphere
that formed as back-arc lithosphere associated with the late Cretaceous Spontang arc. The
Himalaya anomaly, however, is the shallowest anomaly associated with convergence between
India and Asia, and consequently likely represents the youngest slab, interpreted to reflect either
continental crust of Greater India (Replumaz et al., 2010c), or (partly) oceanic crust of Greater
India that according to paleomagnetic data formed in Cretaceous time (Greater Indian Basin of
van Hinsbergen et al., 2012). Given the northward subduction polarity of India below Asia, the
southward dip of the Himalayas slab likely indicates that the slab is overturned (Replumaz et al.,
2010c). Although the inferred nature of the lithosphere of the Himalaya slab differs in the scenarios
of Replumaz et al. (2010c) and van Hinsbergen et al. (2012), the inferred age for top of the slab
are similar, 20±5 Ma, after which time India-Asia convergence can be fully accounted for by intra-
continental shortening and horizontal underthrusting of Indian lithosphere below the Tibetan
Plateau. The age of the base of this slab is difficult to assess. Based on the volume of the Himalaya
slab compared to India-Asia convergence rates, Replumaz et al. (2010b) estimated a ~35 Ma age,
and consensus exists that the Himalaya slab represents Greater Indian lithosphere, i.e. lithosphere

108
Figure 5.43. Hindu Kush anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with similar vertical extent and Relative amplitude
strength, but are considerably different extent. Dip trend is visible in UU-P07, but not in SL2013.

that subducted after the collision of northernmost Indian plate continental lithosphere with Asia
around 50 Ma. We adopt a 43±8, Ma for the base of the Himalaya slab.

5.39. Hindu Kush - HK


A narrow, E-W trending, steeply north-dipping anomaly was described below the Hindu Kush
region of North Pakistan, from deep earthquakes reaching 250 km depth (Pegler and Das, 1998;
Pavlis and Das, 2000; Lister et al., 2008; Zhang et al., 2011) and seismic tomography shows it
reaches a depth of ~600 km (Van der Voo et al., 1999b; Negredo et al., 2007; Li and van der Hilst,
2010; Replumaz et al., 2010b; Kufner et al., 2014; 2016) (Figure 5.43). The Hindu Kush anomaly
is typically seen as subducted Indian plate lithosphere attached to continental crust of NW
India (Negredo et al., 2007; Lister et al., 2008; Replumaz et al., 2010a; Kufner et al., 2016). This
interpretation requires that India has been horizontally underthrust northward below the Sulaiman
Lobe of Pakistan, south of the Hindu Kush, over a N-S distance of ~500 km to reach the location
where the Hindu Kush slab dips steeply northward. We may thus estimate the minimum age for
the moment of subduction of the base of the Hindu Kush slab by reconstructing since when India
and Asia have converged 1100 km (500 km northward underthrusting of India plus 600 km of slab
length) at the longitude of the Hindu Kush slab, assuming no overriding plate shortening, i.e. 30
Ma according to the reconstruction of van Hinsbergen et al. (2011b). Recently, however, Kufner
et al. (2014) suggested that the Hindu Kush slab may correspond to Asian lithosphere instead,
since at depth the Hindu Kush slab cannot be distinguished from the Pamir slab (see section 5.69).
Gaina et al. (2015) developed a scenario assuming that the Hindu Kush slab is Asian, showing that
it must have subducted between the Helmand Block of Afghanistan and the Tadjik depression.
Taking the westward disappearance of the Hindu Kush slab into account, these authors proposed
an (unconstrained) counterclockwise rotation of the Helmand block, keeping the Kandahar arc of

109
Figure 5.44. Hispaniola anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with but are considerably different in lateral and vertical
extent and Relative amplitude strength. Dip trend appears to be near-vertical UU-P07.

the Afghan block contiguous with the Kohistan-Gangdese arc of Karakoram-Tibet. This scenario
would provide ~600 km of Helmand-Tadjik convergence since ~50 Ma. Although speculative, this
age may provide a maximum age for the base of the Hindu Kush slab. We allow a 50-30 Ma age
range for the base of the Hindu Kush slab.

5.40. Hispaniola - His


The Hispaniola anomaly (Figure 5.44) is detached and located below the island of Hispaniola in
the northern Caribbean region, at a depth of 800-1400 km. It was first identified by van Benthem
et al. (2013), who interpreted it as the nGAC slab (northern Great Arc of the Caribbean). We
name the slab Hispaniola according to its current location. Its detached nature and location in the
top of the lower mantle and its position west of the Caribbean slab and east of and above the Cocos
slab suggests that it should be correlated to a geological record of past Cretaceous or Paleogene
subduction in the northern Caribbean region. Van Benthem et al. (2013) interpreted this anomaly
to represent lithosphere that subducted below Cuba and Hispaniola in Cretaceous to Eocene
time, which decoupled from the Venezuela slab (section 5.90), or ‘sGAC’ slab in terminology of
van Benthem et al. (2013), upon the Paleogene northward retreat of the Cuba segment relative to
South America (see also Pindell and Kennan, 2009; Boschman et al., 2014). Cuba and Hispaniola
show a history of subduction of the North American plate, comprising oceanic Proto-Caribbean
lithosphere, a microcontinental terrane, and the North American continental margin below the
oceanic Caribbean plate represented by ophiolite and arc rocks (García-Casco et al., 2008; Pindell
and Kennan, 2009). The oldest record of subduction in Cuba is represented by ~130±5 Ma volcanic
arc and high-pressure, low-temperature metamorphic rocks (García-Casco et al., 2006; Stanek et
al., 2009), and subduction terminated between 45 and 40 Ma shown by the youngest foreland basin
deposits and oldest overlap assemblages (Iturralde-Vinent et al., 2008).

110
Figure 5.45. Hudson anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical and
lateral extent are very similar between tomographic models. There is not a dip trend visible.

5.41. Hudson - Hu
The Hudson anomaly (Figure 5.45) is part of the set of anomalies commonly, also in our previous
compilation in van der Meer et al. (2010), referred to as the Farallon slab, initially defined by Grand
et al. (1997). It is located below northern North America within the lower mantle. The NW-SE
trending Hudson anomaly is interpreted as the northernmost slab of the family of Farallon slabs
and is entirely contained in the lower mantle. At its eastern end it connects to the northern part of
the N-S trending Hatteras slab (section 5.37).
Sigloch and Mihalynuk (2013) referred to the Hudson slab as the ANG slab, and inferred that it
formed due to southwestward subduction of Angayucham oceanic lithosphere since 140 Myr based
on their interpretation of the amalgamation history of the Wrangellia superterrane of the North
American Cordillera. We infer Angayucham subduction led to the formation of the Lougheed
slab instead (section 5.51). Rather, we follow Nokleberg et al. (2000), who suggested that eastward
subduction of Farallon/Panthalassa lithosphere occurred at this latitudinal position at the Stikinia
arc while it was (close to) being accreted to the continental margin subduction zone (Hatteras slab,
section 5.37). Collision started by 72-69 Myr and was accomplished by 55-50 Myr ago (Sigloch
and Mihalynuk; 2013), which we take as age range for the top of the slab.
Johnston and Borel (2007) interpreted a two-stage process, a first involving intra-oceanic
subduction and arc formation and accretion forming the Stikinia-Quesnellia superterrane between
230 Ma to 150 Ma within the (eastern) Panthalassa. In a second stage, from 150 Ma to 55 Ma, this
arc migrated towards North America, collided, and moved northward parallel to the continental
margin. We interpret stage one to have resulted in the Wichita slab (section 5.91) and correlate
their stage two with the Hudson slab.

111
Figure 5.46. Idaho anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with a similar lateral and vertical extent and Relative
amplitude strength. The slab appears to dipping to near-vertical in UU-P07.

5.42. Idaho - Id
The Idaho anomaly (Figure 5.46) is located below western North America from the deep mantle
up to upper part of the lower mantle. It was interpreted as the Idaho slab by van der Meer et al.
(2010) and corresponds to the S2 anomaly of Sigloch et al. (2008) and Shephard et al. (2013) and
Cascadia Root (CR) anomaly of Sigloch and Mihalnyuk (2013). It has been interpreted as Farallon
lithosphere that subducted in the Cenozoic Sigloch et al. (2008) and Shephard et al. (2013) or
as lithosphere of the Kula plate that subducted in the Cretaceous (Ren et al., 2007; Shephard
et al., 2013). Sigloch et al. (2008) suggested that the Idaho slab subducted after, and below the
northern extent of the Farallon slab to the NE (defined here as the Hudson slab, section 5.41)
after the Hudson slab broke off in the late Cretaceous. Van der Meer et al. (2010) argued that such
an interpretation would require that the Hudson slab remained stagnant in the mantle since the
Late Cretaceous, whereas the Idaho slab sank rapidly into the lower mantle, which they considered
unlikely. Instead, they followed Nokleberg et al. (2000), who reconstructed northward to eastward
subduction of Farallon/Panthalassa lithosphere in the latitude range of the Idaho slab based on
paleomagnetic data from the northern parts of the Wrangellia terrane. The Wrangellia terrane
comprises of four island arcs, the Devonian Sicker, the late Paleozoic Skolai, the Late Triassic and
Early Jurassic Talkeetna-Bonanza, and the Late Jurassic through mid-Cretaceous Gravina arcs
(Nokleberg et al. 2000). Van der Meer et al. (2010; 2012) interpreted the Idaho slab as Farallon/
Panthalassa lithosphere resulting in the Talkeetna-Bonanza and Gravina arcs. U-Pb zircon ages of
midcrustal plutons of he Bonanza arc show that these were emplaced between 203.8 Ma and 164
Ma (Canil et al. 2012, D’Souza et al. 2015). Based on their present-day relative locations between
the Mendocino (section 5.59) and Idaho slabs and arcs, we tentatively we suggest that the Idaho
slab is correlated with the Bonanza arc. Following one of the accretion scenarios of Nokleberg et
al. (2000), end of subduction and accretion to the continental margin was modelled by Shephard

112
Figure 5.47. India anomaly. Legend same as figure 5.5. Positive anomalies are identified in the same
location in both tomographic models with a similar vertical extent and Relative amplitude strength.
Lateral extent differs. The slab does not have identifiable dip.

et al. (2013), who modelled accretion to occur at 140 Ma. However, Sigloch and Mihalynuk
(2013), suggested that the final terrane accretion occurred later at 55–50 Myr ago. However,this
late accretion date is specifically referring to the Siletzia terrane, which may be mostly a large
igneous plateau related to the Yellowstone plume (Wells et al. 2014) and therefore not the result
of subduction. Gehrels et al. (2009) argued that on the basis of U-Pb and Hf isotope analysis of
detrital zircons from strata of the Gravina belt that during the mid-Cretaceous the Gravina belt
collapsed and the Alexander-Wrangellia terrane was accreted until 85 Ma to the Stikine and
Yukon Tanana terranes at the continental margin. High-flux magmatism continued during this
accretionary event at 100–80 Ma, which we associate with the final intra-oceanic arc accretion A
dramatic reduction in magmatic flux occurred at ca. 78 Ma, which continued until ca. 55 Ma. We
therefore associate a break-off age of 85-78 Ma in our current compilation, defining the top of the
Idaho slab.

5.43. India - In
The India anomaly (Figure 5.47) is located below India. The India anomaly was classified by Van
der Voo et al. (1999b) as the ‘II’ anomaly, by Replumaz et al. (2010c) as the ‘TH’ anomaly, and it was
also identified by Hafkenscheid et al. (2006), Li et al. (2008), and van Hinsbergen et al. (2012). The
India anomaly is widely interpreted to represent Neotethyan oceanic lithosphere that subducted
below the southern Tibetan Plateau before the inception of India-Asia collision. Initiation of its
subduction probably occurred after the collision of the Lhasa terrane with the Qiangtang terrane
of central Tibet, which occurred ~130 Ma-120 Ma ago (Yin and Harrison, 2000; Kapp et al., 2007;
Li et al., 2017). Subduction of the India slab generated the long-lived Gangdese volcanic arc on the
Lhasa terrane since early Cretaceous time ( Ji et al., 2009) and was associated with the formation of
supra-subduction zone ophiolites in the Lhasa forearc (Huang et al., 2015; Maffione et al., 2015b).

113
Figure 5.48. Izu-Bonin anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model and shows a clear dip trend. Relative amplitude strength, lateral extent and vertical extent
are very different between models.

We date the base of the India slab at 130 ± 10 Ma and note that this is a minimum age. The top of
the slab coincides in age with that of the base of the Himalaya slab of 43 ± 8 Ma (section 5.38), the
subduction of which accommodated India-Asia convergence after break-off of the India slab.

5.44. Izu-Bonin - IB
The Izu-Bonin anomaly (Figure 5.48) is interpreted to reflect Pacific lithosphere that subducted
beneath the Philippine Sea Plate. It is well-imaged in seismic tomographic models of van der
Hilst et al. (1991), Bijwaard et al. (1998), Gorbatov and Kennett (2003), Widiyantoro et al. (1999),
Sugioka et al. (2010), Fukao and Obayashi (2013), and Jaxybulatov et al. (2013) and extends from
the present-day Izu-Bonin-Mariana trench to the base of the upper mantle, where it is horizontally
overlying the 660 km discontinuity. The Izu-Bonin slab is connected to the Mariana slab (section
5.57) in the upper 300-400 km of the mantle, but is disconnected through a vertical slab tear at
greater depth, allowing for the Mariana slab to penetrate steeply into the lower mantle (Miller
et al., 2004; 2005). Previous reconstructions estimated subduction of the Mariana-Izu Bonin
subduction zone to have started at 48 Ma (Seno & Maruyama, 1984) to 50 Ma (Wu et al. 2016).
The Mariana-Izu Bonin forearc has been instrumental in the development of models linking
geochemical evolution of subduction-related magmas in a forearc position to subduction initiation
(Stern and Bloomer, 1992; Dewey and Casey, 2011; Stern et al., 2012). U/Pb and 40Ar/39Ar
ages of the oldest forearc lavas that are believed to have formed during subduction initiation are
consistently 51-52 Ma (Ishizuka et al., 2011; Reagan et al., 2013; Arculus et al., 2015), which we
adopt as age for the onset of subduction of the Izu-Bonin slab.

114
Figure 5.49. Juan de Fuca anomaly. Legend same as figure 5.5. This slab is well imaged in the
UU-P07 model and shows a clear dip trend. Relative amplitude strength, lateral extent and vertical
extent are very different between models.

5.45. Juan de Fuca - JdF


The Juan de Fuca anomaly (Figure 5.49) is located below western North America, is N-S
trending and is interpreted to be the still-subducting slab at the Cascadia subduction zone. In
tomographic studies it has been imaged to depths of 250-400km (Sigloch et al., 2008; Schmandt
and Humphreys, 2011; Chu et al., 2012). The Juan de Fuca slab has been interpreted as the result
of the reinstatement of normal subduction following the accretion of the Siletzia microcontinent,
and a subsequent period of flat-slab subduction (Schmandt and Humphreys, 2011). Above the slab
lies the Cascadian arc, where magmatism started at 45-40 Ma (Schmandt and Humphreys, 2011;
Wells and McCaffrey, 2013). The slab has been interpreted to be broken up by the Yellowstone
plume (Obrebski et al., 2010, Long 2016). Fragmentation of the slab presumably occurred just prior
to the arrival of the plume at the surface, around 19–17 Ma (Obrebski et al., 2010. We adopt this
as the age of the base of imaged slab. Some weak positive anomalies are imaged in the UU-P07
model towards the east and northeast, named the F1 and F2 slabs in Obreski et al. (2010) and
Long (2016), which may represent the remnant slabs of 45-19 Ma subduction, separated from the
Juan de Fuca slab by the Yellowstone plume.

115
Figure 5.50. Kabylides anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model and there is a similar relative amplitude strength, lateral extent and vertical extent are very
different between models. This slab has a dip in the upper mantle, but appears to flat-lying in the
shallow lower mantle.

5.46. Kabylides - Kb
The Kabylides anomaly (Figure 5.50) is located in the upper mantle below the north African
margin of Algeria and was first identified by Spakman and Wortel (2004). It dips steeply, resting
on the 660 km discontinuity, and is interpreted as a slab that is detached above 200 km depth. The
Kabylides anomaly is interpreted to represent part of the African plate that subducted northward
below the Baleares margin since ~85-45 Ma (van Hinsbergen et al., 2014) then still contiguous
with the Gibraltar slab (section 5.34). Subsequent roll-back led to southward retreat of the
subduction zone, and underthrusting of the North African margin in Miocene time (Michard et
al., 2006), followed by slab segmentation and further westward retreat of the Gibraltar slab (section
5.34), leaving the Kabylides slab behind which subsequently detached from the north African
margin (Spakman and Wortel, 2004; Michard et al., 2006; Chertova et al., 2014; van Hinsbergen
et al., 2014). The age of slab break-off is estimated at 15-12 Ma, inferred from the end of African
underthrusting below the leading edge of the European plate exposed in the Kabylides massifs, as
well as break-off induced magmatism (Coulon et al., 2002; Michard et al., 2006).

5.47. Kalimantan - Ka
The Kalimantan anomaly (Figure 5.51) is located below southeast Asia from the middle to the
upper part of the lower mantle (Widiyantoro and van der Hilst, 1996a; 1997; Rangin et al., 1999;
Replumaz et al., 2004; Hall et al., 2008; Zahirovic et al., 2012; Fukao and Obayashi, 2013). It is
frequently interpreted as one single slab together with the Sunda slab (Section 5.86). However, the
Kalimantan anomaly is not present in the western part of the Sunda slab, where the Sunda slab
does not penetrate the upper-lower mantle boundary (Hall and Spakman 2015), and its tip lies
just to the south of the Kalimantan anomaly. In addition, the Kalimantan slab is striking NE-SW,

116
Figure 5.51. Kalimantan anomaly. Legend same as figure 5.5. Relative amplitude strength, vertical
and lateral extent are very similar between tomographic models. The slab does not have a dip trend.

whereas the Sunda slab strikes WNW-ESE. This suggests that the Kalimantan anomaly results
from a history of subduction north of, and starting earlier than the Sunda slab. Slabs to the east of
the Kalimantan anomaly that started to subduct around 55-45 Ma, e.g., Izu-Bonin (section 5.45),
Mariana (section 5.57), Arafura (section 5.10), or Caroline Ridge (section 5.23) still reside in the
upper mantle or penetrated into the top of the lower mantle. The base of the Kalimantan anomaly
is thus likely of Late Cretaceous or Paleogene subduction age. Hall (2012) concluded that westward
subduction below west Sulawesi, Sumba, and Borneo (i.e. Kalimantan) started ~70-65 Ma, and
ended 50-45 Ma, at which time subduction of the Sunda slab started to the south and west (section
5.85). In addition, Hall and Spakman (2015) suggested that the Kalimantan anomaly may host
of two separate slabs, whereby the upper part of the anomaly represents subducted proto-South
China Sea lithosphere that subducted southeastward beneath North Borneo and the Cagayan arc,
now southeast of Palawan, between 45 and 20 Ma. This interpretation was adapted by Wu et al.
(2016), where the western part of the anomaly was marked as Proto South China Sea and the east
as one of two western East Asian Sea anomalies, subducting between 45-35 Ma and 50-20 Ma
respectively. Wu et al. 2016 distinguished another slab in between these anomalies and named it
Molucca Sea West, being a continuation of the slab in the upper mantle, with an interpreted start
of subduction at 30 Ma. In the UU-P07 model it is not possible to confidently distinguish the
boundary between these two or three slabs, and we adopt an age range of subduction of 70-65 to 20
Ma for the composite the Kalimantan slab(s).

117
Figure 5.52. Kamchatka-Kuriles anomaly. Legend same as figure 5.5. This slab is well imaged in the
UU-P07 model and there is a similar Relative amplitude strength, lateral extent and vertical extent
are very different between models. This slab appears to flat-lying on top of the transition zone.

5.48. Kamchatka-Kuriles - Kc
The Kamchatka-Kuriles anomaly (Figure 5.52) is located below the Sea of Okhotsk and is
interpreted to represent a slab that is still subducting at the Kamchatka-Kuriles subduction zone,
penetrating down to the upper part of the lower mantle. We previously named it the Kamchatka
slab (van der Meer et al., 2010) but renamed it to include the widely used Kuriles slab (Spakman
et al., 1989; Van der Hilst et al., 1991; Gorbatov et al., 2000; Jiang et al., 2009; Koulakov et al.,
2011). It represents northwestward-subducted Pacific oceanic lithosphere. The part of the slab
below 660 km discontinuity was interpreted to have an age of at least 65-55 Ma (Gorbatov et
al., 2000). Nokleberg et al. (2000) and Golonka et al. (2003) proposed that the onset of westward
subduction below Kamchatka started earlier and formed the Olyutorka arc since the middle
to Late Cretaceous. Van der Meer et al., 2010) followed that latter interpretation. Following the
more recent tectonic models of Konstantinovskaia (2001), Hourigan et al. (2009), and Shapiro and
Solov’ev (2009), accretion of the intra-oceanic Aichayam-Valangskyi ophiolite terrane including
the Olyutosky arc to the Asian continental margin followed upon a phase of SW-ward subduction
that ended around 52Ma (Konstantinovskaia, 2001; Hourigan et al., 2009), which we correlate
to the Agattu slab (section 5.2). This was followed by a subduction polarity reversal and onset of
NW-ward subduction of Pacific lithosphere during the Middle Eocene (Konstantinovskaia, 2001;
Shapiro and Solov’ev, 2009), simultaneously with the Aleutian slab (section 5.3) to the northeast,
which we interpret to as the age of the base of the Kamchatka-Kuriles slab.

118
Figure 5.53. Komsomolets anomaly. Legend same as figure 5.5. Positive anomalies are identified
in the same location in both tomographic models with a similar vertical extent. Relative amplitude
strength and lateral extent differ between tomographic models.

5.49. Komsomolets - Km
The Komsomolets anomaly (Figure 5.53) is located below northernmost Asia within the deep
mantle and was not previously defined. It is E-W trending and its shallowest and easternmost
location it may connect to the Chukchi slab (section 5.28). Based on the Chukchi slab as well
as the Mongol-Okhotsk slab to the south (Section 5.63), we interpret the anomaly as a Lower
Mesozoic slab subducted in the paleo-Arctic region. In the Arctic, a record of mid-Mesozoic
subduction, along a continental margin, is preserved in the form of the Svyatoy-Nos arc located in
northernmost Siberia (Nokleberg et al., 2000). According to Nokleberg et al. (2000) and Sokolov
(2010) this arc was active in the Late Jurassic to Early Cretaceous and was part of a south dipping
subduction zone, subducting South Anyui oceanic lithosphere. We date the subduction of the
Komsomolets slab by this arc. In the reconstruction of Shephard et al. (2013) the Svyatoy-Nos
arc is not modelled as such but would be located as the time-equivalent continuation of the Oloy
subduction zone in between Siberia and Baltica.

119
Figure 5.54. Lake Eyre anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with a similar vertical extent. Relative amplitude
strength and lateral extent differ between tomographic models.

5.50. Lake Eyre - LE


The Lake Eyre anomaly (Figure 5.54) was defined by Schellart and Spakman (2015). It is located
at 800-1200 km depth in the upper part of the lower mantle below southern Australia. There is no
record of Mesozoic or Cenozoic subduction within Australia, and pointing at a rapid northward
absolute plate motion of Australia suggested by hotspot reference frames, Schellart and Spakman
(2015) suggested that the geological record of subduction of the Lake Eyre slab is best sought
along the northern Australian margin. Those authors correlated the slab to a period of northward
intra-oceanic subduction that started ~70 Ma ago north of Australia. This phase terminated when
Australian continental lithosphere arrived in the trench. Relics of the overriding oceanic lithosphere
are now found as ophiolites on Papua New Guinea and the Pocklinton Trough area. The end of
their emplacement onto the Australian margin occurred around 50 Ma and was suggested as timing
of slab break-off (Schellart and Spakman, 2015). Subsequently, Australia moved northward and
overrode the detached slab.

5.51. Lougheed - Lo
The Lougheed anomaly (Figure 5.55) is located below northernmost North America within the
mid-mantle. It is N-S to NNW-SSE trending and at its southern end lies close to, but is separate
from the NW-SE trending Hudson anomaly that is correlated to the Stikinia arc in northwest
Canada (section 5.41). The Lougheed slab should thus correlate to circum-Arctic subduction to
the north of the Stikinia terrane. Nokleberg et al. (2000) documented the Koyukuk arc in this area
and interpreted this as the result of subduction of Angayucham and Goodnews Ocean lithosphere
below North American continental lithosphere. The Koyukuk arc is mainly of Late Jurassic to Early
Cretaceous age. In the plate motion model of Shephard et al. (2013), these different elements were

120
Figure 5.55. Lougheed anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with a similar vertical, lateral extent and Relative
amplitude strength.

brought together. They concluded that subduction below the Koyukuk and the adjacent Nutesyn
arcs was active between 160-120 Ma, which we adopt for the subduction age range of the anomaly.

5.52. Maldives - Md
The Maldives anomaly (Figure 5.56) is located below the northwestern Indian Ocean from the
deep mantle up to the upper part of the lower mantle. In previous tomographic studies it has been
referred to as the eastern part of the III anomaly (Van der Voo et al., 1999b) or as the IO anomaly
(Hafkenscheid et al., 2006), and was interpreted to result from northward subduction within
the Neotethys ocean. We now distinguish the deeper NW-SE trending Maldives anomaly from
the shallower SW-NE Carlsberg slab (section 5.22) (Gaina et al., 2015), which were previously
considered a single feature (van der Meer et al., 2010). van der Meer et al. (2010) interpreted
an onset of subduction of this anomaly to occur as early as the Late Triassic comparing to
reconstructions of Stampfli and Borel (2004), which we now consider unlikely given the adjacent
Carlsberg, India (section 5.43), and Arabia (section 5.9) slabs that subducted in Cretaceous time.
More recently, the age of the Maldives slab was interpreted from geological evidence of intra-
oceanic subduction from ophiolites found in the Indus-Yarlung suture zone, thrust on the Tibetan
Himalaya (e.g., Spontang and Xigaze ophiolites) (Hébert et al., 2012). These were suggested to
have formed at equatorial latitudes based on paleomagnetic data (Abrajevitch et al., 2005), and
were interpreted to have thrust onto the Tibetan Himalaya around 70 Ma based on arrival of mafic
debris in the Tibetan Himalayan stratigraphy around that time (Searle et al., 1997; Corfield et al.,
2001). Because paleomagnetic data of the Tibetan Himalaya also show equatorial latitudes around
70 Ma (Patzelt et al., 1996), van Hinsbergen et al. (2012) linked the record of the Himalayan
ophiolites to the Maldives anomaly.

121
Figure 5.56. Maldives anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with a similar vertical extent. Relative amplitude
strength and lateral extent differ between tomographic models.

Recently, however, Garzanti and Hu (2015) demonstrated that the~70 Ma mafic debris in the
Tibetan Himalaya is not ophiolite-derived, but related to plume volcanism, perhaps from early-
stage volcanics of the Deccan traps. Ophiolite-derived debris did not arrive in the Tibetan
Himalayan stratigraphy until the Early Eocene instead. In addition, Huang et al. (2015) showed
that the Xigaze ophiolites are unconformably overlain by Tibet-derived forearc sediments with ages
very close to the formation ages of the ophiolite (~130-125 Ma). In addition, they showed that
the previous paleomagnetic data of Abrajevitch et al. (2005) were strongly affected by compaction-
induced inclination shallowing, and provided a ~16.5°N paleolatitude for these ophiolites instead,
i.e. immediately adjacent to the south Tibetan margin. Its current depth in the mantle is consistent
with Cretaceous subduction, and when the global plate circuit (e.g., Seton et al., 2012) is cast in a
hotspot reference frame (e.g., Doubrovine et al., 2012), the Maldives anomaly is located within the
Neotethys ocean below the India-Arabia plate boundary. At this plate boundary, there is evidence of
subduction farther south, culminating in obduction of the Waziristan-Khost ophiolite onto Arabia
around 80 Ma (e.g., Gaina et al., 2015). The nearest record of intra-oceanic subduction may be the
Kohistan arc, and Jagoutz et al. (2015) argued for a low latitude of Cretaceous subduction below
this arc. Borneman et al. (2015), however, documented Asia-derived detritus from Cretaceous
sandstones within the Kohistan stratigraphy, arguing against such a scenario. At this stage, we
therefore refrain from interpreting a geological record of subduction for the Maldives anomaly.

5.53. Malpelo - Mp
The Malpelo anomaly (Figure 5.57) is located below the western Panama Basin and NW South
America in the mid-mantle. It is NW-SE trending and above 1175 km it is close to the Venezuela
slab to the east (section 5.90) as well as the Cocos slab to the north (section 5.29) and Brasilia
slab to the south (section 5.18). The anomaly we here define as the Malpelo slab was previously

122
Figure 5.57. Malpelo anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with a similar relative amplitude and vertical extent.
Lateral extent differs between tomographic models. The slab appears to have a dip towards the SW
in UU-P07.

interpreted by Taboada et al. (2000) to represent subducted Farallon plate lithosphere (labelled FaP
in their figures), beneath the Panama-Costa Rican Arc and the Choco terrane in northwest South
America. Meschede and Frisch (1998) proposed that this subduction zone extended to the north,
and connected to the subduction zone below the Guerrero terrane and Chortis block, which are
now located in present day Honduras, Guatemala, and Mexico. Because the Malpelo slab appears
to have a slight westward dip, van der Meer et al. (2010; 2012) followed models of Dickinson and
Lawton (2001) proposed for the Guerrero arc and interpreted the slab as the result of westward
subducting Mezcalera ocean lithosphere, followed by a collision with the North American margin
in the Late Cretaceous, followed by a subduction polarity switch resulting in eastward subducting
Farallon lithosphere at around 1040 km depth.
The Malpelo slab, however, extends to the south into a position directly west of, and at a similar
depth as the Venezuela slab (section 5.90). This slab is interpreted to result from westward
subduction of Proto-Caribbean lithosphere below the Caribbean plate (van der Meer et al., 2010;
Pindell et al., 2012; van Benthem et al., 2013). Placing the kinematic reconstruction of Boschman
et al. (2014) in the slab-fitted mantle reference frame of van der Meer et al. (2010) places the
eastern margin of the Caribbean plate above the Venezuela slab, but the western Caribbean margin
above the Malpelo slab: there is little space for a westward subducting ocean to the East of the
Malpelo slab. We therefore consider it more likely that the Malpelo slab originated from eastward
subduction west of the Caribbean plate.
The oldest reported evidence of subduction in Central America is arc volcanics of only ~75 Ma
(Buchs et al., 2010; Wegner et al., 2011), and Pindell et al. (2012) suggested that the modern
subduction zone below the eastern Caribbean region may have started ~90-85 Ma. This subduction
zone, however, is still active whilst the Malpelo slab is detached. We consider it more likely that

123
Figure 5.58. Manchuria anomaly. Legend same as figure 5.5. This slab is well imaged in the
UU-P07 model and there is a similar relative amplitude strength, lateral extent and vertical extent
between models. This slab appears to mostly flat-lying on top of the transition zone.

the current phase of subduction generated the slab above and slightly east of the Malpelo slab. This
slab is visible in a narrow depth interval around 1000-900 km, above which a prominent slab gap
is present that coincides with the location of and may be caused by the subduction of the Cocos-
Nazca ridge. Given its narrow range, we have not defined this body as a separate slab.
The Guerrero arc of Mexico may be a better candidate to date the Malpelo slab. This arc formed at a
paleolatitude coinciding with the northern portion of the Malpelo slab. The history of the Guerrero
arc is complex, but is seen as an intra-oceanic arc that was separated from Mexico by a back-arc
region that closed in Early Cretaceous time (Centeno-Garcia et al., 2011). Bajocian–Cenomanian
arc assemblages of the Guerrero terrane (Talavera-Mendoza et al., 2007; Martini et al., 2011; 2013)
and Callovian to Valanginian volcanic sulphides (Mortensen et al., 2008) suggest that the arc was
active from ~170-165 to ~95-90 Ma. We adopt this as age range for the formation of the slab. We
note, however, that our interpretation suggests that the Guerrero arc would have continued far to
the south, to the west of the Caribbean plate. Currently, there is no geological evidence to support
this interpretation and the interpretation of the geological record of the Malpelo slab may need
revision in the future.

5.54. Manchuria - Mc
The Manchuria anomaly (Figure 5.58), also referred to as Japan anomaly (Obayashi et al., 2009) was
interpreted as a slab that dips down westward from the Japan subduction zone at surface, drapes
the transition zone below east Asia and penetrated the upper part of the lower mantle in the west
(van der Hilst et al., 1991; Fukao et al., 2001; Miller and Kennett, 2006; Abdelwahed and Zhao,
2007; Zhao and Ohtani, 2009; Lei, 2012; Chen et al. 2017). In the uppermost mantle it has a N-S
trend and is connected at depths of <300km with the SW-NE trending Kamchatka-Kuriles slab
(section 5.48) towards the north and NNW-SSE trending Izu-Bonin slab (section 5.44) towards

124
Figure 5.59. Manila anomaly. Legend same as figure 5.5. This upper mantle slab is well imaged in
the UU-P07 model and there is a similar relative amplitude strength, lateral extent. Vertical extent
is different between models. This slab near-vertical in the upper mantle.

the south. It has previously been associated with westward subduction of the Pacific and Philippine
Sea plate starting around 40-45 Ma (van der Hilst et al., 1991; Abdelwahed and Zhao, 2007). This
timing was based on the evolution of the Philippine Sea plate only, whereas the Manchuria slab is
in fact located north of the Philippine Sea plate subduction zones. In van der Meer et al. (2010),
we considered an earlier, Late Cretaceous start of Pacific plate subduction more likely, indicated by
a Late Cretaceous-Eocene (110-50 Myr) magmatic phase in the Korean peninsula (Sagong et al.,
2005).
The complications of dating this slab might be related to two ridges subducting underneath Japan
as documented by Isozaki et al. (2010); first the Izanagi–Kula ridge around 120–110 Ma and
then the Kula-Pacific ridge around 70–60 Ma. Ren et al. (2002) documented tectonic inversion
in the Songliao Basin (NE China) at 77-67 Ma, which may date the onset of renewed subduction
following ridge subduction, followed by widespread regional extension and transtension as a result
of roll-back. More recently Seton et al. (2015) and Honda (2016) have the Izanagi-Pacific ridge
subducting between ~60-50 Ma. On the basis of the Korean inversion and the recently modelled
subduction of the last subducted ridge, we interpret the onset of Manchuria slab subduction to start
at 77-50 Ma and associate this timing with the base and westernmost extent of this slab.

5.55. Manila - Ml
The Manila anomaly (Figure 5.59) is located below the northewestern Phillipine Sea in the upper
mantle. The Manila anomaly dips down towards the transition zone at the base of the upper mantle
and has previously been imaged in several regional and global tomographic studies (Rangin et al.,
1999; Lallemand et al., 2001; Zhao and Ohtani, 2009; Zheng et al., 2013; Koulakov et al., 2014;
Wu et al., 2016), named South China Sea slab (Rangin and Spakman, 1999; Lallemand et al.2001),
or Eurasian slab (Zheng et al, 2013; Zhao and Ohtani, 2009; Wu et al., 2016). It is interpreted as

125
Figure 5.60. Maracaibo anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model and has near-vertical dip trend. Relative amplitude strength, lateral extent and vertical extent
are different between models.

eastward dipping South China Sea oceanic lithosphere of the Eurasian plate still subducting at the
Manila trench below the Luzon arc. In the north, continental lithosphere of the South China block
has entered the trench, which led to an arc-continent collision with the Luzon arc, creating the
Taiwan fold-thrust belt (Sibuet and Hsu, 2004; Huang et al., 2014). Due to Pliocene arc-continent
collision, subduction in northern Taiwan is terminating and slab break-off is currently going on,
whereby the Ryukyu subduction zone is propagating westwards, accommodating a subduction
polarity flip (e.g., Ustaszewski et al., 2012). Estimates for the onset of Eurasian plate subduction
below the Luzon arc are 15-10 Ma (Sibuet and Hsu, 2004; Hall, 2012), which we adopt for the age
of the base of the Manila slab.

5.56. Maracaibo - Ma
The Maracaibo anomaly (Figure 5.60) is located in the upper mantle below northern South
America. It was first tomographically imaged by van der Hilst and Mann (1994), and was later
recognized in tomographic models by Taboada et al. (2000), Bezada et al. (2010), and van Benthem
et al. (2013). Van Benthem et al. (2013) shows that the Maracaibo slab has a shallow dip, and
is at least 900 km long from the surface to the base of the upper mantle. Another 300 km may
horizontally overlie the 660 km discontinuity.
The Maracaibo subduction zone is interpreted as actively subducting Caribbean plate lithosphere
that is overridden by the South American continent moving roughly W to WNW relative to
the Caribbean plate, leading to oblique convergence along the NW South American continent
(Kellogg and Bonini, 1982). Originally, the onset of Maracaibo subduction was estimated as ~45
Ma (Kellogg and Bonini, 1982), but recently, Ayala et al. (2012) showed that sedimentary basins in
NW South America became fragmented because of uplift and deformation of the Maracaibo block
as early as 58-55 Ma. Plate kinematic restorations of the Caribbean region (Pindell et al., 2012;

126
Figure 5.61. Mariana anomaly. Legend same as figure 5.5. This slab is well imaged in the UU-P07
model and there is a similar relative amplitude strength and vertical extent between models. This
slab is near-vertical in the upper mantle and top lower mantle.

Boschman et al., 2014) also show an onset of convergence across the Maracaibo margin around 60
Ma, and predict ~1200 km of subduction. This is consistent with the UU-P07 seismic tomographic
image of van Benthem et al. (2013) if the anomaly overlying the 660 km discontinuity is assumed
to be part of the Maracaibo slab. We thus assign a 60-55 Ma age for the base of the Maracaibo slab.

5.57. Mariana - Mr
The Mariana anomaly (Figure 5.61) was first shown in seismic tomography by Spakman et al.
(1989) and Van der Hilst et al. (1991). In subsequent tomographic models, the Mariana anomaly,
which is an steeply westward dipping, N-S trending anomaly that connects to the Mariana trench
and the slab is thus interpreted to represent still-subducting Pacific lithosphere. It was imaged to
penetrate the lower mantle down to 1000-1200 km depth (Bijwaard et al., 1998; Widiyantoro
et al., 1999; Gorbatov and Kennett, 2003; Huang and Zhao, 2006; Rost et al., 2008; Fukao and
Obayashi, 2013; Jaxybulatov et al., 2013; Obayashi et al., 2013; Zahirovic et al., 2014; Wu et al.,
2016). To the north along the same subduction zone, it is connected in the upper 300-400 km to
the Izu-Bonin slab (section 5.44). Below that depth, however, the Mariana and Izu-Bonin slabs are
disconnected through a tear fault that may have formed at the subducted Marcus–Necker Ridge
and the Ogasawara Plateau, with the Izu-Bonin slab lying horizontally on the 660 km discontinuity
and the Mariana slab penetrating through (Miller et al., 2004; 2005). To the south, the Mariana
slab is also bounded by a tear (Miller et al., 2006). Previous reconstructions estimated subduction
of the Mariana-Izu Bonin subduction zone to have started at 48 Ma (Seno and Maruyama,
1984) to 50 Ma (Wu et al. 2016). Prior to 50 Ma, Wu et al. postulate that a western Pacific plate
boundary already existed but was this may have been characterised by highly oblique subduction
or transforms. After 50 Ma, Pacific plate motions changed and fast subduction began (Wu et al.,
2016).

127
Figure 5.62. Mayn anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with similar downward vertical extent. Lateral extent
and relative amplitude strength, are different. There is not a visible dip trend.

The Mariana-Izu Bonin forearc has been instrumental in the development of models linking
geochemical evolution of subduction-related magmas in a forearc position to subduction initiation
(Stern and Bloomer, 1992; Dewey and Casey, 2011; Stern et al., 2012). U/Pb and 40Ar/39Ar
ages of the oldest forearc lavas that are believed to have formed during subduction initiation are
consistently 51-52 Ma (Ishizuka et al., 2011; Reagan et al., 2013), which we adopt as age for the
onset of subduction of the Mariana slab.

5.58. Mayn - Mn
The Mayn anomaly (Figure 5.62) is located below northeastern Siberia in the upper mantle and
uppermost part of the lower mantle. It is N-S trending and towards the south it touches the Bering
Sea slab (section 5.15). It was previously imaged by Gorbatov et al. (2000) and Zhao et al. (2010).
The tectonic evolution and upper mantle structure of the region leads us to interpret the Mayn
anomaly as a separate slab. The location of the slab is consistent with subduction along the Shirshov
Ridge in western Bering Sea. This ridge is generally interpreted to represent a Cenozoic intra-
oceanic arc (Nokleberg et al., 2000; Chekhovich et al., 2012). In the tectonic model of Chekhovich
et al. (2012), the Shirshov ridge underwent a period of imbricate thrusting between 30-15 Ma,
which we adopt as the period of subduction.

5.59. Mendocino - Mdc


The Mendocino anomaly (Figure 5.63), is located in the lower mantle below the northeastern
Pacific Ocean, west of the North American continental margin, west of the Hudson (section
5.41), Hatteras (section 5.37) and Idaho slabs (section 5.42) and south of the North Pacific slab
(section 5.68). It was named ‘X’ by Sigloch (2011) and and later Cascadia Root 2 (CR2) by Sigloch
and Mihalynuk (2013). Van der Meer et al. (2012) interpreted it as an (unnamed) slab remnant

128
Figure 5.63. Mendocino anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with a similar lateral extent. Relative amplitude
strength and vertical extent differ between tomographic models.

at 2300-1700 km depth in east-Panthalassa Ocean marginal subduction zones in the UU-P07


model (Amaru, 2007). Sigloch and Mihalynuk (2013), on the basis of a regional tomographic
model of Sigloch (2011) interpreted the slab to extend to the top of the lower mantle. The age of
subduction is poorly constrained from geological data. Sigloch and Mihalynuk (2013) interpreted
subduction to have occurred between 140-75 Myr on the basis of plate tectonic reconstruction. The
shallowest occurrence of the slab is close to the North American continental margin, suggesting
terrane accretion might have occurred in Cenozoic times. This is consistent with the inferences of
Sigloch (2011) and Sigloch and Mihalynuk (2013), who indicated final terrane accretions (Siletzia,
Pacific Rim) at 55-50 Myr. However as discussed at the Idaho slab, the Siletzia terrane is mostly an
oceanic plateau (Wells et al. 2014) and is not related to arc volcanism. We interpret the Idaho and
Mendocino slab to be genetically linked and both resulting from intra-oceanic subduction at the
Talkeetna-Bonanza arcs. Based on their present-day relative locations between these two slabs and
arcs, we tentatively we suggest that the Mendocino slab is correlated with the Alaskan Talkeetna
arc. This Alaskan arc was formed from 202-201 Ma onwards (Clift et al, 2004, Rioux et al. 2007).
On the basis of rapid exhumation and deposition of the coarse clastic Naknek formation Clift et
al. (2004) suggested it was possibly amalgamated between 160-125 Ma in the greater Wrangellia
terrane. We here interpret these events to represent amalgation with the Bonanza and Gravina arcs,
and correlated with the Idaho slab (section 5.42)

5.60. Mesopotamia - Me
The Mesopotamia anomaly (Figure 5.64) is located below the Zagros mountain belt at the Arabia-
Eurasia plate boundary from the deep mantle up to mid-mantle and connects upward to the Zagros
anomaly residing in the upper mantle. In previous studies it has been referred to as the western
part of the II anomaly Van der Voo et al. (1999b), as the SI and AI anomalies Hafkenscheid et al.

129
Figure 5.64. Mesopotamia anomaly. Legend same as figure 5.5. Positive anomalies are identified
in the same location in both tomographic models with a similar vertical extent. Relative amplitude
strength and lateral extent differ between tomographic models.

(2006), or as the Sb1 anomaly in Agard et al. (2011). In van der Meer et al. (2010), we interpreted
this anomaly as a slab resulting from intra-oceanic subduction within the Neotethys culminating in
the obduction of e.g. the Semail ophiolite of Oman. Closer inspection of the tomographic images
below Arabia shows, however, that the Mesopotamia slab at depths of 1100-1300 km is separated
from a second slab to the SW (now named Arabia slabs following Hafkenscheid et al. (2006), see
section 5.9) that is a better candidate to result from subduction below the ophiolites of Oman and
Iran. The Mesopotamia slab more likely relates to subduction below the Eurasian margin of Iran
(Agard et al., 2011). The oldest geological evidence for subduction below the Iranian continental
fragments, which collided with Eurasia in the Triassic (Muttoni et al., 2009) is the Sanandaj-Sirjan
magmatic arc, active since ~150 Ma (Agard et al., 2011), providing a probable age for the base of
the Mesopotamia slab. Towards the top, the slab appears to be disconnected from the upper mantle
portions of slabs (see Zagros slab, section 5.94). Agard et al. (2011) suggested, based on the age of
rapid exhumation of HP-LT metamorphic rocks in the Zagros Mountains that there may have
been a phase of slab break-off in the latest Cretaceous to Paleogene, ~65±5 Ma which we adopt as
the age of the top of the Mesopotamia slab.

5.61. Mississippi - Mi
The Mississippi anomaly (Figure 5.65) is part of the set of anomalies commonly, and also in our
previous compilation in van der Meer et al. (2010), interpreted as the Farallon slab as originally
defined by Grand et al. (1997). The NW-SE trending Mississippi anomaly has a central position
within the family of Farallon anomalies and is located in the mid-mantle below Central North
America and the northern Caribbean region. At its base it connects with the top of the N-S
trending Hatteras slab (section 5.37). At its top it is close to the Great Basin slab (section 5.35).
The Mississippi anomaly corresponds to the SF3 anomaly of Sigloch and Mihalynuk, (2013),

130
Figure 5.65. Mississippi anomaly. Legend same as figure 5.5. Vertical and lateral extent are very
similar between tomographic models. Relative amplitude strength differs. The slab has a clear dip
trend to the NE.

who interpreted the transition of the Hatteras to the Mississippi slab to have happened when the
South Farallon trench stepped westward after accretion of Shatsky Rise Conjugate plateau. This
led to a slab window with Sonora volcanism including the Tarahumara ignimbrite province (85-
65 Myr ago). We have adopted this as the start of subduction of the Mississippi slab. In the upper
mantle, the slab disintegrates into smaller, dispersed fragments, which have been associated with
subduction during the Laramide orogeny (80-40 Ma) (van der Lee and Nolet, 1997; Sigloch et al.,
2008; Liu and Stegman, 2011). We have taken the end of the Laramide orogeny (40 Ma) as end of
subduction of the slab.

5.62. Mongolia - Mg
The Mongolia anomaly (Figure 5.66) is located below northeast Asia within the mid-mantle. In
previous studies it was interpreted as a Pacific slab (Van der Voo et al., 1999a). In van der Meer
et al. (2010), we used the Kamchatka-Kuriles and Manchuria slabs (shallower in the mantle,
section 5.48 and 5.54) and Mongol-Okhotsk slab (deeper in the mantle, section 5.63) to infer
that the Mongolia anomaly represents (westward) subducted Pacific lithosphere with a Middle to
Late Mesozoic subduction age range. Nokleberg et al. (2000) and Golonka et al. (2003) inferred
westward subduction of Pacific lithosphere at the north-eastern margin of Asia associated with
the continental margin Khingan arc. This arc consists of mainly Barremian (130.8-126.3 Ma) to
Cenomanian (100.5-93.9 Ma) andesites and basalts and related intrusives (Nokleberg et al., 2000),
which we previously used to date the slab. Further south in the Songliao and neighbouring basins
(Ren et al., 2002) documented an earlier start of volcanism, between 155-140 Ma. We therefore
now adopt a somewhat wider age range of subduction with onset and end at 155-126.3 Ma and
100.5-93.9 Ma, respectively.

131
Figure 5.66. Mongolia anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with a similar lateral extent. Relative amplitude strength
and vertical extent differ between tomographic models.

5.63. Mongol-Okhotsk - MO
The Mongol-Okhotsk anomaly (Figure 5.67) is located below northern Siberia from the core-
mantle boundary up to the mid-mantle. The anomaly has been interpreted to represent lithosphere
of the Mongol-Okhotsk ocean, which has been inferred to have subducted until the latest Jurassic-
earliest Cretaceous, based on paleomagnetic and geological data (Van der Voo et al., 1999a; 2015).
The base of the slab merges with the graveyard of slabs under Asia at the base of the mantle (Van
der Voo et al., 1999a). The southern part of the Z-shaped Mongol-Okhotsk slab sensu (Van
der Voo et al., 1999a; 2015) was redefined by van der Meer et al. (2010) as the Central China
slab (section 5.27) and may (also) consist of Paleotethyan lithosphere, as suggested by Stampfli
and Borel (2004). Subduction of and within the Mongol-Okhotsk ocean started well before the
Mesozoic (Tomurtogoo et al., 2005; Donskaya et al., 2013) and pre-Mesozoic lithosphere may
not be visible any longer in the slab graveyard above the core-mantle boundary. Van der Voo et al.
(2015) noted that the deepest part of the Mongol-Okhotsk slab is trending N-S, consistent with
Triassic (~250-220 Ma) orientations of the Mongol-Okhotsk subduction zone, whereas upwards
the slab kinks and eventually becomes oriented ~W-E, similar to the trend of the Mongol-Okhotsk
suture zone. These authors argued that the shape of the slab is consistent with an oroclinal closure
of the Mongol-Okhotsk Ocean suggested by the geological structure of Mongolia and southern
Siberia. The shape of the slab may thus suggest a ~250-220 Ma age of the deepest, N-S striking part
of the slab (below ~2000 km). Closure of the Mongol-Okhotsk ocean follows from paleomagnetic
constraints from the North China and Amurian blocks and Siberia, and is slightly younger than we
previously assumed: 140±10 Ma (Cogne et al., 2005; Van der Voo et al., 2015).
Recently Shephard et al. (2014) performed mantle convection modelling, suggesting that the
Mongol-Okhotsk slab underwent as much as 60 longitude degrees westward motion through
the mantle during its descent to the core-mantle boundary, and should thus be found not further

132
Figure 5.67. Mongol-Okhotsk anomaly. Legend same as figure 5.5. Positive anomalies are
identified in the same location in both tomographic models with a similar vertical extent. Relative
amplitude strength and lateral extent differ between tomographic models. The slab appears to be a
pile on top of the CMB.

east than 35°E, instead of 60-100°E preferred here, and in Van der Voo et al. (1999a; 2015). Our
correlation philosophy, also used in van der Meer et al. (2010; 2012) in linking slabs to their
geological record assumes that slabs do not significantly move laterally relative to each other which,
when viewed globally, implies a preference for slab remnants to sink vertically as was recently
corroborated by Domeier et al. (2016). In other words, the modern distribution of slab remnants
in the deep mantle can be associated with the paleo-subduction zone configuration, which was
found to lead to coherent correlations of plate reconstructions with mantle structure for the last
250-300 Ma (van der Meer et al., 2010; 2012; 2014). The mantle convection model of Shephard
et al. (2014), and similarly Fritzell et al. (2016), predicts that the Mongol-Okhotsk slab has moved
>4000 km westward through the lower mantle since it detached at ~140 Ma (i.e., at ~3 cm/year,
about twice as fast as the sinking rate) relative to the Aegean slab (section 5.1). What is driving
large lateral mantle “winds” in the mantle flow modelling is not known, however, so far there has
been no supporting evidence that such large mantle winds actually exist. Moreover, one would
equally expect that strong mantle winds would deflect or destroy mantle plumes, which then would
question the basic premise of hotspot-based absolute plate motion models that are used to drive
such “slab-prediction” modelling. There is compelling evidence, however, provided by Domeier et
al. (2016), that any lateral mantle flow in the past ~130 Myr did not appreciably perturb the overall
radial sinking of slab and also not appreciably perturb the rising of plumes as is implied by their
use of the Seton et al. (2012) absolute plate motion model. Although we believe that mantle flow
modeling is on the longer term the only viable avenue to quantitatively and dynamically link plate-
tectonic evolution to mantle dynamics, at present the modeling of mantle flow itself is based on
uncertain data and large assumptions, e.g. mantle rheology, and thereby introduces large uncertainty
and complexity to the slab identification problem it tries to solve.

133
Figure 5.68. Nepal anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same lower mantle location in both tomographic models. Relative amplitude strength, vertical and
lateral extent differ between the UU-P07 and S40RTS tomographic models.

5.64. Nepal - Ne
The Nepal anomaly (Figure 5.68) is located in the lower mantle below the Himalaya. Its position
south of the shallower part of the Mongol-Okhotsk slab (section 5.63) and north of the deeper
part of the India slab (section 5.43) suggests that the Nepal anomaly represents lithosphere that
subducted during the mid-Mesozoic, and that its subduction record should be located between
the Himalaya and Mongolia, i.e. in Tibet. The best candidate for its subduction location is the
Bangong-Nujiang suture zone that formed in Early Cretaceous time upon collision of the Lhasa
and Qiangtang terranes of the Tibetan Plateau. All other sutures in Tibet are Triassic or older
in age (Yin and Harrison, 2000; Kapp et al., 2003; Zhu et al., 2013). Van der Voo et al. (2015)
reconstructed the China blocks and Tibet relative to the mantle and portrayed the anomaly here
identified as Nepal slab below the Bangong-Nujiang suture zone in the Early Cretaceous.
Reasonable agreement exists on the age of closure of the Bangong-Nujiang suture zone. The suture
zone contains radiolarian cherts that range in age from mid Jurassic to late Early Cretaceous
(130-120 Ma), flysch deposits of late Early Cretaceous age, and unconformably covering upper
Albian-Aptian (~110-100 Ma) shallow marine limestones (Baxter et al., 2010; Fan et al., 2015).
Calc-alkaline and adakitic magmatic rocks of 115 and 110-100 Ma were interpreted to reflect slab
roll-back and break-off, respectively (Chen et al., 2014; Wu et al., 2014; 2015). The age of the top
of the Nepal slab can thus be dated at ~110-100 Ma. Arc magmatic rocks on southern Qiangtang
have been reported of ~168 Ma (Liu et al., 2013) and 163 Ma (Li et al., 2014b), and northward
subduction below Qiangtang since late Early Jurassic (~180-170 Ma) was concluded as a result
(Xu et al., 2014). Based on a detrital zircon study, subduction below Qiangtang may have started
as early as 210 Ma (Zeng et al., 2015), consistent with the paleomagnetically constrained onset of
northward drift of the Lhasa terrane relative to Gondwana (Li et al., 2016). We use a 210-170 Ma
age range for the base of the Nepal slab.

134
Figure 5.69. New Britain anomaly. Legend same as figure 5.5. This upper mantle slab is well imaged
in the UU-P07 model, but not in the SL2013 model. The dip of this slab is near-vertical.

An alternative model proposes that the Bangong-Nujiang Ocean did not subduct northwards, but
southwards below the Lhasa terrane, and represented the main Paleozoic Tethyan Ocean. This
model proposes that the Lhasa terrane was still connected to Gondwana in the Triassic and rifted
away after the Bangong-Nujiang ocean started to subduct southward below the north Gondwana
margin, opening the Neotethyan ocean as a back-arc from Triassic to Early Cretaceous time (Zhu
et al., 2011; 2013). This scenario implies that the slab below the Lhasa terrane rolled back over a
distance equivalent to the final dimension of the Neotethyan ocean, i.e. >7000 km. However, we
find no anomalies in the lower mantle that span such an enormous width that would support such
a scenario.

5.65. New Britain - NB


The New Britain anomaly (Figure 5.69) corresponds to anomaly A3 of Hall and Spakman (2002;
2004) and was previously imaged by Bijwaard et al. (1998) and Wu et al. (2016). It is a W-E
trending anomaly interpreted as a slab that reaches the base of the upper mantle. It is actively
subducting northward along a trench south of New Britain, consuming Solomon Sea plate
lithosphere. Wu et al. (2016) separated the New Britain slab into two slabs: Solomon Sea West
and Solomon Sea East. Subduction along the New Britain trench is interpreted to be a delayed
response to the collision of the Ontong Java plateau with the original Melanesian subduction zone
along which (part of ) the Papua slab (section 5.70) subducted below the Australian plate until
approximately 26-20 Ma (Hall, 2002; Quarles van Ufford and Cloos, 2005; Knesel et al., 2008;
Holm et al., 2013). The arrest of subduction of the Papua slab along the Melanesian trench was first
followed by subduction of the Solomon Sea plate along the Trobriand trough (Hall, 2002; Quarles
van Ufford and Cloos, 2005; Schellart and Spakman, 2006; Knesel et al., 2008) (Welford slab, see
section 5.92), after which northward subduction started below New Britain creating the Bismarck

135
Figure 5.70. New Hebrides anomaly. Legend same as figure 5.5. This upper mantle slab is well
imaged in the UU-P07 model and there is a similar relative amplitude strength, lateral extent and
vertical extent between models. This slab dips to the NE.

arc since 10-5 Ma (Hall, 2002; Holm et al., 2013) or since 15 Ma (Wu et al, 2016). We adopt
15-10 the age range of the base of the New Britain slab.

5.66. New Hebrides - NH


The New Hebrides anomaly (Figure 5.70) is located west of the northern termination of the Tonga
trench, and is interpreted as lithosphere still subducting at the New Hebrides subduction zone. It is
a NW-SE trending slab that was imaged before by Fukao et al. (2001), Hall and Spakman (2002;
2004), Schellart and Spakman (2012), Fukao and Obayashi (2013), Obayashi et al. (2013), and Wu
et al. (2016). It reaches the base of the upper mantle. The NE-dipping New Hebrides subduction
zone is interpreted to result from subduction polarity reversal due to the arrival of the Ontong-Java
plateau into an originally contiguous Solomon-Vitiaz-Tonga-Kermadec subduction zone, around
15-10 Ma (Hall, 2002; Sdrolias et al., 2004; Schellart et al., 2006; Seton et al., 2012), which we
adopt for the age of the base of the slab.

136
Figure 5.71. North Apennines anomaly. Legend same as figure 5.5. This shallow slab is well imaged
in the UU-P07 model, but not in the SL2013 model. This slab has a dip trend to the SW.

5.67. North Apennines - NA


The North Apennines anomaly (Figure 5.71) is located below Corsica and Tuscany in the upper
mantle. Based on early tomographic models, Spakman et al. (1993) interpreted the North
Apennines anomaly as a detached slab, but subsequent models rather suggested that it is still
connected to the surface (Piromallo and Morelli, 1997; 2003; Bijwaard et al., 1998; Lucente et
al., 1999; Spakman and Wortel, 2004). The North Apennines slab is clearly separated from the
Calabria slab by a several hundred kilometres wide gap, interpreted as the result of a lateral tear
(Wortel and Spakman, 1992; 2000) or a subduction transform fault (Rosenbaum et al., 2008).
Lucente et al. (1999) and Lucente and Speranza (2001) interpreted the North Apennines slab as
contiguous with an anomaly lying in and below the transition zone (the Alps slab, section 5.6) by,
which would result in a 700 km long slab. Spakman and Wortel (2004), however, suggested that the
North Apennines slab reaches a depth of only ~300 km, and is a separate body from the Alps slab.
Geological reconstructions of the northern Mediterranean region consistently conclude that the
southward subduction zone of the Alps reached as far south as Corsica, based on the finding of
a (currently) west-verging thin-skinned fold-thrust belt including high-pressure metamorphic
rocks on Corsica (Brunet et al., 2000a). A transition from east- to west-dipping subduction below
Corsica, coinciding with the start of the North Apennines subduction zone, is generally estimated
to have occurred around 35-30 Ma (Rosenbaum et al., 2002a; Jolivet et al., 2009; Argnani, 2012;
Advokaat et al., 2014b). Because the North Apennines slab is oriented at a high angle to the
Africa-Europe convergence direction, almost all of its length should have been accommodated by
back-arc extension in the northwest Mediterranean, which at the latitude of the North Apennines
slab is on the order of 300 km (Faccenna et al., 2001; Faccenna et al., 2004). We therefore follow
the interpretation of Spakman and Wortel (2004), and assign an age to the base of the North
Apennine slab of 35-30 Ma.

137
Figure 5.72. North Pacific anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both tomographic models with a similar vertical extent. Relative amplitude
strength and lateral extent differ between tomographic models.

5.68. North Pacific - NP


The North Pacific anomaly (Figure 5.72) is located below the northern Pacific Ocean and southern
Alaska from the mid-mantle up to upper part of the lower mantle. In previous tomographic
studies it has been interpreted as Kula (Qi et al., 2007),Pacific slab (Ren et al., 2007) and was also
identified as ‘K’ in Sigloch (2011). The Aleutian slab (section 5.3) higher in the mantle to the north
that subducted since 56-46 Ma led us to infer that the North Pacific anomaly represents Early
or pre-Cenozoic subducted lithosphere (van der Meer et al., 2010). Shapiro and Solov’ev (2009)
inferred northward subduction of Kula oceanic lithosphere below a southward extension of the
North American plate, resulting in the intra-oceanic Kronotsky-Commander arc now found in
Kamchatka. This arc was dated by Levashova et al. (2000) to be active from 73±7 Ma (Campanian-
Maastrichtian) to 40 ± 2 Ma (Eocene), perhaps with a switch in subduction polarity in the
Paleocene, resulting in subduction of North American plate.
The reconstruction of Shapiro and Solov’ev (2009), who restored the Kronotksy-Commander arc
as part of the Pacific plate after its Eocene cessation, would place the arc in a positon that is in
agreement with the location of the North Pacific slab. We therefore follow their interpretation
and surpass our previous interpretation (van der Meer et al., 2010) which was based on an older
reconstruction (Engebretson et al., 1985) of the paleo-Pacific Ocean. Following its extinction the
Kronotsky-Commander arc was transported westward as an aseismic ridge to the Eurasian plate
and collided with Kamchatka at ~5 Ma (Shapiro and Solov’ev, 2009).

5.69. Pamir - Pa
Subduction in the Pamir (Figure 5.73) was first demonstrated by Burtman and Molnar (1993),
and became evident from deep seismicity down to 250 km depth (Pegler and Das, 1998; Pavlis
and Das, 2000; Koulakov, 2011; Zhang et al., 2011; Schneider et al., 2013; Sippl et al., 2013). The

138
Figure 5.73. Pamir anomaly. Legend same as figure 5.5. This upper mantle slab is well imaged in the
UU-P07 model and there is a similar relative amplitude strength, lateral extent and vertical extent
between the UU-P07 and SL2013 models. This slab dips to the SW.

Pamir anomaly is interpreted to contain Asian lithosphere that subducted southward below the
Pamir salient, which is a half-orocline associated with major counterclockwise rotations in the west,
and a transform fault in the east – the Kashgar-Yecheng transform system (Bourgeois et al., 1997;
Cowgill, 2010; van Hinsbergen et al., 2011a; Sobel et al., 2013). Seismic tomographic images of
the Pamir slab reveal that its base reaches down to ~300-400 km depth (Van der Voo et al., 1999b;
Negredo et al., 2007) (Figure 5.77). The nature of the lithosphere of the Pamir slab is difficult to
establish. The Pamir subduction zone is not associated with accretion of sediments, but instead with
subduction erosion of the overriding plate (Sobel et al., 2013). Because the Pamir subduction zone
is currently entirely intracontinental, the Pamir slab is generally interpreted as continental Asian
lithosphere, even though it seems to behave in an oceanic manner, dipping near-vertical into the
mantle (Sobel et al., 2013). Jackson et al. (2002) suggested that the Pamir slab may have represented
a trapped basin of strongly attenuated continental or perhaps oceanic crust similar to the South
Caspian Sea.
Van Hinsbergen et al. (2011a) kinematically restored the Pamir salient and showed that the ~300-
400 km of Pamir subduction is consistent with the northward motion of the northern Pamir
relative to the adjacent Tarim basin, i.e. ~370 km based on estimated displacements in the western
Kunlun Shan fold-thrust belt and along the Kashgar-Yecheng strike-slip system of Cowgill et al.
(2003) and Cowgill (2010). Cowgill (2010) estimated the onset of activity of these structures at
30±7 Ma. Sobel et al. (2013), using thermochronological data from the northern Pamir, estimated
an onset of subduction around 25 Ma, at the young end of the spectrum. We conservatively keep
the wide age range of Cowgill (2010) to date the base of the Pamir slab.

139
Figure 5.74. Papua anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both tomographic models with a similar vertical extent and relative amplitude.
Lateral extent is not clearly marked in the tomographic models. The slab is flat-lying on top of the
transition zone.

5.70. Papua - Pu
The Papua anomaly (Figure 5.74) corresponds to the northern part of the A7 anomaly of Hall and
Spakman (2002; 2004) and underlies an area from Papua New Guinea to the New Hebrides and
from the east Australian margin to the Solomon islands, flat-lying at the top of the lower and base
of the upper mantle. At its top it is south of the base of the New Hebrides (section 5.66) and
New Britain (section 5.67) slabs that subducted since ~10-15 Ma. Hall and Spakman (2002; 2004)
noted that the Papua anomaly is not everywhere well-defined and may in fact represent more than
one slab, whereby the southwesternmost part of the slab below NW Australia (the Welford slab,
section 5.92) may be disconnected from the northern and eastern parts of the anomaly, which is
the interpretation we follow here to account for the subduction records of both the Melanesian and
Trobriand troughs. Hall and Spakman (2002; 2004) interpret the northern segment of the anomaly
from Papua New Guinea to the New Hebrides to result from subduction at the Melanesian arc,
from 45 Ma until the collision with the Ontong Java plateau, ~25 Ma, based on the reconstruction
of Hall (2002). We adopt a somewhat large age range of 26-20 Ma for the collision of the Ontong
Java plateau based on published estimates of Hall (2002), Quarles van Ufford and Cloos (2005),
Knesel et al. (2008), and Holm et al. (2013) as estimate for the age of the top of the slab. Hall
(2002), Cloos (2005) and Gaina and Müller (2007) interpreted the onset of subduction below the
Melanesian arc to occur ~45 Ma ago, and Wu et al. (2016) since 50 Ma. Schellart et al. (2006)
interpreted westward subduction below the Melanesian arc to have been ongoing since at least 90
Ma. We therefore adopt a large, 90-45 Ma range for the age of the base of the slab.

140
Figure 5.75. Reggane anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both the UU-P07 and S40RTS tomographic models with a similar vertical extent
and relative amplitude. Lateral extent differs. The slab does not have dip.

5.71. Reggane - Re
The Reggane anomaly (Figure 5.75) is a small anomaly located in the upper part of the lower
mantle below the western Sahara in the lower mantle and is resolved in both the UU-P07 and
S40RTS models. It was previously identified by Vissers et al. (2016). The Algeria slab directly to
the east, which we correlate with Early Cretaceous ophiolite emplacement over Adria (section 5.4)
suggests that the Reggane anomaly likely represents lithosphere that subducted sometime in the
Late Jurassic or Early Cretaceous in the westernmost Mediterranean region.
An Early Cretaceous phase of subduction between Iberia and southern France was postulated by
Sibuet et al. (2004) based on reconstructions of the Bay of Biscay and Central Atlantic Ocean.
Vissers and Meijer (2012a; b) revisited the ocean basin reconstructions and reconciled these with
the shortening history of the Pyrenees. They confirmed that ocean basin reconstructions require
~500 km of Early Cretaceous convergence across the Pyrenees, preceding intra-continental
shortening, and advocated a short phase of subduction accommodating the counterclockwise
rotation of Iberia. Paleomagnetic data confirm the ~35-40° of counterclockwise rotation of
Iberia relative to Eurasia predicted from ocean basin reconstructions and constrain the timing of
rotation, and hence Pyrenean subduction, to ~125-110 Ma (Gong et al., 2008; Ruiz-Martinez et
al., 2012). Vissers and Meijer (2012b) interpreted a phase of high-temperature metamorphism
that affected the central Pyrenees around 110-100 Ma as the result of slab break-off. Vissers et al.
(2016) therefore dated the base and top of the Reggane slab at 126-121 Ma, corresponding to the
varying ages assigned to the M0 isochron, and 105±5 Ma, respectively. On the basis of refuting the
paleomagnetic database of Vissers et al. and claiming the tomographic model used was inconclusive,
Barnett-Moore et al. (2016) stated the Reggane slab was the result of enigmatic western Neotethys
subduction instead of of Pyrenean subduction. Van Hinsbergen et al. (2017) showed that the
Reggane anomaly appears in 8 tomographic models, including the ones used by Barnett-Moore

141
Figure 5.76. Rio Negro anomaly. Legend same as figure 5.5. This upper mantle slab is well imaged
in the UU-P07 model and there is a similar relative amplitude strength, lateral extent and vertical
extent between the UU-P07 and SL2013 models. This slab dips to the NE.

et al. (2016), and showed that their criticism of Barnett-Moore et al. (2016) on the paleomagnetic
database that requires Pyrenean subduction is unfounded.

5.72. Rio Negro - RN


The Rio Negro anomaly (Figure 5.76) is located in the upper mantle below western Patagonia,
southern South America. It is N-S trending and partly flat-lying. It has previously been interpreted
to represent still-subducting Nazca lithosphere below the Andean margin (Aragón et al., 2011). We
interpret a break in magmatism in the overriding plate between ~30-15 Ma to correspond with the
gap between this slab and the deeper San Matias slab (section 5.77). Aragón et al. (2011) estimated
that subduction was re-established by ~23 Ma ago. Arc magmatism was re-established at around 15
Ma (Munizaga et al., 2002). We adopt the 23-15 Ma age range date the base of the slab.

142
Figure 5.77. Rockall anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both the UU-P07 and S40RTS tomographic models. Lateral, vertical and relative
amplitude differ. The slab does not have dip.

5.73. Rockall - Ro
The Rockall anomaly (Figure 5.77) is located west of the British Isles under the eastern North
Atlantic Ocean and lies on the core-mantle boundary. It was first identified in van der Meer et al.
(2010). The Mongol-Okhotsk slab (section 5.63) to the east, the Wichita slab (section 5.91) to the
west and the Balkan slab (section 5.12) to the southeast were used by van der Meer et al. (2010) to
interpret the anomaly as the Rockall slab that resulted from Early Mesozoic subduction between
Siberia and Laurussia in the paleo-Arctic Ocean. The paleogeographic location is similar to the
Upper Triassic-Lower Cretaceous continental margin Kony Murgal arc (Nokleberg et al., 2000).
Shephard et al. (2013) renamed this to the Koni-Taigonous arc and narrowed its age range down to
(Late) Triassic-Late Jurassic, which we adopt as age range for the Rockall slab.

143
Figure 5.78. Ryukyu anomaly. Legend same as figure 5.5. This upper mantle slab is well imaged
in the UU-P07 model and there is a similar relative amplitude strength, lateral extent and vertical
extent between the UU-P07 and SL2013 models. This slab dips to the NW.

5.74. Ryukyu - Ry
The Ryukyu anomaly (Figure 5.78) is located in eastern Asia below the East China Sea in the upper
mantle. The Ryukyu slab has previously been imaged in numerous regional and global tomographic
models (Bijwaard et al. 1998; Lallemand et al., 2001; Zhao and Ohtani, 2009; Zheng et al., 2013;
Koulakov et al., 2014; Wu et al., 2016). The slab is connected to the surface at the Ryukyu arc-
Okinawa trench subduction zone, and is interpreted as Philippine Sea plate lithosphere that is still
subducting northward below the South China block of the Eurasian plate. Towards the northeast it
is interacts with the Izu-Bonin and Manchuria ( Japan) slabs (sections 5.44 and 5.54, respectively)
at a trench-trench-trench triple junction. Towards the southwest, the trench is propagating
westwards south of the island of Taiwan, where subduction of the southeastward dipping Manila
slab (section 5.55) is coming to a halt due to collision of the Philippine Sea plate with the South
China block (e.g., Ustaszewski et al., 2012). According to the kinematic restoration of SE Asia of
Hall (2002), subduction along the Ryukyu trench started at ~25 Ma, whereas in the reconstruction
of Wu et al. 2016, subduction started at ~15 Ma. In the central Ryukyu arc, volcanism started at ~21
Ma (Chung et al., 2000), and we adopt a 25-21 Ma age range for the base of the Ryukyu slab.

144
Figure 5.79. Sakhalin anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both the UU-P07 and S40RTS tomographic models. Lateral, vertical extent and
relative amplitude differ. The slab does not have dip.

5.75. Sakhalin - Sa
The Sakhalin anomaly (Figure 5.79) is located below northeast Asia within the upper part of the
lower mantle. In previous tomographic studies it has been interpreted to as a subducted slab of the
Okhotsk plate (Gorbatov et al., 2000). On the basis of the Kamchatka slab to the east and shallower
in the mantle, van der Meer et al. (2010) inferred that the Sakhalin slab likely subducted in Late
Mesozoic to Early Cenozoic times. In the tectonic model of Nokleberg et al. (2000) subduction
of Okhotsk lithosphere initiated at the continental margin below the East Sikhote-Alin arc in the
Late Cretaceous. The Sikhote arc consists of Cenomanian-Danian arc volcanics, which we adopt
as the age range for the Sakhalin slab. Subduction ended in the Late-Cretaceous-Paleocene due to
accretion of the Okhotsk block, resulting in an eastward jump to form the Kamchatka-Kuriles arc
and slab (section 5.47).

145
Figure 5.80. Sangihe anomaly. Legend same as figure 5.5. This shallow slab is well imaged in the
UU-P07 model, but not in the SL2013 model. This slab has a dip trend to the NW.

5.76. Sangihe - Sn
The Sangihe anomaly (Figure 5.80) was imaged by Widiyantoro and van der Hilst (1997), Bijwaard
et al. (1998), Rangin et al. (1999), and Hall and Spakman (2002; 2004; 2015), and Wu et al. (2016)
and is interpreted as one of the two actively subducting slabs attached to the Molucca Sea plate,
the other one being the Halmahera slab (section 5.36). The Sangihe slab dips to the NW below the
island of Sulawesi and the Sangihe arc and reaches the base of the upper mantle. At shallow upper
mantle levels the slab is separated into several slabs (Hall & Spakman 2015, Wu et al. 2016), from
north to south including the Philippine Trench slab, Molucca Sea West slab, Sulu and Celebes
Sea South slab (Wu et al., 2016). The deepest part of the Sangihe slab, named Molucca Sea West
slab in Wu et al. (2016) possibly penetrates into the lower mantle, where it merges with the large
Kalimantan anomaly (section 5.48). In the reconstruction of Hall (2002), subduction below the
Sangihe arc started ~25 Ma, whereas the reconstruction of Wu et al. 2016 assume a start at ~30 Ma.
We therefore use 30-25 Ma as age for the base of the slab.

146
Figure 5.81. San Matias anomaly. Legend same as figure 5.5. Positive anomalies are identified in
approximate locations in both the UU-P07 and SL2013-S40RTS tomographic models. Lateral,
vertical extent and relative amplitude differ. In the UU-P07 model, the slab is either flat-lying at
the transition zone, or may have a dip to the SE.

5.77. San Matias - SM


The San Matias anomaly (Figure 5.81) is located below southeastern South America, offshore
Patagonia, at the top lower mantle and at the base of the upper mantle. It is a N-S trending
anomaly and has previously been interpreted to have subducted eastward below South America,
followed by its detachment due to the arrival of the Farallon-Aluk ridge at the Andean margin
(Aragón et al., 2011). This led to a break in magmatism between ~30-15 Ma, which we interpret to
correspond with the slab window between this slab and the shallower Rio Negro slab. We interpret
the onset of subduction of the slab to correspond with the Paleocene–Eocene Pilcaniyeu belt with
an age of 68–49 Ma (Giacosa and Heredia, 2004).

147
Figure 5.82. Sao Francisco anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both the UU-P07 and S40RTS tomographic models. Lateral, vertical extent
and relative amplitude are similar. The slab is vertical in the deep mantle and seems to be piling up
at the CMB.

5.78. Sao Francisco - SF


The Sao Francisco anomaly (Figure 5.82) is located below eastern South America, within the
deepest mantle and resting on the core-mantle boundary. It is NE-SW trending at the core-
mantle-boundary and changes to N-S trending at shallower levels. It is disconnected from and lies
eastward of the Brasilia slab (section 5.18). At the CMB, it connects northward with the Atlantis
slab (section 5.11) suggesting a possible common origin. As we interpreted the Atlantis slab to
result from Triassic Panthalassa Ocean subduction along southwestern Laurentia, we infer the
Sao Francisco slab to have formed at the northwestern Gondwana (South American) continental
margin. At this location, which would fit the NE-SW slab trend, Triassic metamorphism (243-225
Ma) affected granites, basalts and sedimentary rocks in the Tahamí Terrane, interpreted to have
taken place in an Andean-type orogeny on the western side of Pangea as part of an orogenic belt
(Restrepo et al., 2011). We connect this record to the subduction of the Sao Francisco slab and use
it to date the Sao Francisco slab.

5.79. Scotia - Sc
The Scotia anomaly (Figure 5.83) was first documented in the tomographic model of Bijwaard et al.
(1998). It is interpreted to represent still-subducting South American plate lithosphere consumed
at the South Sandwich subduction zone, and dips westward below the Scotia Sea. Resolution tests
show that the deeper part of the anomaly is poorly resolved, and our interpretation is therefore
tentative. The cause of formation of the Scotia subduction zone is puzzling. It formed within the
South American plate, south of the Falkland plateau, even though the South American plate and
Antarctic plate were in an extensional phase at that time and the spreading between South America
and Antarctica was being accommodated at the mid-oceanic ridge in the Weddell Sea. Because

148
Figure 5.83. Scotia anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both the UU-P07 and S40RTS tomographic models in the lower mantle, but
differ considerably in the upper mantle between the UU-P07 and SL2013 model. In the lower
mantle, lateral, vertical extent and relative amplitude are similar. The slab is dipping to the west, and
is flat-lying at the top of the lower mantle.

subduction of the Scotia slab was for most of its history not associated with plate convergence
but is fully accommodated by extension in the Scotia and South Sandwich basins, the age for
the onset of subduction is best estimated from the age of the onset of extension in the overriding
plate. Oceanic lithosphere in the South Sandwich back-arc is Late Miocene and younger, and in
the Scotia Sea is interpreted as ~27 Ma and younger (Eagles, 2005; Lodolo et al., 2006). Closing
these basins juxtaposes continental fragments along the southern fringes of the Scotia Sea (e.g., the
South Orkney microcontinent) with the southern margin of the Falkland Plateau, suggesting that
basin formation started within the Mesozoic passive margin of the Falkland plateau that formed
during the breakup of South America and Antarctica in the Late Jurassic or Early Cretaceous.
This suggests that oceanization in the Scotia Sea at 27 Ma was preceded by a period of pre-drift
extension. Estimates for the onset of extension in this area from Tierra de Fuego is Eocene (~55-45
Ma) (Ghiglione et al., 2008), by which time subduction of the Scotia slab must have been active
(Eagles et al., 2006; Livermore et al., 2007; Dalziel et al., 2013; Nerlich et al., 2013). This episode
was preceded by a phase of slow convergence between West Antarctica and Patagonia concluded
from marine magnetic anomalies of the Atlantic and Antarctic oceans, which likely date the onset
of subduction (e.g., König and Jokat, 2006; Eagles, 2010; Eagles and Jokat, 2014). We adopt an
80-55 Ma age range for the base of the Scotia slab. The anomalies in the Weddell Sea become
younger towards the north, and are youngest at the fault zone that bounds the South Orkney
microcontinent from the Weddell Sea, where they are ~10 Ma old (Müller et al., 2008). This may
indicate that the originally E-W trending portion of the Scotia subduction zone to the south of
the South Orkney microcontinent became inactive due to the migration of the subduction zone to

149
Figure 5.84. Sisimut anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both the UU-P07 and S40RTS tomographic models. Lateral, vertical extent difer
and relative amplitude is similar. The shape of the slab is unclear in UU-P07, but appears to be
dipping to the south in the S40-RTS model.

the Weddell Sea ridge and consequent slab break-off, leaving the west-dipping South Sandwich
subduction zone as the only active portion of the system.

5.80. Sisimut - Si
The Sisimut anomaly (Figure 5.84) is located below northernmost North America and Greenland
within the mid-mantle. It is NW-SE trending and does not have a neighbouring slab in close
vicinity. Van der Meer et al. (2012) suggested that the slab may result from paleo-Arctic subduction
and loosely inferred the anomaly we now define as Sisimut slab to be of Early Cretaceous age.
It was recently studied in detail by Shephard et al. (2016) who named it the Greenland anomaly.
Within the region of the Sisimut slab, Nokleberg et al., 2000) documented southward subduction
of Angayucham and Anyui lithosphere below the continental-margin Nutesyn arc. This arc is
Late Jurassic to Early Cretaceous in age (Nokleberg et al., 2000; Parfenov et al., 2009). In the
plate motion model of Shephard et al. (2013), these different elements were integrated and they
concluded that subduction below the Koyukuk and Nutesyn arcs was active between 160-120 Ma,
which we adopt for ages for the top and bottom of the slab, in line with Shephard et al. (2016).

150
Figure 5.85. Sistan anomaly. Legend same as figure 5.5. Vertical and lateral extent are very similar
between tomographic models. Relative amplitude strength differs and is stronger in UU-P07. The
slab has no dip trend.

5.81. Sistan - St
The Sistan anomaly (Figure 5.85) is a N-S striking anomaly below eastern Iran and western
Afghanistan in the upper part of the lower mantle. Given its orientation, its position to the west
of the India slab (section 5.43), northeast of the Mesopotamia slab (section 5.60) and northwest
of the Carlsberg slab (section 5.22), the Sistan anomaly is best explained by subduction that
terminated with the formation of the Sistan suture that currently still overlies this slab. The Sistan
suture is N-S trending and separates the Lut Block of central Iran and the Helmand Block of
Afghanistan (Camp and Griffis, 1982; Tirrul et al., 1983). This suture formed at the expense of an
Early Cretaceous oceanic basin (Babazadeh and De Wever, 2004) that was in an overriding plate
position relative to the Mesopotamia slab (section 5.60). Eclogite and blueschist rocks in melanges
of the Sistan suture (Fotoohi Rad et al., 2005) were originally suggested to have formed in Early
Cretaceous time (~125 Ma) based on 40Ar/39Ar thermochronology with large age uncertainties
of more than 10 Myr (Fotoohi Rad et al., 2009). A recent re-dating of these rocks by 40Ar/39Ar,
Rb/Sr and U/Pb geothermochronology was not able to reproduce these Early Cretaceous ages and
rather produced a tight clustering of ages between 83 and 89 Ma (Bröcker et al., 2013), by which
time subduction must have been underway to produce pressures in excess of 20 kbar (Fotoohi Rad
et al., 2005). Given the stratigraphic ages of Albian-Aptian in the HP mélange of the Sistan suture
(~125-100 Ma) we adopt an onset of subduction of the Sistan Ocean of 100-90 Ma. An ~59 Ma
granite in the Sistan suture has a geochemistry consistent with arc magmatism (Delavari et al.,
2014), ~46-25 Ma magmatism in the area is interpreted as post-collisional, perhaps reflecting
delamination and asthenospheric inflow that may have been triggered by slab break-off (Rezaei-
Kahkhaei et al., 2010; Pang et al., 2013; Mohammadi et al., 2016). We adopt a 59-46 Ma end of
the subduction of the Sistan slab.

151
Figure 5.86. Socorro anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same location in both the UU-P07 and S40RTS tomographic models. Lateral extent is similar, but
vertical extent and relative amplitude differ. The dip of the slab is to the SW in UU-P07.

5.82. Socorro - So
The Socorro anomaly (Figure 5.86) is located below western North America within the mid-
mantle. It was first described in van der Meer et al. (2010) and corresponds to anomaly CR-3
of Sigloch and Mihalynuk (2013). The Idaho slab (section 5.42) and the Socorro anomaly cover
similar depth ranges and dips, and have a more westerly location than the Hatteras and Cocos slabs
of the Farallon family (sections 5.37 and 5.29, respectively). Similar to the Idaho slab we therefore
interpreted an intra-oceanic, Jurassic-Cretaceous origin for this slab. The Wrangellia superterrane of
the northwestern North American margin contains terranes derived from both northerly and near-
equatorial paleolatitudes according to paleomagnetic data (Nokleberg et al., 2000). In conjunction
with our interpretation of the Idaho slab, we associate eastward subduction of Panthalassa/Farallon
lithosphere below the southern part of the Wrangellia terranes to have formed the Socorro slab,
starting in the Early to Middle Jurassic. Accretion of the terrane to Laurentia started in the Late
Jurassic (Trop et al., 2005) and ended in the Early to Late Cretaceous (Nokleberg et al., 2000).
Sigloch and Mihalynuk (2013) inferred that the Socorro slab was overridden by the continental
margin at 55-50 Myr, culminating in the Siletzia and Pacific Rim terrane accretions, which we
adopt as the age of breakoff.

152
Figure 5.87. South Loyalty Basin anomaly. Legend same as figure 5.5. Positive anomalies are
identified in the same location in both the UU-P07 and S40RTS tomographic models. Lateral
extent is similar, but vertical extent and relative amplitude differ. There is no clear dip of the slab.

5.83. South Loyalty Basin - SLB


The NW-SE trending South Loyalty Basin anomaly (Figure 5.87) is located below the Tasman Sea
within the upper part of the lower mantle. It was first identified by Schellart et al. (2009). In the
east, the flat-lying South Loyalty Basin anomaly is difficult to distinguish from the steeply dipping
Tonga-Kermadec-Hikurangi anomaly (section 5.87), which makes estimating the depth of its base
difficult. We follow the suggestion of Schellart et al (2009) and Schellart and Spakman (2012), and
assign a 1000 to 1200 km depth range for the anomaly, which we interpret as the South Loyalty
Basin slab.
Its position west of the Tonga-Kermadec-Hikurangi slab (section 5.87) requires a subduction
location between the Tonga-Kermadec trench and Australia, where the geological record of the
New Caledonia subduction zone (Schellart et al., 2009) is the only candidate for correlation.
Schellart et al. (2006) estimated that this subduction zone was active from 50-45 Ma until 30-20
Ma. Matthews et al. (2015) suggested that subduction halted earlier, 45 Ma. The New Caledonia
subduction zone terminated with the obduction of an ophiolite derived from the overriding plate
forearc over a Paleozoic to Mesozoic volcano-sedimentary complex belonging to the Gondwana-
derived Zealandia microcontinent (Luyendyk, 1995; Bache et al., 2013) on the island of New
Caledonia. The ophiolite is underlain by high pressure, high-temperature amphibolites interpreted
as a metamorphic sole, which was dated with 40Ar/39Ar on hornblende and U/Pb dating of
zircons at ~56 Ma (Cluzel et al., 2012); sheeted dykes in the New Caledonia ophiolite have 53 Ma
ages (Cluzel et al., 2006). This shows a somewhat earlier onset of New Caledonia subduction than
previously interpreted, likely a few million years before the ages derived from the sole. We adopt a
60-56 Ma subduction initiation age and a 45-30 Ma break-off age in conjunction with the recent
kinematic restoration of van de Lagemaat et al. (2017).

153
Figure 5.88. South Orkney Island anomaly. Legend same as figure 5.5. Positive anomalies are
identified in the same location in both the UU-P07 and S40RTS tomographic models. Lateral
extent and relatrive amplitude is similar, but vertical extent differs. There is no clear dip of the slab.

5.84. South Orkney Island - SOI


The South Orkney Island anomaly (Figure 5.88) is NW-SE trending and is located in the mid-
mantle, from below southeastern Patagonia to below the Weddell Sea. Based on the shallower
Scotia slab (section 5.79), it most likely represents paleo-Pacific lithosphere that subducted at the
proto-Andean or Gondwanide margin. Martin (2007) interpreted break-up of the Gondwanide
margin by slab rollback and back-arc basin spreading to occur from the Early-Middle Jurassic
(190-175 Myr) to the Middle Cretaceous. Subduction rollback was terminated by the Palmer Land
tectonic event in West Antarctica from 113 – 103 Ma and by the inversion of the Rocas Verdes/
Magallanes Basin in Patagonia around 94 Ma (Vaughan et al., 2002; Fildani and Hessler, 2005),
which we interpret to represent the end of the subduction of the slab.

5.85. Sunda - Su
The Sunda anomaly (Figure 5.89) was previously imaged by Fukao et al. (1992), Puspito and
Shimazaki (1993), Puspito et al. (1995), Widiyantoro and van der Hilst (1996; 1997), Bijwaard
et al. 1998; Replumaz et al. (2004), Spakman and Hall (2010), Pesicek et al. (2010), Widiyantoro
et al. (2011), Koulakov (2013), Fukao and Obayashi (2013), Zahirovic et al. (2014), Hall and
Spakman (2015), and Wu et al. (2016). It is interpreted as Australian and Indian plate lithosphere
that is actively subducting northward below Sundaland, along the Sumatra and Java trenches. To
the northwest it is disconnected through a slab window below the Andaman Islands from the
Burma slab (section 5.19). In the East it connects to the Banda slab (section 5.13). In the west,
tomographic images show that the Sunda slab reaches the base of the upper mantle, but does not
connect with deeper anomalies. In the east, however, the Sunda slab merges with the SW-NE
trending Kalimantan anomaly (section 5.48) and several authors (e.g. Replumaz et al., 2004;
Zahirovic et al., 2014) suggested that these slabs may be contiguous. The different orientation of

154
Figure 5.89. Sunda anomaly. Legend same as figure 5.5. Positive anomalies are identified in
the same location in both the UU-P07 and S40RTS tomographic models in the lower mantle,
but differ considerably in the upper mantle between the UU-P07 and SL2013 model. In the
lower mantle, lateral, vertical extent and relative amplitude are similar. The slab is dipping to the
northeast, and is flat-lying at the top of the lower mantle.

the Kalimantan slab from the Sunda slab, however, suggests that these slabs result from separate
subduction events instead (Hall & Spakman 2015, Wu et al., 2016). For the Kalimantan anomaly
(section 5.47) the kinematic restoration of Hall (2012) combined with the interpretation of mantle
structure by Hall and Spakman (2015) suggests a NW-ward subduction event below western
Sulawesi from the latest Cretaceous to the middle Eocene. This same reconstruction suggested that
the Sunda trench was in Late Cretaceous to Paleocene time a transform boundary, which became
inverted as a subduction zone ~50-45 Ma ago upon the onset of Australia-Eurasia convergence,
subducting Indo-Australian oceanic crust towards the NE. We adopt this as the age for the base
of the Sunda slab. Disagreement between these reconstructions exists for the eastern extent of the
Sunda slab. This subduction either continued as northward subduction of Indo-Australian oceanic
lithosphere, offset by a dextral transfer zone (Hall and Spakman, 2015) or as southward subduction
of the western part of an East Asian Sea (Wu et al., 2016).

5.86. Telkhinia slabs - Te


A north–south trending belt of positive wave-speed anomalies under the central Pacific
Ocean has been interpreted to represent a series of Triassic-Jurassic intra-oceanic subduction
zones, called the Telkhinia subduction zones by van der Meer et al. (2012) (Figure 5.90). These
anomalies are detected in both the UU-P07 model and S40RTS model in the lower mantle at
depths greater than 1500 km. Three distinct intra-oceanic subduction zones were correlated to
these anomalies. At these positions in the lower mantle seismic scatterers have been detected,
interpreted to be caused by remnants of subducted and folded former oceanic crust under the
central Pacific Ocean by Kaneshima & Helfrich (2010) and Ma et al. (2016) and under the

155
Figure 5.90. Telkhinia anomalies, Legend same as figure 5.5. Positive anomalies are identified in
the same locations in both UU-P07 and S40RTS tomographic models. Lateral, vertical extents and
relative amplitudes differ.

northern Pacific Ocean by Schumacher and Thomas (2016). The tomographic analyses of van der
Meer et al. (2012) and waveform modelling of He and Wen (2009) showed that the top of the
Pacific LLSVP was essentially split as a result of sinking slabs. In both P and S-wave tomographic
models, the amplitudes of the anomalies are weaker than for other slabs identified at equivalent
depths associated with circum-Pangaea subduction zones. This may be explained by the proximity
of the slabs to the hotter LLSVP (van der Meer et al., 2012). Subsequent tomographic studies
corroborated the presence of central-Pacific lower mantle anomalies (Simmons et al., 2012; French
and Romanowicz, 2014, Suzuki et al., 2016) or improved the imaging of the genetically related
intra-oceanic Mendocino slab (section 5.59) to the east (Sigloch and Mihalynuk, 2013).
The Telkhinia slabs were correlated to exotic Triassic to Lower Cretaceous volcanic arcs now
accreted in the far east Asian margin, currently incorporated in the Kolyma-Omolon and Andyr-
Koryak arcs in Siberia, and the Oku-Niikappu arc in northern Japan (van der Meer et al., 2012).
For this correlation, van der Meer et al. (2012) used the sinking rate of slabs of van der Meer et al.
(2010) to interpret the Telkhinia slabs as the result of Early-Mid Mesozoic subduction, consistent
with the age of these exotic arc relics. Paleomagnetic and paleontological constraints from these
arcs provided an indication of their paleolatitude, and van der Meer et al. (2012) inferred their
paleolongitude relative to the mantle by connecting them to the Telkhinia slabs. To avoid circular
reasoning, we have not included the inferred ages of tops and bottoms of the Telkhinia slabs in our
compilation used to constrain sinking rates of slabs in the mantle.

5.87. Tonga-Kermadec-Hikurangi - TKH


The Tonga–Kermadec–Hikurangi anomaly (Figure 5.91) is located below the south Fiji Basin
in the upper mantle and the upper part of the lower mantle up to a depth of ~1200 km. It was
interpreted in the uppermost lower mantle as a single N-S trending slab by van der Hilst (1995),

156
Figure 5.91. Tonga-Kermadec-Hikurangi anomaly. Legend same as figure 5.5. The upper mantle
part of this slab slab is well imaged in the UU-P07 model but is less continuous in SL2013 and
S40RTS. Overall there is a similar relative amplitude strength and vertical extent between the
UU-P07 and SL2013 models. Lateral extent differs between models. This slab dips to the SW.

Bijwaard et al. (1998), Fukao et al. (2001), Hall and Spakman (2002; 2004), Gorbatov and Kennett
(2003), Schellart et al. (2009), Schellart and Spakman (2012), and Fukao and Obayashi (2013).
In the south, the Kermadec section of the slab penetrates almost straight through the 660 km
discontinuity, but northward, a horizontal section of the slab, overlying the 660 km discontinuity
becomes prominent. This is explained by increased eastward roll-back of the slab towards the north
around an Euler pole relative to the mantle close to the southern tip of the slab in the Hikurangi
segment (Schellart and Spakman, 2012).
The onset of subduction along the Tonga-Kermadec-Hikurangi subduction zone is debated. Hall
(2002) suggested a 45 Ma onset based on the onset of Pacific-Australia convergence, whereas
Schellart et al. (2006) suggested that westward subduction along the Tonga subduction zone had
been ongoing since at least 85 Ma. Recently Wu et al. (2016) suggested 50 Ma as a start and on the
basis of the LLNL-G3Dv3 tomographic model of Simmons et al. (2012) that the Tonga-Kermadec
Hikurangi slab at ~1000km depth may be connected with the central Mariana slab. Prior to 50
Ma Wu et al. postulate that a western Pacific plate boundary already existed but was this may have
been characterised by highly oblique subduction or transforms. After 50 Ma, Pacific plate motions
changed and fast subduction began below the Philippine Sea Plate (Wu et al. 2016). Geochemical
analysis of dredge samples from the Tonga forearc found the first evidence of arc magmatism
around 50 Ma (Meffre et al., 2012), although these may be interpreted to reflect the initiation of
the New Caledonia subduction zone, correlated to the South Loyalty Basin slab (Schellart et al.,
2009) (section 5.83). In a recent plate kinematic analysis, van de Lagemaat et al. (2017) showed
that when subduction of the South Loyalty Basin slab along the New Caledonia trench is taken
into account, there is no or little convergence across the Tonga Trench until as young as 30 Ma,
and argued that correlating the onset of Tonga-Kermadec subduction to subduction initiation at

157
Figure 5.92. Trans-Americas anomaly. Legend same as figure 5.5. A broad positive anomaly ise
identified in the same location in both UU-P07 and S40RTS tomographic models with similar
lateral, vertical extent and relative amplitude. The slab is piling up above the CMB.

the Philippine Sea Plate at 52-51 Ma, or to the onset of Pacific-Australia convergence at 45 Ma
is kinematically unlikely. Given the ongoing debate on the age of subduction initiation, we adopt
a wide age range of 85-30 Ma for the onset of the Tonga-Kermadec-Hikurangi subduction zone.

5.88. Trans-Americas - TA
The Trans-Americas anomaly (Figure 5.92) is located below the Cocos plate and Central
America from the core-mantle boundary up to the deep mantle. It has been detected in previous
seismological (Niu and Wen, 2001; Thomas, 2004; Hutko et al., 2006; Kito et al., 2007; 2008) and
tomographic studies (van der Hilst et al., 2007, Ko et al. 2017). We infer by correlation to the base
of the Cocos slab (section 5.29) to the northeast and the base of the Idaho slab (section 5.42) to the
north that the Trans-Americas anomaly represents lithosphere that subducted during the Middle
Mesozoic or before. Van der Meer et al. (2010) previously interpreted the Trans-Americas slab to
result from subduction of Farallon/Panthalassa lithosphere in the Permian to Triassic associated
with the Sonoma orogeny at the western margin of Laurentia (Ziegler, 1989; Ward, 1995; Cawood
and Buchan, 2007). Alternatively, and perhaps more likely, the slab may have an early Mesozoic
intra-oceanic origin with its associated arc accreting to the western North American margin
during the Middle Mesozoic. The Stikinia-Quesnellia arc may fit such a scenario. This arc initiated
in the Middle-Late Triassic and accreted to Laurentia in the Early-Middle Jurassic (Nokleberg
et al., 2000; Johnston and Borel, 2007). However, the location of the Stikinia-Quesnellia terrane,
currently extending over 1200 km (Nokleberg et al. 2012), is poorly constrained and may in fact
represent more intra-oceanic arcs of varying ages and paleo-locations. As pointed out by Shephard
et al. (2013), other slabs (e.g., Wichita, Hudson, sections 5.91, and 5.41, respectively), would have
been correlated as well with this arc. Alternatively, Boschman and van Hinsbergen (2016) suggested
that the Trans-Americas slab may have detached as a result of triple junction migration at the

158
Figure 5.93. Ushkhy anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same locations in both UU-P07 and S40RTS tomographic models. Lateral, vertical extent and
relative amplitude are similar.

Izanagi-Farallon plate boundary, which culminated through instable triple junction formation in
the 190 Ma birth of the Pacific plate. Further research on this topic may require us to revise the age
of the slab, which we now tentatively place between 233 and 166 Ma covering the age of Stikinia-
Quesnellia subduction (see Wichita slab, section 5.91) as well as the birth of the Pacific plate.

5.89. Ushky - Ush


The Ushky anomaly (Figure 5.93) is located below northeast Asia in the upper mantle and
uppermost part of the lower mantle. It is E-W trending and at ~920 km depth it connects to the
N-S trending Sakhalin slab (section 5.75) in the west, and the NE-SW trending Kamchatka-
Kuriles slab (section 5.48) in the east. The anomaly’s location is consistent with the inferred paleo-
position of the continental margin Okhotsk-Chukotka arc. According to Nokleberg et al. (2000),
the arc started activity in the Cenomanian-Santonian and ended in the Miocene. In recent model
of Shephard et al. (2013), the arc was presumed to be active in the middle-Late Cretaceous (108–
67.1 Ma), based on the studies of Layer et al. (2001), Stone et al. (2009), and Vishnevskaya and
Filatova (2012), which we adopt as the age range of the slab. This was followed by the accretion
of the Okhotsk block to Siberia, followed by an eastward jump of subduction to the Kamchatka-
Kuriles trench (section 5.49).

159
Figure 5.94. Venezuela anomaly. Legend same as figure 5.5. A broad positive anomaly is identified
in the same locations in both UU-P07 and S40RTS tomographic models. Lateral, extent and
relative amplitude are similar. Vertical extent differs between models.

5.90. Venezuela - Ve
The Venezuela anomaly (Figure 5.94) is located below northern South America from the mid-
mantle up to the upper part of the lower mantle. It was first identified in van der Meer et al. (2010).
It was interpreted as the sGAC slab (Southern Great Arc of the Caribbean) by van Benthem et
al. (2013), who inferred a Cretaceous to Eocene subduction period for this slab. Subduction of
the Venezuela slab is likely associated with volcanic arc rocks in the southern Caribbean region
(Boschman et al., 2014), which on e.g. Tobago date back to ~130 Ma (Neill et al., 2012). Pindell et
al. (2012) suggested that subduction of the Venezuela slab started sometime before the geological
records mentioned above formed, and estimated an onset of subduction of 135 Ma. Arrest of the
southern Caribbean arcs in the latest Cretaceous (Neill et al., 2011), and subsequent overriding
of South America over the Venezuela slab in kinematic reconstructions (Pindell and Kennan,
2009; Boschman et al., 2014) suggests that break-off of the Venezuela slab occurred around 65±5
Ma, after which the northern part of the slab (Hispaniola slab, see section 5.40) retreated farther
northward (see also van Benthem et al., 2013).

5.91. Wichita - Wc
The Wichita anomaly (Figure 5.95) is located below central North America from the core-
mantle boundary up to deep mantle. It was first described in van der Meer et al. (2010), where
we interpreted the anomaly as the Wichita slab that is the result of intra-oceanic subduction
consuming Panthalassa lithosphere, creating the Mesozoic part of the intra-oceanic Paleozoic-
Mesozoic Stikinia-Quesnellia arc. This arc consists of an extensive suite of mainly Late Triassic
and Early Jurassic volcanic and granitic plutonic rocks, each extending for a distance of about
1,200 km (Nokleberg et al., 2000). Based on the long extent, another part of this arc is perhaps
associated with the Trans-Americas slab, section 5.88). Van der Meer et al. (2012) highlighted this

160
Figure 5.95. Wichita anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same locations in both UU-P07 and S40RTS tomographic models. Relative amplitude is similar.
Lateral and vertical extent differ between models.

slab as effectively forming the separation between the Slide Mountain and Thalassa Oceans. The
interpretation of this deep slab and the Trans Americas slab was to some extent corroborated by
the plate model of Shephard et al. (2013), who also identified material lying along the core-mantle
boundary at these locations using six tomographic models. They tentatively suggested that the
Wichita slab could represent Cache Creek Ocean and/or Farallon lithosphere that subducted along
the Talkeetna– Bonanza and Gravina arcs, or alternatively an older remnant of Slide Mountain
Ocean lithosphere subducted below earlier Stikinia-Quesnellia arcs. Johnston and Borel (2007)
interpreted a two-stage process, a first involving intra-oceanic subduction and arc formation and
accretion forming Stikinia-Quesnellia between 230 Ma to 150 Ma within the (eastern) Panthalassa.
In a second stage, from 150 Ma to 55 Ma, this arc migrated towards North America, collided,
and moved northward parallel to the continental margin. We correlate stage one with the Wichita
slab and stage two with subduction leading to the Hudson slab (section 5.41). Barresi et al. (2015)
analysed the lower Hazelton Group in central Stikinia, which comprises three volcanic-intrusive
complexes that nconstitute almost 16 km of volcanic stratigraphy. U–Pb zircon ages indicate that
volcanism initiated by ca. 204 Ma (latest Triassic). Detrital zircon populations from the basal
conglomerate contain abundant 205–233 Ma zircons, derived from regional unroofing of older
Triassic intrusions. They also linked the spatial co-occurrence of Hazelton Group volcanic rocks
with a belt of economic Cu–Au porphyry deposits (ca. 205–195 Ma) throughout northwestern
Stikinia. Strata in the upper Hazelton Group) with a U–Pb zircon age of 178.90±0.28 Ma represent
waning island-arc volcanism. Milidragovic et al. (2016) incorporated the above results and included
the absolute dating of the Hickman (222-220 Ma) and Three Sisters plutons (172±6 Ma) and as
the first and final evidence of volcanism in the Mesozoic Stikinia terrane. Based on these studies
above, we adopt a 233-220 Ma start of subduction and 178-166 Ma as the end defining the ages of
the base and top of the Wichita slab respectively.

161
Figure 5.96. Welford anomaly. Legend same as figure 5.5. The upper mantle part of this slab slab
is well imaged in the UU-P07 model but is less continuous in SL2013 and S40RTS. Overall there
is a similar relative amplitude strength, but lateral and vertical extent differ between tomographic
models. This slab dips to the NE.

5.92. Welford - We
The Welford anomaly (Figure 5.96) corresponds to the southwestern part of the A7 anomaly of
Hall and Spakman (2002; 2004) and underlies NE Australia, flat-lying at the top of the lower
and base of the upper mantle. Hall and Spakman (2002; 2004) note that their A7 anomaly is not
everywhere well-defined and may in fact represent more than one slab, whereby the northern part
of the slab below NW Australia (the Papua anomaly, section 5.70) may be disconnected from the
southwestern parts of the anomaly, which is the interpretation we follow here.
Kinematic reconstructions of the region NE of Australia consistently interpret a southwestward
Trobriand subduction zone to form as far-field response to collision of the Ontong Java plateau
with the Melanesian arc at 26-20 Ma (Hall, 2002; Quarles van Ufford and Cloos, 2005; Knesel et
al., 2008). This trench would have been active until ~10 Ma, after which northward subduction of
the New Britain slab (section 5.65) created the Bismarck arc since 10-5 Ma (Hall, 2002; Holm et
al., 2013), accommodating northward absolute plate motion of Australia.

162
Figure 5.97. Yukon anomaly. Legend same as figure 5.5. Positive anomalies are identified in the
same locations in both UU-P07 and S40RTS tomographic models. Relative amplitude and vertical
extent is similar. Lateral extent differs between models.

5.93. Yukon - Yuk


The Yukon anomaly (Figure 5.97) is located below northwestern North America and is within the
upper part of the lower mantle. Previously, van der Meer et al. (2010) interpreted it to be part of
the North Pacific slab (section 5.68). By closely examining the tectonic evolution and upper mantle
structure of the region we now prefer an interpretation that these are separate slabs. The Yukon slab
is E-W trending, and at ~900-1100 km it connects to the SW-NE trending North Pacific slab.
The location of the slab is in agreement with the location of the Upper Cretaceous-Lower Tertiary
continental margin Kluane arc in the reconstructions of Nokleberg et al. (2000) and Shapiro and
Solov’ev (2009). We prefer this connection over the previous interpretation of van der Meer et al.
(2010), which was based on an older reconstruction of the paleo-Pacific Ocean of Engebretson et
al. (1985).
Trop and Ridgway (2007) infer that the Kluane arc is of Campanian-Maastrichtian (83-66 Ma)
age. In the Late Maastrichtian- Early Paleocene (68–61 Ma) the continental margin was uplifted
and a regional unconformity developed (Trop and Ridgway, 2007). We consider this the best
estimate for the moment of slab detachment and use this to infer the age of the top of the slab.

163
Figure 5.98. Zagros anomaly. Legend same as figure 5.5. The upper mantle part of this slab slab
is well imaged in the UU-P07 model but is broader in SL2013 and S40RTS. Overall there is a
similar relative amplitude strength and vertical extent. Lateral extent differs between tomographic
models. This slab dips to the NE in the upper mantle and appears flat-lying at the top of the lower
mantle.

5.94. Zagros - Za
A northward dipping anomaly is located below the Zagros mountains of Iran in the upper mantle
and in the upper part of the lower mantle (Hafkenscheid et al., 2006; Zor, 2008; Chang et al., 2010;
Agard et al., 2011; Koulakov, 2011) (Figure 5.98). The anomaly is interpreted as the Zagros slab,
disconnected from the deeper Mesopotamia slab (section 5.60) by a gap that likely corresponds
to a phase of slab break-off. Along most of the Zagros mountain range, the Zagros slab is no
longer connected to the surface, and displays a clear gap suggesting a recent phase of slab break-off.
Agard et al. (2011) suggested, based on the age of rapid exhumation of HP-LT metamorphic rocks
found in the Zagros suture zone that there may have been a phase of slab break-off in the latest
Cretaceous to Paleogene, ~65±5 Ma. We inferred that this corresponds to the age of the top of the
Mesopotamia slab (section 5.60), and we also use this age as the maximum age of subduction of
the base of the Zagros slab. The most recent phase of slab break-off occurred in the last ~10 Myr
(Agard et al., 2011), which we adopt for the age of the top of the slab.

164
Sinking of slabs in the lower mantle and implications for the lower mantle viscosity
profile

As a first application and illustration of the potential use of our Atlas of the Underworld we explore
the sinking and deformation history of slab remnants in the mantle and the consequences thereof
for depth variation of mantle viscosity involved in sinking of slab in the lower mantle.

Variable sinking of slabs in the lower mantle

The more than three-fold increase of age-depth relations for tops and bases of slabs in comparison
to those of the 28 slabs of van der Meer et al. (2010) enables us to search for age-depth trends
in average sinking velocity. Figure 5.99A shows the results of our age-depth compilation and
Figure 5.99B displays the average sinking velocities that are derived from the age-depth relations.
All pertaining numbers and uncertainties are listed in Table A1. In Figures 5.100 and 5.101 these
results are split into depth intervals for the slab tops. A straight line connects the top and base

Figure 5.99. Age-depth relations and corresponding average sinking rates of tops and bases of slabs.
A) Age-depth relations for the top (green triangles) and base (blue diamonds) of tomographically
imaged slabs determined in the Atlas of the Underworld. Depth is determined from the
tomographic model UU-P07 and the age information corresponds for the top of a slab anomaly
to the end of subduction / slab detachment and for the anomaly base to the start of subduction
with depth and age uncertainties derived from the tomographic models and from the pertinent
geological literature, respectively. Anomalous slabs: Ar=Arabia; Ka=Kalimantan; TKh=Tonga–
Kermadec–Hikurangi; GI=Georgia Islands B) The average sinking rate of the top and base of
slabs determined from the age-depth data of A) as Vav(depth)=depth/age. Velocity uncertainties
are determined from the combined age and depth uncertainties in a). The depth uncertainty is the
same as in A). ”Slab-top”-symbols at the surface represent present-day subduction velocities as an
instantaneous estimate of average sinking velocity. The three orange vertical lines at 10, 15 and 20
mm/yr are plotted for identifying subtle depth-trends.

165
Figure 5.100. The age-depth relations from Figure 5.99A illustrated for each slab (lines connecting
slab top and base) and for different depth-domains for the top of slabs as indicated in each panel.
Anomalous slabs: Ar=Arabia; Br=Brasilia; GI=Georgia Islands; Ka=Kalimantan; C=Cocos;
Mp=Malpelo; TKh=Tonga–Kermadec–Hikurangi.

symbols for each slab in these figures. The first observation from Figure 5.99A is that sinking of
slab remnants occurred across the entire mantle during the past ~300 Myr and that there is no
long-term (>60 Myr) stagnation (e.g., Fukao et al. 2001, 2009; Fukao and Obayashi 2013) visible in
this downward flow, neither in the upper mantle transition zone, nor in the top of the lower mantle,
suggesting a single layer convective system for the mantle. Long-term (>60 Myr) stagnation would
lead to a flattening of the sinking pattern of Figure 5.99A in a particular depth range. We infer that
all subduction zones that were confined to Mesozoic times are associated with slab remnants that
now reside in the lower mantle. The deepest slab we identify required between ~200 Myr (Rockall
and Trans-Americas slabs) and ~300 Myr (Atlantis, Balkan, Georgia Islands slabs) to reach the
core-mantle boundary (CMB) (Figure 5.99A). The average whole-mantle sinking rate (mantle
thickness / transit time) we determine from our global compilation is 12.0±2.5 mm/yr (Figure
5.99A,B), similar to, or within range of previous estimates (van der Meer et al. 2010; Butterworth
et al. 2014; Domeier et al. 2016). However, we also identify depth trends in average sinking rate.
The fastest slabs reach the base of the upper mantle at ~660 km within ~10 Myr while some
slabs reside in the upper mantle until ~75 Myr since start of subduction (Figure 5.99A). This
spread of about ~65 Myr in subduction age can be attributed to a variety of subduction behaviour
comprising slow and fast subduction or slab retreat leading to slab flattening in the upper mantle
transition zone where the flat slab meets resistance to entering the lower mantle (e.g. Goes et al.
2008). This spread in subduction age increases to 80-110 Myr across the top 800-900 km of the
lower mantle, associated with what has been interpreted as a slab stagnation zone (Fukao et al.

166
Figure 5.101. Average sinking rate to depth relations for Figure 5.99B. Illustrated for each slab
(lines connecting slab top and base and for different depth-domains for the top of slabs as indicated
in each panel. Anomalous slabs: Ar=Arabia; Br=Brasilia; GI=Georgia Islands; Ka=Kalimantan;
C=Cocos; TKh=Tonga–Kermadec–Hikurangi.

2001, 2009; Fukao and Obayashi 2013). From the depth of 1500-1700 km until the CMB, this
spread in subduction age stays rather constant. Inspection of Figure 5.100 shows that slabs with
tops in the lower mantle sink with comparable rates suggested by the more or less parallel lines
connecting top and base of each slab (Figures 5.100C-F). We therefore suggest that the lower
mantle spread in subduction ages at each depth is largely inherited from that acquired in the upper
mantle. The much-reduced variation in average sinking velocities in the lower mantle, particularly
below ~1500 km retains this sinking pattern. The implication is that the position of a sinking slab
in Figure 5.99A translates for each slab into a first-order indication regarding the upper mantle
speed of paleo-subduction to be slow, fast, or intermediate, which puts qualitative constraints on for
instance plate tectonic reconstructions, or the geodynamic context of mountain building, or informs
independently on the possibility of e.g. large trench retreat leading to a flat slab at the base of the
upper mantle that strongly decreases the net average sinking rate.
Generally, we observe a deceleration of the average sinking rate with increasing depth. The bases
of the majority of actively subducting slabs (Figure 5.101A) show a much smaller average sinking
rate than the present-day convergence velocity at the trench. The latter is plotted at the surface as an
instantaneous estimate of average sinking rate. A few slabs with currently small subduction speed
at the trench show the opposite trend but the average sinking of their bases is not anomalously
high. By far most detached slabs (Figures 5.101B-F) show deceleration with depth implying that
the slab top is sinking faster than the slab base and thus that the slab must be thickening, either
by bulk deformation or by buckling. One particular exception is the detached Himalayas slab (Hi;

167
section 5.38, Figure 5.101B), which may accelerate because of the proposed strong slab suction due
to sinking of the lower-mantle Tethys slab remnants underneath (Faccenna et al. 2013). We note,
however, that the interpretation of the age range of subduction of the Himalaya slab is based on
its apparent volume (Replumaz et al., 2010b; van Hinsbergen et al., 2017), and few independent
geological constraints exist. We will argue in a following section that the general deceleration
trend results from an increase in viscosity accommodating slab sinking in the first half of the lower
mantle, where the spread in average sinking rates decreases from 10-40 mm/yr near the top to
10-15 mm/yr in the mid-mantle, below which there is not much change. Further, the spread in
sinking rates at a particular depth may reflect differences in forced (Himalaya example) or free
sinking, or lateral differences in mantle viscosity, e.g. due to long-term mantle plume activity in
regions near the edges of long-lived Large Low Shear Velocity Provinces (LLSVPs) (e.g. Torsvik et
al. 2008, 2010)
Figures 5.99-5.101 also allow identifying outlying slabs. Three exceptionally fast sinking slabs
have reached depths below 1500 km (Figure 5.99A,B) of which only the Tonga–Kermadec–
Hikurangi (TKH) slab is still actively subducting and may therefore still be influenced by upper
mantle subduction. However, the fast sinking of two detached slabs (Arabia, Kalimantan) requires
a different cause. These fast sinking lower mantle slabs occur in regions of mantle upwelling
associated with ‘Plume Generation Zones’ at the edges of the deep LLSVPs (e.g. Torsvik et al.
2008, 2010). The Arabia slab sinks close to the well-known Afar upwelling at the NE-edge of the
African LLSVP; the Kalimantan slab sinks in mantle that has reduced seismic velocities associated
with the Hainan upwelling that branches of the western edge of the Pacific LLSVP (Hall and
Spakman 2015). Although we cannot exclude an effect of active subduction we note that the
Tonga-Kermadec slab, with an average sinking velocity as high as 28 mm/yr, also sinks in an overall
slow seismic-velocity mantle associated with the Samoan plume (Chang et al. 2016) above the
southern edge of the Pacific LLSVP. Two particularly slowly sinking slabs identifiable in Figures
5.100E and 5.100C are the Georgia Islands slab (GI) and the Malpelo slab (Mp). In case of the
GI slab the delayed sinking may be due to early plume-slab interaction at a time when the slab may
have been flat in the upper mantle (section 5.33) in which case plume rise may have stalled sinking,
while the interpretation of the Mp slab is stated to be uncertain (section 5.53). Other slabs that
stand out are the Cocos (Co, section 5.29) and Brasilia (Ba, section 5.18) slabs (Figures. 5.100A,
5.101A) that we propose are still connected to active subduction and constitute the longest slabs on
Earth.

Average versus in situ slab sinking rates

While previously we only estimated a lower mantle averaged slab sinking rate of 12±3 mm/yr from
a much smaller slab catalogue (van der Meer et al. 2010), a major finding here is the depth variation
of depth-average sinking rates Vav (d) (Figures 5.99B; 101) in the lower mantle as a function of
depth d, as described qualitatively in the previous section. First, we like to emphasize that the
Vav (d) sinking rates are an integral measure of the actual slab sinking rates V(d,t) at depth d and
geological time t. For instance, the average sinking rate of the base of the Calabria slab (section
5.20) is 10 mm ± 3 mm/yr, but because it is temporarily flat-lying on the 660 km discontinuity
(e.g. Spakman and Wortel 2004) its actual sinking velocity is possibly V(660,t)≈0 mm/yr. Generally,
such temporal aspects of the sinking of slab are unknown and therefore we approximate V(d,t) by
the time-stationary in situ sinking velocity Vin_situ(d), which represents the time-averaged V(d,t) at

168
each mantle depth d. This allows us to define the relation Vav(d)=1/d ∫0d Vin_situ(z) dz, which after
differentiating leads to Vin_situ(d)=Vav(d) + d V'av(d). From this equation, the in situ time-stationary
sinking velocity Vin_situ(d) can be obtained for a given profile Vav(d) of average sinking velocity
for a particular (fictitious) slab. We note that we see no use yet for solving the integral equation
for Vin_situ(d) by some method of geophysical inversion. The lateral scatter in Vav(d) at fixed depth
may relate to, as indicated earlier, forcing by regional mantle flow or to slabs subject to different
mantle viscosities during their downward journey. The number and limited geographical spread of
Vav(d) estimates does not allow yet for solving for regional differences in the Vav(d) profile yielding
a formal inversion of the integral equation fairly meaningless because of the large uncertainties
expected.
Therefore, we adopt a forward approach in which we configured several slab-sinking profiles of
Vav (d) for fictitious slabs for which we calculated the corresponding time-stationary sinking rates
Vin_situ(d). The results are summarized in Figure 5.102. We conducted three batches of calculations
to illustrate the sensitivity of Vin_situ(d) for smooth changes in Vav (d). Figure 5.102A shows 8
fictitious Vav (d) profiles for slabs that assume a constant sinking rate of 12 mm/yr below ~1700 km
while at the depth of 660 km Vav(660) is designed to start from 10 mm/yr to 24 mm/yr in steps of
2 mm/yr. The smooth variation in the top part mimics the different sinking trends of slabs observed
in Figure 101, while the bottom part simulates the near constant Vav (d) in deeper half of the lower
mantle. The corresponding in situ sinking rates Vin_situ(d) (Figure 5.102D) shows that generally slab
sinking occurs at speeds less than 15 mm/yr and decelerating slabs in particular with speeds less
than 10 mm/yr. We note that at the CMB, Vav is lowered to 11.25 mm/yr to satisfy that Vin_situ
(CMB)=0 mm/yr. Slightly accelerating Vav(d)-profiles lead to a modest speed up of the slab while
the deceleration leads to slow-down to as small as 4-5 mm/yr. The deceleration occurs from 660 km
to ~1300 km. The lowest sinking rates occur in a depth range of about 300 km centred around 1300
km, which will take these slabs about 60 Myr to transit. Assuming incompressible deformation,
the vertical shortening rates are approximated by the factor Vin_situ(660)/Vin_situ(d) while the
horizontal surface area of a slab is extended reciprocally. The vertical shortening for each of the 8
slabs is plotted in Figure 5.102G and shows shortening with factors up to ~4. Shortening factors
between 2-3 have been proposed for lower mantle slabs in the Tethyan realm (Hafkenscheid et
al. 2006) which would change the plate-like character of upper mantle slab into laterally extended
amorphous “blobs” in the lower mantle, as observed with tomography. Below this slab deceleration
zone, slab anomalies can partially stretch because of increasing sinking rate, but will not assume
their plate-like shape. Figures 5.102B,E,H, show results from a similar batch of experiments with
the only difference that Vav=14 mm/yr below the depth of 1700 km and Vav (CMB)=13.12 mm/y.
The larger deep-mantle average sinking rate causes a general increase in rates Vin_situ(d) of a few
mm/yr while the thickening factor stays below 2.5 because deceleration is smaller.
In the last batch of experiments, Figures 5.102C,F,I, we probe into the sensitivity of Vin_situ(d)
to extreme as well as middle-of-the road Vav-profiles including small variations in these trends. As
a whole, this experiment encompasses the full range of our dataset, with the exception of the three
outliers (Ar, Ka, TKH) discussed above. The small variations we made in the Vav-profiles of each
subset of three profiles serve forging that a minimum of Vin_situ(d)>4 mm/yr occurs around in the
depth range from 900-1500 km. Choices made for Vav above ~1700 km also determine the small
variations in Vav below this depth such that Vav stays within limits of the age-depth relations of
Figure 5.99A. Obtaining a minimum in Vin_situ(d) requires that all Vav-profiles are decelerating in
the top of the lower mantle. A small drop in Vav below 2750 km of 0.5 - 0.7 mm/yr suffices to
satisfy the 0 mm/yr sinking rate at the CMB.

169
170
The resulting Vin_situ(d)-profiles of Figure 5.102F show in situ velocities decelerating from
12-22 mm/yr at the top of the lower mantle to ~4-8 mm/yr at particular depths between 1000-
1500 km. Below the particular depth were the minimum in Vin_situ(d) occurs all profiles speed up
to rates between 10-15 mm/yr in the deeper mantle in order to satisfy the average velocity Vav(d)
below ~1700 km, i.e. to keep within the inferred spread of age-depth relations (Figure 5.99A). The
corresponding shortening factors (Figure 5.102J) range generally between 2 and 3.5 with excursions
to factors larger than 5 for two extremely fast deceleration profiles. A factor of 5 would shorten
a 660 km long upper mantle slab to a ~130 km thick “pancake” in the depth range 1000-1250
km. Although this cannot be excluded, such strong shortening has so far not been suggested from
tomography or numerical modeling of subduction. Note that even slabs that sink with a minimum
Vav-profile (darker green profiles in Figure 5.102I) in which Vav reduces from 12-13 mm/yr at
the top of the lower mantle to only 9 mm/yr around 1250 km, lead to a slab deceleration with
shortening factors larger than 3. However, slab that subducts slowly in the upper mantle does not
need to thicken strongly in the lower mantle. An example is the Aegean slab that, as observed in
tomography, seems to transit the deceleration zone without considerable thickening. Differences
in upper mantle subduction speed may thus explain why some slab seems to stagnate while
others do not, as noted previously by Fukao and Obayashi (2013). These results show generally,
and expectedly, that a gentle decrease in Vav(d) between 660 km and 1500 km corresponds to a
much more pronounced decrease in the in situ sinking rate Vin_situ(d) of a slab leading to vertical
shortening. Similarly, a smooth 30% increase in Vav(d) from 9 mm/yr to 12 mm/yr from 1600 km
to 2750 km corresponds to an almost 100% increase of Vin_situ(d) from 8 mm/yr to 15 mm/yr.
Based on the average sinking velocities obtained from the Atlas (Figure 5.99B), we suggest
that in situ sinking rates Vin_situ(d) generally decrease in the top few hundred kilometres of the
lower mantle from 10-25 mm/yr to 4-7 mm/yr somewhere in the depth range 1000-1500 km.
Corresponding vertical shortening factors are between 2 and 3.5 but could be larger for more
rapid deceleration. A minimum sinking rate of ~5-10 mm/yr across a 200-300 km depth interval
where slab sinking is impeded most in the top 900 km of the lower mantle, would predict a slab
transit time between 20-60 Myr. Such gradual transition in sinking is not easily inferred from the
age-depth relations of Figure 5.99A, but would explain the increase in horizontal spread in these

Figure 5.102. Average sinking velocity converted to in situ sinking rate, vertical slab shortening
factors. a) Configured lower mantle sinking profiles Vav (d) for 8 fictitious slabs (dashed lines).
Below 2000 km Vav (d)=12 mm/yr is assumed which is reduced below 2750 km to reach 0 mm/
yr in situ sinking at the CMB. At 660 km the starting value of Vav= 10 mm/yr (left curve)
incremented in steps of 2 mm/yr to 24 mm/year (right curve). Grey symbols correspond to the
age-depth data of Figure 5.99A. B) As A) but for a constant average sinking velocity of 14 mm/yr
below 2000 km; C) Configured lower mantle sinking profiles Vav (d) for 9 fictitious slabs for slowly
sinking slabs (green curves) fast sinking slabs (red/orange curve), or following the average depth
trend (magenta curves). These curves where constrained by a minimum sinking speed of ~ 4 mm/yr
above 1500 km and were furthermore configured to illustrate the sensitivity of in situ sinking rates
Vin_situ(d) for systematic as well as small changes in the average sinking rates Vav(d). Below 2750 km
the Vav (d) profiles where artificially reduced by 0.5-0.7 mm/yr to match 0 mm/yr vertical sinking at
the CMB in line with the trend in the age-depth relations of Figure 5.99A (grey symbols); D-F) in
situ time-stationary sinking velocity Vin_situ(d) corresponding (color) to the average sinking velocity
curves of A-C). g-i) vertical shortening factors Vin_situ(660)/Vin_situ(d) corresponding to the in situ
sinking curves d-f ).

171
Figure 5.103. Viscosity profiles (coloured) corresponding to the in situ sinking profiles of Figure 5.102d, 5.102e and 5.102f, respectively, for
an ellipsoidal shaped slab with a vertical axis length of 600 km, a width of 2000 km, a thickness of 80 km and a temperature contrast of 250
K that sinks from the top of the lower mantle. The grey and black curves in the background are published mantle viscosity profiles shown for
comparison: CA, CB the prefered A- and B-profiles of Cízková et al. (2012); CF (Cadec and Fleitout 2003); EX2 and EX4 (Bower et al. 2013);
Ki (King 2016b); Lam (Lambeck et al. 2014); Lau (Lau et al. 2016); MF (Mitrovica and Forte, 2004); SC (Steinberger and Calderwood, 2006);
model VM5a from Argus et al. (2014); models V1 and V2 from Shahnas et al. (2017).

172
relations from ~65 Myr at the base of the upper mantle to roughly 80-110 Myr. Flat slab in the
upper-mantle transition zone will sink in the lower mantle within ~65 Myr or earlier while the
continuous flow of slabs from the top to the bottom of the lower mantle excludes long-term (e.g. >
100 Myr) stagnation due to compositional buoyancy effects (Morra et al. 2010; Ballmer et al. 2015).
Below ~1500 km in situ sinking rates are below 15 mm/yr for most lower mantle slabs. Only few
slabs are exceptionally faster than the above range, which we have discussed in the previous section.

Toward constraining the radial profile of lower mantle viscosity associated with slab sinking

The vertical shortening and lateral thickening of slab explains the primary tomographic
observations that led to the proposition of the slab stagnation zone between 660-1000 km (Fukao
et al. 2001, 2009; Fukao and Obayashi 2013) or down to 1500 km (Morra et al. 2010). Recent
explanations of lateral slab thickening vary from a gradual increase in mantle viscosity in the
top of the lower mantle, or a sudden viscosity increase near or below 1000 km, to compositional
effects raising slab buoyancy (Morra et al. 2010; Marquardt and Miyagi, 2015; Rudolph et al. 2015;
Ballmer et al. 2015). Despite these and other investigations there is, however, still no consensus on
the particular shape of the lower mantle viscosity profile (Karato 2010; King 2016a,b).
The continuous through-flow of slab demonstrates that the “slab stagnation zone” is rather a
“slab deceleration zone” that requires a gradual increase in mantle viscosity associated with slab
sinking below 660 km until ~1500 km. A rheological origin for the underlying increase of mantle
viscosity across the depth range of 660-1500 km, matching with our deceleration pattern, was
recently identified as a gradual increase in the strength of ferropericlase (Marquardt and Miyagi
2015), which is the weakest mineral accommodating lower mantle deformation (Karato 2010;
Girard et al. 2016). The same authors note that below ~1500 km the increase in viscosity becomes
smaller while in the deeper mantle viscosity can even reduce resulting from the Fe-spin crossover.
The latter may even cause a mid-mantle minimum in viscosity (Shahnas et al. 2017). Our patterns
of Vav(d) and Vin_situ(d) below ~1500-1700 km implicitly support these predictions in terms of a
slight acceleration in in situ sinking speeds for those slabs that initially decelerate in the top of the
lower mantle (Figure 5.102). We note for depths below 1600 km a small increase of the minimum
average sinking rates from 9 to 12 mm/yr (Figure 5.99B), which may possibly be attributed to
the Fe spin crossover lowering mantle viscosity (Shahnas et al. 2017). Concurrently, any decrease
in thermal expansivity and slow but progressive thermal assimilation of the slab remnant with
increasing residence time would gradually lower the density contrast between slab remnant and
ambient mantle and thus reduce the amplitude of the negative slab buoyancy. This would seemingly
promote slower sinking, but to still match the inferred sinking trends (Figures 5.99; 5.102) such
buoyancy decrease will actually require a reduction in mantle viscosity in the deeper mantle.
Any compositional contribution to the deceleration (Morra et al. 2010; Ballmer et al. 2015) is
insufficient to cause long-term (> 100 Myr) stagnation, as this is not observed (Figure 5.99A).
To investigate the implication of the in situ sinking profiles (Figure 5.102B,F,J) for the depth
variation in lower mantle viscosity we assume, considering the very low Reynolds number, that
detached slab sinks with a depth-dependent terminal velocity equal to the in situ velocity Vin_situ(d).
To estimate viscosity from terminal viscosity we follow the work of Kerr and Lister (1991) on
sinking of ellipsoidal olivine grains in magma. This is done for an ellipsoidal slab with a temperature
difference of 250 K estimated from seismic anomaly amplitudes (Goes et al. 2004), and with a
vertical axis of 600 km, a long horizontal axis (trench-length) of 1000 km, and a short horizontal

173
axis (slab width) of 80 km. Details are given in the appendix. The calculation includes depth-
dependent thermal expansivity (Steinberger and Calderwood 2006), but ignores effects of thermal
assimilation, as these are difficult to incorporate. An approximation for the effect of slab shape
change resulting from deceleration/acceleration is made by shortening/extending the principal axes.
Figure 5.103 shows the results of converting the Vin_situ-sinking profiles of Figures 5.102D-F into
viscosity profiles. Because of several assumptions and uncertainties (Appendix in online article),
the viscosity profiles computed serve primarily to be indicative of the general depth trend of lower
mantle viscosity related to slab sinking. The general observation is that slab decelerating in the top
half of the lower mantle experiences increasing viscosity. The viscosity peaks in the top half of the
lower mantle depending on sinking profile. Below 1500-1600 km viscosity slowly decreases due to
reducing density contrasts resulting from decreasing thermal expansivity and increasing background
density. Any reduction of the temperature contrast by diffusive heat transport would only enhance
this decreasing trend. Most important is that the total variation in viscosity with depth is rather
limited and does not depend on the slab shape, more extensive calculations show. We note that the
variability we model via our choices for average sinking profiles (Figures 5.102A-C) could possibly
reflect the effect of global lateral differences in temperature dependence of viscosity, e.g. slab sinking
in plume regions versus relatively sinking in relatively cold mantle.
Lower mantle viscosity associated with the sinking of slabs varies by less than a factor of 6
with depth, our results suggest. This is much less variation than seen in most of the published
viscosity profiles of which a subset is shown in Figure 5.103. Viscosity profiles that are explicitly
constrained by the average rate of slab sinking are the A- and B-profiles (CA and CB in Figure
5.103A) of Cízková et al. (2012). These resulted primarily from searching for viscosity profiles
that cause slabs to transit the mantle within 250-300 Myr, i.e. with an average sinking rate of 12
mm/yr, while allowing for depth variability in viscosity. Our results agree best with their A-family
of profiles. Also viscosity profiles EX2 and EX4 of Bower et al. (2013) match this transit time
constraint. In contrast to this work, all other published viscosity profiles have been estimated from
observations distinctly different from slab sinking rates. This concerns observations of the global
dynamic response to redistribution of ice-water surface loads, gravity observations being linked
to mantle structure, mantle dynamics, and Earth-rotational dynamics, constrained by mineral
physics (e.g. Hager et al. 1985; Ricard et al. 1993, Lambeck et al. 1998, 2014; Peltier 1998; Peltier
and Drummond 2010; Cadec and Fleitout 2003; Mitrovica and Forte, 2004; Steinberger and
Calderwood, 2006; Argus et al., 2014; Rudolph et al. 2015; Lau et al. 2016; Shahnas et al., 2017;
King 2016b). These studies probe the bulk viscosity field assuming a linear material response to
volumetrically distributed forcing of mantle flow. Most studies suggest a clear depth variation while
others suggest a more constant lower mantle viscosity. These differences may depend on the type
of data constraint used or simply on the detail of depth parameterization adopted. In contrast, our
inferences of slab sinking speeds pertain to the mantle rheology surrounding sinking slabs. There
can be a difference if lower mantle rheology can be stress-dependent, as is recently suggested
(Girard et al. 2016), and if slab sinking can generate the differential stresses required to activate
such rheology. This, however, needs to be further investigated.
Recent seismic tomography imaging of mantle plumes suggest that broad vertically rising
plumes change character by deflecting or thinning above ~1000 km (French and Romanowicz
2015). This concurs with upward decreasing mantle viscosity, alike the viscosity trends in reverse
direction suggested here for slab sinking, allowing for more plume mobility and thinning of the
plume diameter, which is also confirmed by 3-D numerical models (e.g. Leng and Gurnis 2012;

174
Rudolph et al. 2015). This suggests that viscosity trends shown here are of wider relevance than slab
sinking only and may also pertain to the bulk mantle.
Lastly, we note that proper assessment of lower mantle viscosity based on the sinking rate
findings should involve work similar to that of Cízková et al. (2012), while mantle convection
simulations can compare their prediction of average and in situ sinking rates to those found here.

Concluding remarks

In the preceding paragraphs 94 positive wave speed anomalies in the mantle have been interpreted
as subducted slabs. We only considered P-wavespeed anomalies that were also imaged, although at
a different resolution, in two S-wavespeed models: SL2013sv for the upper mantle and S40RTS
for the lower mantle. A preliminary comparison with recently constructed voting maps from 14
recent P- and S-tomography models (Shephard et al., submitted) reveals that the majority of
features identified in the voting maps match those identified here in the Atlas of the Underworld
(G. Shephard, personal communication). We identified the Cocos (Co) and Brasilia (Ba) slabs as
the longest slabs on Earth.
The strategy we followed for associating slab anomalies to geological evidence for subduction
is based on a regional-continental scale interpretation of geological records relative to each other
involving slab anomalies that are found with a similar spatial distribution at depth in the mantle.
This only assumes vertical sinking of detached slab remnant, which was recently corroborated
by Domeier et al. (2016). The interpretation of the start and end of subduction was based on
geological information with age ranges encompassing the estimates given in the pertinent
literature. Where geology implied distinctly different subduction events in a particular region,
this allowed for inspecting any internal lateral structure in large lower mantle slab anomalies. For
example, lateral structure in the huge Farallon slab anomaly has led to a subdivision into several
subduction segments: the Cocos, Hatteras and Hudson slabs, resulting from different subduction
zones. By treating these separately also enables a clearer discussion on their tectonic origins for
which proposed solutions differ between the Farallon Ocean plate subducting underneath the
American continental margin (i.e. Grand et al., 1997, Bunge & Grand 2000), or subduction of
several Panthalassa Ocean plates dominantly under the continental margin (van der Meer et al.,
2010, 2012), or dominantly at intra-oceanic subduction zones (Sigloch & Milhalynuk, 2013).
Future, more accurate, plate tectonic models relative to the mantle will be needed to converge to
one of these scenarios or combinations thereof. In this context, we emphasize again that in the
construction of the Atlas of the Underworld no assumptions were made about slab sinking rates.
Only for the interpretation of the enigmatic Maldives and Malpelo slab anomalies we cross-
checked with predictions from tectonic reconstructions cast in an absolute plate motion frame, but
for the vast majority of our orogen-slab associations we did not use any guidance from absolute
plate motion models, particularly not to allow the Atlas to become a source of information for
constructing a mantle frame based on subduction (van der Meer et al. 2010).
Under the Indo-Australian realm, many lower mantle slabs anomalies are imaged,
presumably of Mesozoic origin (van der Meer et al., 2012; Simmons et al. 2015). In plate tectonic
reconstructions these would be located in the transition of the Tethys to Panthalassa Oceans, which
remains poorly reconstructed. The mantle structure below the Pacific and East Asia may provide
an unprecedented means to interpret the Panthalassa plate kinematic history back to the Triassic,

175
linking the few presumed relics of intra-Panthalassa and Panthalassa-Tethys junction subduction
that accreted around the Pacific to subducted slab remnants.
We did not identify slabs that correlate to subduction zones older than ~300 Myr. This is
perhaps because much older subduction remnants may have been thermally assimilated in the hot
thermal boundary layer overlying the CMB. In addition, at the base of the mantle the occurrence of
rheologically weaker post-perovskite phase, mantle viscosities are considerably lower (Figure 5.103)
allowing for lateral advection and strong deformation of some slab material.
The Atlas includes several slabs that appear to have interacted with the two large low shear wave
velocity bodies (LLSVPs) on the core-mantle boundary. The LLSVPs are generally assumed to have
been long-lived, stable features along which edges plumes may have been generated throughout
the Phanerozoic or beyond (e.g., Burke et al., 2008; Torsvik et al., 2008; 2010; 2014). The Atlas
documents that the modern shapes of both LLSVPs appear to have been influenced by subducted
slabs. The Pacific LLSVP contains positive seismic velocity bodies in its centre that we correlated to
Telkhinia slabs (van der Meer et al., 2012). The Perm Anomaly below Siberia was recently assumed
to have migrated ~1500 km along the CMB as a function of subduction in the Paleopacific realm
(Flament et al., 2017), but our correlations suggest a more straightforward solution in which it
was separated from the large African LLSVP by the Balkan slab that we correlated to Permian
subduction in the Urals. In this scenario, the Perm anomaly may well have acted as generation zone
for the Siberian Traps (e.g., Torsvik et al., 2008; 2010).
The age-depth trend of slab sinking (Figures 5.99A, 100) and average sinking speeds (Figures
5.99B, 101) we derived from the Atlas suggest significant depth variability and possibilities to
retrieve the speed of subduction in the upper mantle for slab that are now in the lower mantle.
All slabs sink to the CMB. In the slab deceleration zone (660 km -1500 km), slab requires less
than ~60 Myr to cross a depth range of ~300 km where sinking speed is lowest (4-10 mm/yr).
Differences in the speed of subduction in the upper mantle may explain why some slabs thicken
strongly while others thicken only moderately. Our results do not agree with long-term stagnation
due to compositional buoyancy effects. The spread in sinking rates may reflect lateral heterogeneity
in the lower mantle viscosity as we identified some fast sinking slabs in regions of long-term plume
activity. Estimated in situ time-stationary sinking rates generally decrease in the upper 900 km of
the lower mantle depending on the assumed average-sinking speed profile and then show increase
below ~1500 km. Conversion to mantle viscosity profiles show generally a viscosity increase
somewhere between 660 km and 1500 km, while below 1500 km viscosity tends to slowly decrease.
Generally our estimated profiles show a depth-variation in viscosity far less than one order in
magnitude. The mismatch we find with many published viscosity profiles is puzzling if lower mantle
rheology around sinking slabs is not stress dependent. The depth trend in viscosity we qualitatively
infer agrees with that implied by the observed morphology change of rising plumes and recent
experimental finding regarding the rheological behavior of ferropericlase.
In conclusion, this Atlas of the Underworld aims to provide a global starting point, or
framework for linking plate tectonic history of the past ~300 Myr to present-day mantle structure.
This also shows great potential to provide novel constraints on lower mantle viscosity associated
with the sinking of slab remnants. Advances, corrections, and extensions are expected from more
detailed tomographic models and better dating of paleo-subduction evolution. We aimed at brief
descriptions per slab also to welcome post-peer review input on our Atlas via the website: http://
www.atlas-of-the-underworld.org/.

176
Epilogue

In this thesis, I aimed at searching for new ways of constraining paleo-geographic, -atmosphere
and -sea level reconstructions, through an extensive investigation of mantle structure in seismic
tomographic models. To this end, I explored evidence for paleo-subduction in these models and
how this may allow for a new avenue towards reconstructing the plate tectonic history of our planet.
The major findings of each chapter can be summarized as follows:

In Chapter 1 (van der Meer et al. 2010), 28 slabs were characterized in terms of ranges of depth
and timing of subduction. This effort led to 1) a determination of the sinking rates of slabs as a
novel implicit observation of the rate of lower mantle flow; 2) to constrain how far back in time
‘mantle memory’ of past subduction extends, being a maximum of 250-300 Myr; and 3) to test
whether slab remnants of paleo-subduction can be used as a constraint on absolute plate position
yielding a new approach to constrain absolute plate reconstructions by ‘slab-fitting’ reconstructions.

This methodology is applied in Chapter 2 (van der Meer et al. 2012) to study the mantle
underneath the Pacific Ocean. Geological evidence from east Asia suggested that accreted exotic
terranes with Lower Mesozoic volcanic arcs should have originated from central Panthalassa Ocean
locations, yet how these fitted in plate kinematic reconstructions remained poorly constrained.
Remnants of paleo-subduction were identified in the mantle, revealing the location of Mesozoic
intra-Panthalassa subduction zones and shedding light on the plate tectonic configuration of the
Panthalassa Ocean of which limited oceanic crust has been preserved until present-day.

In Chapter 3 (van der Meer et al. 2014) we interpret paleo-subduction zone configurations at
a series of depth slices of the mantle and calculate total slab, and hence subduction zone length
versus depth. With the age-depth relationship from Chapter 1, depth was converted into time of
subduction. This provided a first-order estimate of total global subduction zone length for the last
235 Myr. Total subduction zone length was shown to scale to ocean floor production rates at mid-
ocean ridges such that the average subduction rate at a trench is 6±1 cm/yr, similar to today’s rates.
Then, subduction zone length was used to compute volcanic degassing of CO2 from subduction
and ridge spreading. This served as input into a carbon-cycle climate model, improving previous
estimates inferred from eustatic sea-level variation from the 1970’s. The new degassing estimate
resulted in an improved fit between modelled atmospheric CO2 and proxy data. In addition it
provided a good first-order fit of 87Sr/86Sr ratios from marine carbonates, a record of paleo-seawater
composition.

This correlation between the 87Sr/86Sr record and plate tectonic activity was improved by correcting
for weathering of continents in Chapter 4 (van der Meer et al. 2017). Using the corrected 87Sr/86Sr
record we calculated global ocean floor production rates to the Neoproterozoic and used this to
compute a eustatic sea level curve independent from sequence stratigraphy. We compared this curve
with previous sea level and flooded shelf area curves derived from sequence stratigraphy and from
plate motion models. Aside from providing a novel method of estimating sea level variation as a
function of global plate tectonic activity, it extended sea level curves back in time by an additional
~300 Myr, to ~840 Myr.

177
In Chapter 5 the identification and geological interpretation of slabs of Chapter 1, 2 is expanded
to 94 slabs. I dubbed this compilation the ‘Atlas of the Underworld’, which provides a complete
overview of the present interpretation of slab remnants by the scientific community and ourselves.
The slabs' subduction age-depth data provided new average and in situ sinking rates. The fastest
slabs in the lower mantle are still actively subducting, long slabs. The slowest slabs, are detached,
flat-lying and may have been slowed in their upper mantle sinking history. Slab deceleration in the
top of the lower mantle is modelled as well as the required associated slab thickening or buckling.
Instead of stagnated slabs, we observe that all slabs sink further towards the core-mantle-boundary.
Slab acceleration may occur below 1700 km depth. The data enabled quantification of lower mantle
viscosity and agree with recent mineral experiments.

In this thesis, I tried to document the global record of paleo-subduction, in combination with
the geological record, to enable break-through insights for geodynamic, paleo-geographic, paleo-
climatic and paleo-sea level studies of Earth History. However, the results of Chapters 1 to 3,
should be considered a first pass of what constraints paleo-subduction from tomographic models
may bring. Considerable differences exist between tomographic models, both in terms of data used
and methodology. Limitations of the used P-wave tomographic model (Amaru, 2007) in terms of
coverage and resolution particularly below parts of the Pacific and Indian Oceans, in combination
with uncertainties in timing of subduction from the geological record, result in - at best - first
order estimates (10’s-100 Myr scale) of mantle flow, absolute constraints of paleolongitude
and global subduction lengths through time. Although Chapter 5 included other, independent
tomographic models, interpretations may be improved using more tomographic models and a
thorough discussion of the interpretation of geological records on a slab for slab case (e.g. Shephard
et al. 2012; Vissers et al. 2016; Schellart et al. 2009; Schellart and Spakman 2015), for all slabs.
To facilitate this, the Atlas-of-the-underworld.org website is designed with discussion forums for
every slab. Regarding the effects of plate tectonics on the atmosphere (Chapter 3) and hydrosphere
(Chapter 4), improvement may be made by incorporating more variables in the analysis, such as the
effects of plume volcanism (Müller et al. 2008, Conrad et al. 2013, Mills et al. 2014), sedimentation
and weathering (Goddéris et al. 2014) and spatial and temporal variation of input sources (87Sr/86Sr,
carbon content of subducting crust). With improved data and interpretations of paleo-subduction,
shorter timescales might be achievable (<10’s Myr). The Atlas of the Underworld (Chapter 5), is
therefore not an end, but a new starting point for increasing detail in studies of mantle dynamics,
paleogeographic, paleo-atmosphere and paleo-sea level reconstruction.

Douwe van der Meer

Ruislip, United Kingdom, June 18th 2017

178
References
Abdelwahed, M.F., Zhao, D., 2007. Deep structure of the Japan subduction zone. Physics of the Earth and Planetary Interiors
162, 32–52.
Abrajevitch, A.V., Ali, J.R., Aitchison, J.C., Badengzhu, Davis, A.M., Liu, J., Ziabrev, S.V., 2005. Neotethys and the India–
Asia collision: Insights from a palaeomagnetic study of the Dazhuqu ophiolite, southern Tibet. Earth and Planetary
Science Letters 233, 87–102.
Acharya, S.K., Mitra, N.D., 1986. Regional geology and tectonic setting of northeast India and adjoining region. Geology of
Nagaland Ophiolite, Memoir of the Geological Survey of India 6–12.
Advokaat, E.L., van Hinsbergen, D., Kaymakci, N., Vissers, R.L.M., Hendriks, B.W.H., 2014a. Late Cretaceous extension
and Palaeogene rotation-related contraction in Central Anatolia recorded in the Ayhan-Büyükkışla basin. International
Geology Review 56, 1813–1836.
Advokaat, E.L., van Hinsbergen, D., Maffione, M., Langereis, C.G., Vissers, R.L.M., Cherchi, A., Schroeder, R., Madani, H.,
Columbu, S., 2014b. Eocene rotation of Sardinia, and the paleogeography of the western Mediterranean region. Earth
and Planetary Science Letters 401, 183–195.
Agard, P., Jolivet, L., Vrielynck, B., Burov, E., Monié, P., 2007. Plate acceleration: The obduction trigger? Earth and Planetary
Science Letters 258, 428–441.
Agard, P., Omrani, J., Jolivet, L., Whitechurch, H., Vrielynck, B., Spakman, W., Monié, P., Meyer, B., Wortel, M.J.R., 2011.
Zagros orogeny: a subduction-dominated process. Geological Magazine 148, 692–725.
AkininV., Miller, E.L., 2011. Evolution of calc-alkaline magmas of the Okhotsk-Chukotka volcanic belt. Petrology 19,
237–277.
Algeo, T.J., Seslavinsky, K.B., 1995. Reconstructing Eustasy and Epeirogenic Trends from Palezoic Continental Flooding
Records, in: Haq, B.U. (Ed.), Sequence Stratigraphy and Depositional Response to Eustatic, Tectonic and Climate
Forcing. Kluwer Academic Publishers, The Netherlands, pp. 209–246.
Allègre, C.J., Louvat, P., Gaillardet, J., Meynadier, L., Rad, S., Capmas, F., 2010. The fundamental role of island arc weathering
in the oceanic Sr isotope budget. Earth Planet. Sci. Lett. 292, 51–56. doi:10.1016/j.epsl.2010.01.019
Al-Riyami, K., Robertson, A.H.F., Dixon, J.E., Xenophontos, C., 2002. Origin and emplacement of the Late Cretaceous
Baer–Bassit ophiolite and its metamorphic sole in NW Syria. Lithos 65, 225–260.
Amante, C., Eakins, B.W., 2009. ETOPO1 1 Arc-Minute Global Relief Model: Procedures, Data Sources and Analysis.
NOAA Tech. Memo. NESDIS NGDC-24. doi:10.7289/V5C8276M
Amaru, M.L., 2007. Global travel time tomography with 3-D reference models. Geologica Ultraiectina 274, 174 p.
Anczkiewicz, R., Platt, J.P., Thirlwall, M.F., Wakabayashi, J., 2004. Franciscan subduction off to a slow start: evidence from
high-precision Lu–Hf garnet ages on high grade-blocks: Earth and Planetary Science Letters 225. 147–161.
Aragón, E., D'Eramo, F., Castro, A., Pinotti, L., Brunelli, D., Rabbia, O., Rivalenti, G., Varela, R., Spakman, W., Demartis,
M., Cavarozzi, C.E., Aguilera, Y.E., Mazzucchelli, M., Ribot, A., 2011. Tectono-magmatic response to major
convergence changes in the North Patagonian suprasubduction system; the Paleogene subduction–transcurrent plate
margin transition. Tectonophysics 509, 218–237.
Arculus, R.J., Ishizuka, O., Bogus, K.A., Gurnis, M., Hickey-Vargas, R., Aljahdali, M.H., Bandini-Maeder, A.N., Barth, A.P.,
Brandl, A., Drab, L., do Monte Guerra, R., Hamada, M., Jiang, F., Kanayama, K., et al., 2015. A record of spontaneous
subduction initiation in the Izu–Bonin–Mariana arc. Nature Geoscience 8, 728–733.
Argnani, A., 2012. Plate motion and the evolution of Alpine Corsica and Northern Apennines. Tectonophysics 579, 207–219.
Argus, D.F., Peltier, W.R., Drummond, R. and Moore, A.W., 2014. The Antarctica component of postglacial rebound model
ICE-6G_C (VM5a) based on GPS positioning, exposure age dating of ice thicknesses, and relative sea level histories.
Geophysical Journal International, 198(1), 537-563.

179
Avdeev, B., Niemi, N.A., 2011. Rapid Pliocene exhumation of the central Greater Caucasus constrained by low-temperature
thermochronometry. Tectonics 30, doi: 10.1029/2010TC002808.
Ayala, R.C., Bayona, G., Cardona, A., Ojeda, C., Montenegro, O.C., Montes, C., Valencia, V., Jaramillo, C., 2012. The
paleogene synorogenic succession in the northwestern Maracaibo block: Tracking intraplate uplifts and changes in
sediment delivery systems. Journal of South American Earth Sciences 39, 93–111.
Babazadeh, S.A., De Wever, P., 2004. Early Cretaceous radiolarian assemblages from radiolarites in the Sistan Suture (eastern
Iran). Geodiversitas 26, 185–206.
Bache, F., Mortimer, N., Sutherland, R., Collot, J., Rouillard, P., Stagpoole, V., Nicol, A., 2013. Seismic stratigraphic record of
transition from Mesozoic subduction to continental breakup in the Zealandia sector of eastern Gondwana. Gondwana
Research 26, 1060-1078.
Bailey, W.R., Holdsworth, R.E., Swarbrick, R.E., 2000. Kinematic history of a reactivated oceanic suture: the Mammonia
Complex Suture Zone, SW Cyprus: Journal of the Geological Society of London 157. 1107–1126.
Ballmer, M.D., Schmerr, N.C., Nakagawa, T. and Ritsema, J., 2015. Compositional mantle layering revealed by slab stagnation
at~ 1000-km depth. Science advances, 1(11), p.e1500815.
Barboza-Gudiño, J.R., Orozco-Esquivel, M.T., Gómez-Anguiano, M., Zavala-Monsiváis, A., 2008. The Early Mesozoic
volcanic arc of western North America in northeastern Mexico: Journal of South American Earth Sciences 25. 49–63.
Barnett-Moore, N., Hosseinpour, M., Maus, S., 2016. Assessing discrepancies between previous plate kinematic models of
Mesozoic Iberia and their constraints. Tectonics 35, 1843-1862.
Barresi, T., Nelson, J.L., Dostal, J. and Friedman, R., 2015. Evolution of the Hazelton arc near Terrace, British Columbia:
stratigraphic, geochronological, and geochemical constraints on a Late Triassic–Early Jurassic arc and Cu–Au porphyry
belt. Canadian Journal of Earth Sciences, 52(7), 466-494.
Baxter, A.T., Aitchison, J.C., Ali, J.R., Zyabrev, S., 2010. Early cretaceous radiolarians from the spongtang massif, Ladakh,
NW India: Implications for Neo-Tethyan evolution. Journal of the Geological Society of London 167, 511–517.
Becker, T.W., Conrad, C.P., Buffett, B., Müller, R.D., 2009. Past and present seafloor age distributions and the temporal
evolution of plate tectonic heat transport. Earth Planet. Sci. Lett. 278, 233–242. doi:10.1016/j.epsl.2008.12.007
Berner, R.A., 1994. GEOCARB II; a revised model of atmospheric CO 2 over Phanerozoic time. Am. J. Sci. 294, 56–91.
doi:10.2475/ajs.294.1.56
Berner, R.A., 2006. Inclusion of the Weathering of Volcanic Rocks in the GEOCARBSULF Model. American Journal of
Science 306, 295–302.
Bertrand, G., Rangin, C., 2003. Tectonics of the western margin of the Shan plateau (central Myanmar): implication for the
India–Indochina oblique convergence since the Oligocene. Journal of Asian Earth Sciences 21, 1139–1157.
Bezada, M.J., Humphreys, E.D., Toomey, D.R., Harnafi, M., Davila, J.M., Gallart, J., 2013. Evidence for slab rollback in
westernmost Mediterranean from improved upper mantle imaging. Earth and Planetary Science Letters 368, 51–60.
Bezada, M.J., Levander, A., Schmandt, B., 2010. Subduction in the southern Caribbean: Images from finite-frequency P wave
tomography. Journal of Geophysical Research 115, B12333–19, doi: 10.1029/2010JB007682.
Biggin, A.J., Steinberger, B., Aubert, J., Suttie, N., Holme, R., Torsvik, T.H., van der Meer, D.G., van Hinsbergen, D.J.J.,
2012. Possible links between long-term geomagnetic variations and whole-mantle convection processes. Nat. Geosci. 5,
526–533. doi:10.1038/ngeo1521
Bijwaard, H., Spakman, W., Engdahl, E.R., 1998. Closing the gap between regional and global travel time tomography:
Journal of Geophysical Research 103. 30055–30078.
Biryol, C.B., Beck, S.L., Zandt, G., Özacar, A.A., 2011. Segmented African lithosphere beneath the Anatolian region inferred
from teleseismic P-wave tomography. Geophysical Journal International 184, 1037–1057.
Blakey, R.C., 2003. Carboniferous-Permian paleogeography of the assembly of Pangea, in: Wong, T.E. (Ed.), Fifteenth
International Congress on Carboniferous and Permian Stratigraphy: Utrecht, the Netherlands, Royal Netherlands
Academy of Arts and Sciences. pp. 443–465.

180
Blakey, R.C., 2008. Gondwana paleogeography from assembly to breakup—A 500 m.y. odyssey. Geol. Soc. Am. Spec. Pap.
441, 1–28. doi:10.1130/2008.2441(01)
Blanco, M.J., Spakman, W., 1993. The P-wave velocity structure of the mantle below the Iberian Peninsula: evidence for
subducted lithosphere below southern Spain. Tectonophysics 221, 13–34, doi: 10.1016/0040-1951(93)90025-F.
Boekhout, F., Spikings, R., Sempere, T., Chiaradia, M., Ulianov, A., Schaltegger, U., 2012. Mesozoic arc magmatism along the
southern Peruvian margin during Gondwana breakup and dispersal. Lithos 146-147, 48–64.
Bogdanov, N.A., Vishnevskaya, S., Kepezhinskas, K., Sukhov, A.N., Fedorchuk, A.V., 1987. Geology of the Southern Koryak
Plateau [in Russian]. Nauka, Moscow.
Bonnet, S., Crave, A., 2003. Landscape response to climate change: Insights from experimental modeling and implications for
tectonic versus climatic uplift of topography. Geology 31, 123–126. doi:10.1130/0091-7613(2003)031<0123:LRTCCI>
2.0.CO;2
Borneman, N.L., Hodges, K.V., Soest, M.C., Bohon, W., Wartho, J.-A., Cronk, S.S., and Ahmad, T., 2015. Age and structure
of the Shyok suture in the Ladakh region of northwestern India: Implications for slip on the Karakoram fault system.
Tectonics 34, 2011–2033.
Boschman, L.M., van Hinsbergen, D.J., 2016. On the enigmatic birth of the Pacific plate within the Panthalassa Ocean.
Science Advances 2, e1600022.
Boschman, L.M., van Hinsbergen, D.J.J., Torsvik, T.H., Spakman, W., Pindell, J., 2014. Kinematic reconstruction of the
Caribbean region since the Early Jurassic. Earth-Science Reviews, 138, 102-136.
Bourgeois, O., Cobbold, R., Rouby, D., Thomas, J.-C., Shein, V., 1997. Least squares restoration of Tertiary thrust sheets in
map view, Tajik depression, central Asia. Journal of Geophysical Research 102, 27553–27573.
Bower, D.J., Gurnis, M., Flament, N., 2015. Assimilating lithosphere and slab history in 4-D Earth models. Physics of the
Earth and Planetary Interiors 238, 8–22.
Bower, D.J., Gurnis, M., Seton, M., 2013. Lower mantle structure from paleogeographically constrained dynamic Earth
models. Geochemistry, Geophysics, Geosystems 14, 44–63.
Bradley, D.C., 2011. Secular trends in the geologic record and the supercontinent cycle. Earth-Sci. Rev. 108, 16–33.
doi:10.1016/j.earscirev.2011.05.003
Bröcker, M., Rad, G.F., Burgess, R., Theunissen, S., Paderin, I., Rodionov, N., Salimi, Z., 2013. New age constraints for the
geodynamic evolution of the Sistan Suture Zone, eastern Iran. Lithos 170-171, 17–34.
Brunet, C., Monié, P., Jolivet, L., Cadet, J.P., 2000a. Migration of compression and extension in the Tyrrhenian Sea, insights
from 40Ar/39Ar ages on micas along a transect from Corsica to Tuscany. Tectonophysics 321, 127–155.
Brunet, M.F., Spakman, W., Ershov, A.V., Nikishin, A.M., 2000b. Information given by the tomography on the geodynamics
of the Caucasus–Caspian area. European geophysical society annual meeting 2000, 25-29.
Buchs, D.M., Arculus, R.J., Baumgartner, O., Baumgartner-Mora, C., Ulianov, A., 2010. Late Cretaceous arc development
on the SW margin of the Caribbean Plate: Insights from the Golfito, Costa Rica, Azuero, Panama, complexes.
Geochemistry, Geophysics, Geosystems 11, doi: 10.1029/2009GC002901.
Bunge, H.P., Grand, S.P., 2000. Mesozoic plate-motion history below the northeast Pacific Ocean from seismic images of the
subducted Farallon slab: Nature 405, 337–340.
Burgess, P.M., Gurnis, M., 1995. Mechanisms for the formation of cratonic stratigraphic sequences. Earth Planet. Sci. Lett.
136, 647–663. doi:10.1016/0012-821X(95)00204-P
Burgess, P.M., Prince, G.D., 2015. Non-unique stratal geometries: implications for sequence stratigraphic interpretations.
Basin Res. 27, 351–365. doi:10.1111/bre.12082
Burke, K., 1988. Tectonic evolution of the Caribbean. Annual Review of Earth and Planetary Sciences 16, 201–230.
Burke, K., Steinberger, B., Torsvik, T.H., Smethurst, M.A., 2008. Plume Generation Zones at the margins of Large Low
Shear Velocity Provinces on the core–mantle boundary. Earth and Planetary Science Letters 265, 49–60.
Burtman, S., Molnar, P., 1993. Geological and geophysical evidence for deep subduction of continental crust beneath the
Pamir. Geological Society of America Special Paper 281, 76p.

181
Burton, R., Kendall, C.G.S.C., Lerche, I., 1987. Out of our depth: on the impossibility of fathoming eustasy from the
stratigraphic record. Earth-Sci. Rev. 24, 237–277. doi:10.1016/0012-8252(87)90062-6
Butterworth, N.P., Talsma, A.S., Müller, R.D., Seton, M., 2014. Geological, tomographic, kinematic and geodynamic
constraints on the dynamics of sinking slabs. Journal of Geodynamics 73, 1–13. doi:10.1016/j.jog.2013.10.006
Čadek, O. and Fleitout, L., 2003. Effect of lateral viscosity variations in the top 300 km on the geoid and dynamic topography.
Geophysical Journal International, 152(3), pp.566-580.
Camp, E., Griffis, R.J., 1982. Character, genesis and tectonic setting of igneous rocks in the Sistan suture zone, eastern Iran.
Lithos 15, 221–239.
Carminati, E., Wortel, M., Spakman, W., Sabadini, R., 1998a. The role of slab detachment processes in the opening of the
western–central Mediterranean basins: some geological and geophysical evidence. Earth and Planetary Science Letters
160, 651–665.
Carminati, E., Wortel, M.J.R., Meijer, T., Sabadini, R., 1998b. The two-stage opening of the western-central Mediterranean
basins: a forward modeling test to a new evolutionary model. Earth and Planetary Science Letters 160, 667–679.
Cawood, A., Buchan, C., 2007. Linking accretionary orogenesis with supercontinent assembly. Earth-Science Reviews 82,
217–256.
Çelik, Ö.F., Delaloyle, M.F., Feraud, G., 2006. Precise 40 Ar– 39 Ar ages from the metamorphic sole rocks of the Tauride
Belt Ophiolites, southern Turkey: implications for the rapid cooling history. Geological Magazine 143, 213–227.
Centeno-Garcia, E., Busby, C., Busby, M., Gehrels, G., 2011. Evolution of the Guerrero composite terrane along the Mexican
margin, from extensional fringing arc to contractional continental arc. Geological Society of America Bulletin 123,
1776–1797.
Chang, S.J., Ferreira, A.M. and Faccenda, M., 2016. Upper-and mid-mantle interaction between the Samoan plume and the
Tonga-Kermadec slabs. Nature communications, 7.
Chang, S.-J., van der Lee, S., Flanagan, M.P., Bedle, H., Marone, F., Matzel, E.M., Pasyanos, M.E., Rodgers, A.J.,
Romanowicz, B., Schmid, C., 2010. Joint inversion for three-dimensional S velocity mantle structure along the Tethyan
margin. Journal of Geophysical Research 115, B08309–doi:10.1029–2009JB007204.
Chauvel, C., Lewin, E., Carpentier, M., Arndt, N.T., Marini, J.-C., 2008. Role of recycled oceanic basalt and sediment in
generating the Hf-Nd mantle array. Nat. Geosci. 1, 64–70.
Chekhovich V.D., Sukhov, A.N., Sheremet, O.G., Kononov, M.V., 2012. Cenozoic geodynamics of the Bering Sea region.
Geotectonics 46, 212–231.
Chekhovich, V.D., Sheremet, O.G. and Kononov, M.V., 2014. Strike-slip fault system in the Earth’s crust of the Bering Sea:
A relic of boundary between the Eurasian and North American lithospheric plates. Geotectonics, 48(4), 255-272.
Chen, C., Zhao, D., Tian, Y., Wu, S., Hasegawa, A., Lei, J., Park, J.H. and Kang, I.B., 2017. Mantle transition zone, stagnant
slab and intraplate volcanism in Northeast Asia. Geophysical Journal International, p.ggw491.
Chen, Y., Zhu, D.-C., Zhao, Z.-D., Meng, F.-Y., Wang, Q., Santosh, M., Wang, L.-Q., Dong, G.-C., Mo, X.-X., 2014. Slab
breakoff triggered ca. 113Ma magmatism around Xainza area of the Lhasa Terrane, Tibet. Gondwana Research 26,
449–463.
Chertova, M.V., Spakman, W., van den Berg, A.P., Geenen, T., van Hinsbergen, D.J.J., 2014. Underpinning tectonic
reconstructions of the western Mediterranean region through dynamic slab evolution from 3D numerical modeling.
Journal of Geophysical Research 119, 5876–5902.
Christensen, U.R., 1996. The influence of trench migration on slab penetration into the lower mantle. Earth and Planetary
Science Letters 140, 27–39.
Christie-Blick, N., 1991. Onlap, offlap, and the origin of unconformity-bounded depositional sequences. Mar. Geol. 97,
35–56. doi:10.1016/0025-3227(91)90018-Y
Chu, R., Schmandt, B., Helmberger, D.V., 2012. Juan de Fuca subduction zone from a mixture of tomography and waveform
modeling. Journal of Geophysical Research 117, doi: 10.1029/2012JB009146.

182
Chung, S.L., Wang, S.L., Shinjo, R., Lee, C.S., Chen, C.H., 2000. Initiation of arc magmatism in an embryonic continental
rifting zone of the southernmost part of Okinawa Trough. Terra Nova 12, 225–230.
Cízková, H., van den Berg, A.P., Spakman, W., Matyska, C., 2012. The viscosity of Earth’s lower mantle inferred from sinking
speed of subducted lithosphere. Physics of The Earth and Planetary Interiors 200-201, 56–62.
Clift, P.D., Pavlis, T., DeBari, S.M., Draut, A.E., Rioux, M. and Kelemen, P.B., 2005. Subduction erosion of the Jurassic
Talkeetna-Bonanza arc and the Mesozoic accretionary tectonics of western North America. Geology, 33(11), 881-884.
Cloetingh, S., Haq, B.U., 2015. Inherited landscapes and sea level change. Science 347, 1258375. doi:10.1126/science.1258375
Cloetingh, S., McQueen, H., Lambeck, K., 1985. On a tectonic mechanism for regional sealevel variations. Earth Planet. Sci.
Lett. 75, 157–166. doi:10.1016/0012-821X(85)90098-6
Cloos, M., 2005. Collisional Delamination in New Guinea. Geological Society of America Special Papers 400, 51 p.
Cluzel, D., Jourdan, F., Meffre, S., Maurizot, P., Lesimple, S., 2012. The metamorphic sole of New Caledonia ophiolite:
40Ar/ 39Ar, U-Pb, and geochemical evidence for subduction inception at a spreading ridge. Tectonics 31, doi:
10.1029/2011TC003085.
Cluzel, D., Meffre, S., Maurizot, P., Crawford, A.J., 2006. Earliest Eocene (53 Ma) convergence in the Southwest Pacific:
evidence from pre-obduction dikes in the ophiolite of New Caledonia. Terra Nova 18, 395–402.
Cocks, L.R.M., Torsvik, T.H., 2007. Siberia, the wandering northern terrane, and its changing geography through the
Palaeozoic. Earth-Science Reviews 82, 27–74.
Cocks, L.R.M., Torsvik, T.H., 2011. The Palaeozoic geography of Laurentia and western Laurussia: A stable craton with
mobile margins. Earth-Science Reviews 106, 1–51.
Cogné, J.-P., Humler, E., 2004. Temporal variation of oceanic spreading and crustal production rates during the last 180 My.
Earth Planet. Sci. Lett. 227, 427–439. doi:10.1016/j.epsl.2004.09.002
Cogne, J.P., KravchinskyA., Halim, N., Hankard, F., 2005. Late Jurassic-Early Cretaceous closure of the Mongol Okhotsk
Ocean demonstrated by new Mesozoic palaeomagnetic results from the Trans-Baikal area (SE Siberia). Geophysical
Journal International 163, 813–832.
Coltice, N., Rolf, T., Tackley, P.J., Labrosse, S., 2012. Dynamic Causes of the Relation Between Area and Age of the Ocean
Floor. Science 336, 335–338. doi:10.1126/science.1219120
Coltice, N., Seton, M., Rolf, T., Müller, R.D., Tackley, P.J., 2013. Convergence of tectonic reconstructions and mantle
convection models for significant fluctuations in seafloor spreading. Earth Planet. Sci. Lett. 383, 92–100. doi:10.1016/j.
epsl.2013.09.032
Condie, K., Pisarevsky, S.A., Korenaga, J., Gardoll, S., 2015. Is the rate of supercontinent assembly changing with time?
Precambrian Res., Supercontinental Cycles and Geodynamics 259, 278–289. doi:10.1016/j.precamres.2014.07.015
Conrad, C.P., 2013. The solid Earth’s influence on sea level. Geol. Soc. Am. Bull. 125, 1027–1052. doi:10.1130/B30764.1
Corfield, R.I., Searle, M.P., Pederson, R.B., 2001. Tectonic Setting, Origin, and Obduction History of the Spontang
Ophiolite, Ladakh Himalaya, NW India. Journal of Geology 109, 715–736.
Coulon, C., Fourcade, S., Maury, R.C., Bellon, H., Louni-Hacini, A., Cotten, J., Coutelle, A., Hermitte, D., 2002. Post-
collisional transition from calc-alkaline to alkaline volcanism during the Neogene in Oranie (Algeria): magmatic
expression of a slab breakoff. Lithos 62, no. 3, 87–110.
Cowgill, E., 2010. Cenozoic right-slip faulting along the eastern margin of the Pamir salient, northwestern China. Geological
Society of America Bulletin 122, no. 1-2, 145–161, doi: 10.1130/B26520.1.
Cowgill, E., Forte, A.M., Niemi, N.A., Avdeev, B., Tye, A., Trexler, C., Javakhishvili, Z., Elashvili, M., Godoladze, T., 2016.
Relict basin closure and crustal shortening budgets during continental collision: An example from Caucasus sediment
provenance. Tectonics, 35, doi: 10.1002/2016TC004295.
Cowgill, E., Yin, A., Harrison, T.M., Wang, X.F., 2003. Reconstruction of the Altyn Tagh fault based on U-Pb
geochronology: Role of back thrusts, mantle sutures, and heterogeneous crustal strength in forming the Tibetan Plateau.
Journal of Geophysical Research 108, no. B7, 2346–doi:10.1029–2002JB002080, doi: 10.1029/2002JB002080.

183
Cox, G.M., Halverson, G.P., Stevenson, R.K., Vokaty, M., Poirier, A., Kunzmann, M., Li, Z.-X., Denyszyn, S.W., Strauss,
J.V., Macdonald, F.A., 2016. Continental flood basalt weathering as a trigger for Neoproterozoic Snowball Earth. Earth
Planet. Sci. Lett. 446, 89–99. doi:10.1016/j.epsl.2016.04.016
Creager, J.S., Scholl, D.W., 1973. Initial reports of the Deep Sea Drilling Project. Printing Office, Washington, DC.
Csontos, L., Voros, A., 2004. Mesozoic plate tectonic reconstruction of the Carpathian region. Palaeogeography,
Palaeoclimatology, Palaeoecology 210, 1–56.
Curray, J.R., 1964. Transgressions and regressions, in: Miller, R.L. (Ed.), Papers in Marine Geology, Shepard Commemorative
Volume. Macmillan, New York, pp. 175–203.
D’Souza, R.J., Canil, D. and Creaser, R.A., 2016. Assimilation, differentiation, and thickening during formation of arc crust in
space and time: The Jurassic Bonanza arc, Vancouver Island, Canada. Geological Society of America Bulletin, 128(3-4),
543-557.
Dalziel, I.W.D., Lawver, L.A., Norton, I.O., Gahagan, L.M., 2013. The Scotia Arc: Genesis, Evolution, Global Significance.
Annual Review of Earth and Planetary Sciences 41, p. 767–793.
Dasgupta, R., Hirschmann, M.M., 2006. Melting in the Earth's deep upper mantle caused by carbon dioxide. Nature 440,
659–662. doi:10.1038/nature04612
Dasgupta, R., Hirschmann, M.M., 2010. Earth and Planetary Science Letters. Earth and Planetary Science Letters 298,
1–13. doi:10.1016/j.epsl.2010.06.039
de Boorder, H., Spakman, W., White, S.H., Wortel, M., 1998. Late Cenozoic mineralization, orogenic collapse and slab
detachment in the European Alpine Belt. Earth and Planetary Science Letters 164, 569–575.
De Jonge, M.R., Wortel, M., Spakman, W., 1994. Regional scale tectonic evolution and the seismic velocity structure of the
lithosphere and upper mantle: the Mediterranean region. Journal of Geophysical Research 99, 12091–12108.
De Jonge, M.R., Wortel, M.J.R., Spakman, W., 1993. From tectonic reconstruction to upper mantle model: an application to
the Alpine-Mediterranean region. Tectonophysics 223, 53–65.
DeCelles, G., Ducea, M.N., Kapp, P., Zandt, G., 2009. Cyclicity in Cordilleran orogenic systems. Nature Geoscience 2,
251–257.
Delavari, M., Amini, S., Schmitt, A.K., McKeegan, K.D., Harrison, T.M., 2014. U–Pb geochronology and geochemistry of
Bibi-Maryam pluton, eastern Iran: Implication for the late stage of the tectonic evolution of the Sistan Ocean. Lithos
200-201, 197–211.
Demouy, S., Paquette, J.L., de Saint Blanquat, M., Benoit, M., Belousova, E.A., O'Reilly, S.Y., García, F., Tejada, L.C.,
Gallegos, R., Sempere, T., 2012. Spatial and temporal evolution of Liassic to Paleocene arc activity in southern Peru
unraveled by zircon U–Pb and Hf in-situ data on plutonic rocks. Lithos 155, 183–200.
Dessert, C., Dupré, B., Gaillardet, J., François, L.M., Allègre, C.J., 2003. Basalt weathering laws and the impact of basalt
weathering on the global carbon cycle. Chem. Geol., Controls on Chemical Weathering 202, 257–273. doi:10.1016/j.
chemgeo.2002.10.001
Dewey, J.F., Casey, J.F., 2011. The origin of obducted large-slab ophiolite complexes, in Brown, D. and Ryan, D. eds., Arc-
continent collision. Frontiers in Earth Sciences, 431–444.
Dickens, G.R., 2011. Down the Rabbit Hole: Toward appropriate discussion of methane release from gas hydrate systems
during the Paleocene-Eocene thermal maximum and other past hyperthermal events. Climate of the Past 7, 831–846.
Dickinson, W.R., 2006. Geotectonic evolution of the Great Basin. Geosphere 2, 353–17.
Dickinson, W.R., Lawton, T.F., 2001. Carboniferous to Cretaceous assembly and fragmentation of Mexico. Geological
Society of America Bulletin 113, 1142–1160.
Dilek, Y., Furnes, H., 2009. Structure and geochemistry of Tethyan ophiolites and their petrogenesis in subduction rollback
systems. Lithos 113, 1–20.
Dimo-Lahitte, A., Monié, P., Vergély, P., 2001. Metamorphic soles from the Albanian ophiolites: Petrology, 40Ar/39Ar
geochronology, and geodynamic evolution. Tectonics 20, 78–96.

184
Domeier, M., Doubrovine, V., Torsvik, T.H., Spakman, W., Bull A., 2016. Global correlation of lower mantle structure and
past subduction. Geophysical Research Letters 43, 4945-4953.
Donskaya, T.V., Gladkochub, D.P., Mazukabzov, A.M., Ivanov, A.V., 2013. Late Paleozoic – Mesozoic subduction-related
magmatism at the southern margin of the Siberian continent and the 150 million-year history of the Mongol-Okhotsk
Ocean. Journal of Asian Earth Sciences 62, 79–97.
Doubrovine, P. V., B. Steinberger, T. H. Torsvik, 2012. Absolute plate motions in a reference frame defined by
moving hot spots in the Pacific, Atlantic, and Indian oceans, Journal of Geophysical Research 117-B09101,
doi:10.1029/2011JB009072.
Dziewonski, A.M., 1984. Mapping the lower mantle: Determination of lateral heterogeneity in P velocity up to degree and
order 6. Journal of Geophysical Research 89-B7, 5929-5952.
Eagles, G., 2005. Tectonic evolution of the west Scotia Sea. Journal of Geophysical Research 110, B02401–19, doi:
10.1029/2004JB003154.
Eagles, G., 2010. The age and origin of the central Scotia Sea. Geophysical Journal International 183, 587–600.
Eagles, G., and Jokat, W., 2014. Tectonic reconstructions for paleobathymetry in Drake Passage. Tectonophysics 611, 28–50.
Eagles, G., Livermore, R., Morris, P., 2006. Small basins in the Scotia Sea. The Eocene Drake Passage gateway. Earth and
Planetary Science Letters 242, 343–353.
Elliot, D.H., Fanning, C.M., 2008. Detrital zircons from upper Permian and lower Triassic Victoria Group sandstones,
Shackleton Glacier region, Antarctica: Evidence for multiple sources along the Gondwana plate margin. Gondwana
Research 13, 259–274.
Engebretson, D.C., Cox, A., Gordon, R.G., 1985. Relative motions between oceanic and continental plates in the Pacific
Basin. Geological Society of America Special Paper 206. 59p..
Engebretson, D.C., Kelley, K.P., Cashman, H.J., Richards, M.A., 1992. 180 million years of subduction. GSA Today, v. 2,
92–95.
Faccenna, C., Becker, T.W., Auer, L., Billi, A., 2014. Mantle dynamics in the Mediterranean. Reviews of Geophysics 52, 283-
332.
Faccenna, C., Becker, T.W., Jolivet, L. and Keskin, M., 2013. Mantle convection in the Middle East: Reconciling Afar
upwelling, Arabia indentation and Aegean trench rollback. Earth and Planetary Science Letters, 375, 254-269.
Faccenna, C., Becker, T.W., Pio Lucente, F., Jolivet, L., Rossetti, F., 2001. History of subduction and back-arc extension in the
Central Mediterranean. Geophysical Journal International 145, 809–820.
Faccenna, C., Bellier, O., Martinod, J., Piromallo, C., Regard, V., 2006. Slab detachment beneath eastern Anatolia: A possible
cause for the formation of the North Anatolian fault. Earth and Planetary Science Letters 242, 85–97.
Faccenna, C., Jolivet, L., Piromallo, C., Morelli, A., 2003. Subduction and the depth of convection in the Mediterranean
mantle. Journal of Geophysical Research 108, 2099, doi:10.1029–2001JB001690.
Faccenna, C., Oncken, O., Holt, A.F. and Becker, T.W., 2017. Initiation of the Andean orogeny by lower mantle subduction.
Earth and Planetary Science Letters, 463, pp.189-201
Faccenna, C., Piromallo, C., Crespo-Blanc, A., Jolivet, L., Rossetti, F., 2004. Lateral slab deformation and the origin of the
western Mediterranean arcs. Tectonics 23, doi: 10.1029/2002TC001488.
Fan, J.-J., LI, C., Liu, Y.-M., Xu, J.-X., 2015. Age and nature of the late Early Cretaceous Zhaga Formation, northern Tibet:
constraints on when the Bangong–Nujiang Neo-Tethys Ocean closed. International Geology Review 57, 342–353.
Filatova, N.I., Vishnevskaya, V.S., 1997. Radiolarian stratigraphy and origin of the Mesozoic terranes of the continental
framework of the northwestern Pacific (Russia). Tectonophysics 269, 131–150. doi:doi: DOI: 10.1016/S0040-1951
(96)00156-4
Fildani, A., Hessler, A.M., 2005. Stratigraphic record across a retroarc basin inversion: Rocas Verdes–Magallanes Basin,
Patagonian Andes, Chile. Geological Society of America Bulletin 117, 1596–20.
Fischer, A.G., 1981. Climatic oscillations in the biosphere, in: Nitecki, M.H. (Ed.), Biotic Crises in Ecological and
Evolutionary Time. Academic Press, New York, N.Y., pp. 103–131.

185
Fischer, A.G., 1984. The two Phanerozoic supercycles, in: Berggren, W.A., Van Couvering, J. (Eds.), Catastrophies and Earth
History. Princeton Press, Princeton, N.J., pp. 129–150.
Flament, N., Coltice, N., Rey, P.F., 2013. The evolution of the 87Sr/86Sr of marine carbonates does not constrain continental
growth. Precambrian Res., Evolving Early Earth 229, 177–188. doi:10.1016/j.precamres.2011.10.009
Flament, N., Williams, S., Müller, R.D., Gurnis, M. and Bower, D.J., 2017. Origin and evolution of the deep thermochemical
structure beneath Eurasia. Nature communications, 8, 14164.
Forney, G.G., 1975. Permo-Triassic Sea-Level Change. J. Geol. 83, 773–779.
Forte, A.M., Cowgill, E., Bernardin, T., Kreylos, O., Hamann, B., 2010. Late Cenozoic deformation of the Kura fold-thrust
belt, southern Greater Caucasus. Geological Society of America Bulletin 122, 465–486.
Forte, A.M., Cowgill, E., Whipple, K.X., 2014. Transition from a singly vergent to doubly vergent wedge in a young orogen:
The Greater Caucasus. Tectonics 33, 2077-2101.
Fotoohi Rad, G.R., Droop, G.T.R., Amini, S., Moazzen, M., 2005. Eclogites and blueschists of the Sistan Suture Zone,
eastern Iran: A comparison of P–T histories from a subduction mélange. Lithos 84, 1–24.
Fotoohi Rad, G.R., Droop, G.T.R., Burgess, R., 2009. Early Cretaceous exhumation of high-pressure metamorphic rocks of
the Sistan Suture Zone, eastern Iran. Geological Journal 44, 104–116.
French, S.W. and Romanowicz, B., 2015. Broad plumes rooted at the base of the Earth's mantle beneath major hotspots.
Nature, 525(7567), 95-99.
French, S.W., Romanowicz, B.A., 2014. Whole-mantle radially anisotropic shear velocity structure from spectral-element
waveform tomography. Geophysical Journal International 199, 1303–1327.
Fritzell, E.H., Bull, A.L., Shephard, G.E., 2016. Closure of the Mongol–Okhotsk Ocean: Insights from seismic tomography
and numerical modelling. Earth and Planetary Science Letters 445, 1–12.
Fukao, Y., Obayashi, M. and Nakakuki, T., 2009. Stagnant slab: a review. Annual Review of Earth and Planetary Sciences, 37,
19-46.
Fukao, Y., Obayashi, M., 2013. Subducted slabs stagnant above, penetrating through, and trapped below the 660 km
discontinuity. Journal of Geophysical Research 118, 5920–5938.
Fukao, Y., Obayashi, M., Inoue, H., Nenbai, M., 1992. Subducting slabs stagnant in the mantle transition zone. Journal of
Geophysical Research 97, 4809–4822.
Fukao, Y., Widiyantoro, S., Obayashi, M., 2001. Stagnant Slabs in the Upper and Lower Mantle Transition Region. Reviews
of Geophysics 39, 291–323.
Gaffin, S., 1987. Ridge volume dependence on seafloor generation rate and inversion using long term sealevel change.
American Journal of Science 287, 596–611.
Gągała, Ł., Vergés, J., Saura, E., Malata, T., Ringenbach, J.-C., Werner, P., Krzywiec, P., 2012. Architecture and orogenic
evolution of the northeastern Outer Carpathians from cross-section balancing and forward modeling. Tectonophysics
532-535, 223–241.
Gaherty, J.B., Hager, B.H., 1994. Compositional vs. thermal buoyancy and the evolution of subducted lithosphere.
Geophysical Research Letters 21, 141–144.
Gaina, C., Müller, D., 2007. Cenozoic tectonic and depth/age evolution of the Indonesian gateway and associated back-arc
basins. Earth-Science Reviews 83, 177–203.
Gaina, C., van Hinsbergen, D.J.J., Spakman, W., 2015. Tectonic interactions between India and Arabia since the
Jurassic reconstructed from marine geophysics, ophiolite geology, and seismic tomography. Tectonics 34,
doi:10.1002/2014TC003780.
García-Casco, A., Iturralde-Vinent, M.A., Pindell, J., 2008. Latest Cretaceous Collision/Accretion between the Caribbean
Plate and Caribeana: Origin of Metamorphic Terranes in the Greater Antilles. International Geology Review 50,
781–809.
García-Casco, A., Torres-Roldan, R., Iturralde-Vinent, M.A., Trujillo, G.M., Cambra, K.E.N., Lázaro, C., Vega, A.R., 2006.
High pressure metamorphism of ophiolites in Cuba. Geologica Acta 4, 63-88.

186
Garcia-Castellanos, D., and Villasenor, A., 2011, Messinian salinity crisis regulated by competing tectonics and erosion at the
Gibraltar arc: Nature 480, 359–363.
Garnero, E.J., McNamara, A.K., 2008. Structure and Dynamics of Earth’s Lower Mantle. Science 320, 626 –628.
Garzanti, E., Hu, X., 2015. Latest Cretaceous Himalayan Tectonics: Obduction, Collison Or Deccan-Related Uplift?.
Gondwana Research 28 165-178.
Gaschnig, R.M., Vervoort, J.D., Tikoff, B. and Lewis, R.S., 2017. Construction and preservation of batholiths in the northern
US Cordillera. Lithosphere, 9(2), 315-324.
Gehrels, G., Rusmore, M., Woodsworth, G., Crawford, M., Andronicos, C., Hollister, L., Patchett, J., Ducea, M., Butler, R.,
Klepeis, K. and Davidson, C., 2009. U-Th-Pb geochronology of the Coast Mountains batholith in north-coastal British
Columbia: Constraints on age and tectonic evolution. Geological Society of America Bulletin, 121(9-10), 1341-1361.
Geist, E.L., Vallier, T.L., Scholl, D.W., 1994. Origin, transport, and emplacement of an exotic island-arc terrane exposed in
eastern Kamchatka, Russia. Geological Society of America Bulletin 106, 1182–1194.
Ghiglione, M.C., Yagupsky, D., Ghidella, M., Ramos, A., 2008. Continental stretching preceding the opening of the Drake
Passage: Evidence from Tierra del Fuego. Geology 36, 643–645.
Ghose, N.C., Agrawal, O.P., Chatterjee, N., 2010. Geological and mineralogical study of eclogite and glaucophane schists in
the Naga Hills Ophiolite, Northeast India. Island Arc 19, 336–356.
Giacosa, R.E., Heredia, N., C., 2004. Structure of the North Patagonian thick-skinned fold-and-thrust belt, southern central
Andes, Argentina (41°–42°S). Journal of South American Earth Sciences 18, 61–72.
Gibbs, M.T., Bluth, G.J., Fawcett, P.J., Kump, L.R., 1999. Global chemical erosion over the last 250 My: variations due to
changes in paleogeography, paleoclimate, and paleogeology. American Journal of Science 299, 611–651.
Girard, J., Amulele, G., Farla, R., Mohiuddin, A. and Karato, S.I., 2016. Shear deformation of bridgmanite and
magnesiowüstite aggregates at lower mantle conditions. Science, 351(6269), 144-147.
Gnos, E., Immenhauser, A., Peters, T., 1997. Late Cretaceous/early Tertiary convergence between the Indian and Arabian
plates recorded in ophiolites and related sediments. Tectonophysics 271, 1–19.
Gnos, E., Khan, M., Mahmood, K., Khan, A.S., Shafique, N.A., Villa, I.M., 1998. Bela oceanic lithosphere assemblage and its
relation to the ReÂunion hotspot. Terra Nova 10, 90–95.
Goddéris, Y., Donnadieu, Y., Le Hir, G., Lefebvre, V., Nardin, E., 2014. The role of palaeogeography in the Phanerozoic
history of atmospheric CO2 and climate. Earth-Sci. Rev. 128, 122–138. doi:10.1016/j.earscirev.2013.11.004
Goddéris, Y., François, L.M., 1995. The Cenozoic evolution of the strontium and carbon cycles: relative importance
of continental erosion and mantle exchanges. Chem. Geol., The Mantle-Ocean connection 126, 169–190.
doi:10.1016/0009-2541(95)00117-3
Goddéris, Y., Le Hir, G., Macouin, M., Donnadieu, Y., Hubert-Théou, L., Dera, G., Aretz, M., Fluteau, F., Li, Z.X.,
Halverson, G.P., 2016. Paleogeographic forcing of the strontium isotopic cycle in the Neoproterozoic. Gondwana Res.
doi:10.1016/j.gr.2016.09.013
Goes, S., Cammarano, F. and Hansen, U., 2004. Synthetic seismic signature of thermal mantle plumes. Earth and Planetary
Science Letters, 218(3), pp.403-419.
Goes, S., Capitanio, F.A. and Morra, G., 2008. Evidence of lower-mantle slab penetration phases in plate motions. Nature,
451(7181), 981-984.
Golonka, J., 2007a. Phanerozoic Paleoenvironment and Paleolithofacies Maps : late Paleozoic. Geol. Akad. Gór.-Hut. Im
Stanisława Staszica W Krakowie T. 33, z. 2, 145–209.
Golonka, J., 2007b. Phanerozoic Paleoenvironment and Paleolithofacies Maps : Mesozoic. Geol. Akad. Gór.-Hut. Im
Stanisława Staszica W Krakowie T. 33, z. 2, 211–264.
Golonka, J., 2007c. Late Triassic and Early Jurassic paleogeography of the world. Paleogeogr. Paleoclimatol. Paleoecol. 244,
297–307. doi:doi: 10.1016/j.paleo.2006.06.041
Golonka, J., 2009a. Phanerozoic Paleoenvironment and Paleolithofacies Maps : Cenozoic. Geol. Akad. Gór.-Hut. Im
Stanisława Staszica W Krakowie T. 35, z. 4, 507–587.

187
Golonka, J., 2009b. Phanerozoic paleoenvironment and paleolithofacies maps : Early Paleozoic. Geol. Akad. Gór.-Hut. Im
Stanisława Staszica W Krakowie T. 35, z. 4, 589–654.
Golonka, J., 2012. Paleozoic paleoenvironment and paleolithofacies maps of Gondwana. AGH University of Science and
Technology Press, Kraków.
Golonka, J., Bocharova, N.Y., Ford, D., Edrich, M.E., 2003. Paleogeographic reconstructions and basins development of the
Arctic. Marine and Petroleum Geology 20, 211-248.
Golonka, J., Ross, M.I., Scotese, C.R., 1994. Phanerozoic Paleogeographic and Paleoclimatic Modeling Maps. Pangea Glob.
Enironments Resour. - Mem., Canadian Society of Petroleum Geologists Special Publications 17, 1–47.
Gong, Z., Langereis, C.G., Mullender, T.A.T., 2008. The rotation of Iberia during the Aptian and the opening of the Bay of
Biscay. Earth and Planetary Science Letters 273, 80–93.
González-León, C.M., Solari, L., Solé, J., Ducea, M.N., Lawton, T.F., Bernal, J.P., Becuar, E.G., Gray, F., Martínez, M.L. and
Santacruz, R.L., 2011. Stratigraphy, geochronology, and geochemistry of the Laramide magmatic arc in north-central
Sonora, Mexico. Geosphere, GES00679-1.
Gorbatov, A., Kennett, B.L.N., 2003. Joint bulk-sound and shear tomography for Western Pacific subduction zones. Earth
and Planetary Science Letters 210, 527–543.
Gorbatov, A., Widiyantoro, S., Fukao, Y., Gordeev, E., 2000. Signature of remnant slabs in the North Pacific from P‐wave
tomography. Geophysical Journal International 142, 27–36.
Gorman, P.J., Kerrick, D.M., Conolly, J.A.D., 2006. Modeling open system metamorphic decarbonation of subducting slabs.
Geochemistry Geophysics Geosystems, v. 7, Q04007.
Govers, R., Wortel, M.J.R., 2005. Lithosphere tearing at STEP faults: Response to edges of subduction zones. Earth and
Planetary Science Letters 236, 505–523.
Grand, S., van der Hilst, R.D., Widiyantoro, S., 1997. Global seismic tomography: a snapshot of convection in the earth. GSA
Today 7, 1–7.
Gurnis, M., 1988. Large-scale mantle convection and the aggregation and dispersal of supercontinents. Nature 332, 695–699.
doi:10.1038/332695a0
Gurnis, M., 1990. Bounds on global dynamic topography from Phanerozoic flooding of continental platforms. Nature 344,
754–756. doi:10.1038/344754a0
Gurnis, M., 1992. Rapid Continental Subsidence Following the Initiation and Evolution of Subduction. Science 255, 1556–
1558. doi:10.1126/science.255.5051.1556
Gurnis, M., Moresi, L., Dietmar Müller, R., 2000. Models of mantle convection incorporating plate tectonics: The Australian
region since the Cretaceous, in The History and Dynamics of Global Plate Motions, Geophysical Monograph Series,
American Geophysical Union, Washington, D. C., 211–238.
Gutiérrez-Alonso, G., Fernandez-Suarez, J., Weil, A.B., Brendan Murphy, J., Damian Nance, R., Corfu, F., Johnston, S.T.,
2008. Self-subduction of the Pangaean global plate. Nature Geoscience 1, 549–553.
Gutscher, M.A., Malod, J., Rehault, J.P., Contrucci, I., Klingelhoefer, F., Mendes-Victor, L., Spakman, W., 2002. Evidence for
active subduction beneath Gibraltar. Geology 30, 1071–1074.
Hacker, B.R., 1990. Simulation of the metamorphic and deformational history of the metamorphic sole of the Oman
ophiolite. Journal of Geophysical Research 95, 4895–4907.
Hacker, B.R., 1994. Rapid Emplacement of Young Oceanic Lithosphere: Argon Geochronology of the Oman Ophiolite.
Science 265, 1563–1565.
Hacker, B.R., Mosenfelder, J.L., Gnos, E., 1996. Rapid emplacement of the Oman ophiolite: Thermal and geochronologic
constraints. Tectonics 15, 1230–1247.
Hadlari,T., Midwinter, D., Poulton, T.P., Matthews, A., 2017. A Pangean rim of fire: Reviewing the Triassic of western
Laurentia. Lithosphere, L643-1.
Hafkenscheid, E., Buiter, S.J.H., Wortel, M.J.R., Spakman, W., Bijwaard, H., 2001. Modelling the seismic velocity structure
beneath Indonesia: a comparison with tomography. Tectonophysics 333, 35–46.

188
Hafkenscheid, E., Wortel, M.J.R., Spakman, W., 2006. Subduction history of the Tethyan region derived from seismic
tomography and tectonic reconstructions. Journal of Geophysical Research 111, B08401–doi: 10.1029–2005JB003791.
Hager, B.H., Clayton, R.W., Richards, M.A., Comer, R.P. and Dziewonski, A.M., 1984. Lower mantle heterogeneity,
dynamic topography and the geoid.
Hall, C.M., Kesler, S.E., Russell, N., Piñero, E., Sánchez, R., Pérez, M., Moreira, J., Borges, M., 2004. Age and tectonic
setting of the Camaguey volcanic-intrusice arc, Cuba: Cretaceous extension and uplift of the Greater Antilles. Journal of
Geology 112, 521–542.
Hall, R., 2002. Cenozoic geological and plate tectonic evolution of SE Asia and the SW Pacific: computer-based
reconstructions, model and animations. Journal of Asian Earth Sciences 20, 353–431.
Hall, R., 2012. Late Jurassic–Cenozoic reconstructions of the Indonesian region and the Indian Ocean. Tectonophysics 570-
571, 1–41.
Hall, R., Spakman, W., 2002. Subducted slabs beneath the eastern Indonesia-Tonga region: insights from tomography. Earth
and Planetary Science Letters 201, 321–336.
Hall, R., Spakman, W., 2004. Mantle structure and tectonic evolution of the region north and east of Australia. Geological
Society of Australia Special Publication 22, 361–381.
Hall, R., Spakman, W., 2015. Mantle structure and tectonic history of SE Asia, Tectonophysics 658, 14-45.
Hall, R., van Hattum, M.W.A., Spakman, W., 2008. Impact of India–Asia collision on SE Asia: The record in Borneo.
Tectonophysics 451, 366–389.
Hallam, A., 1984. Pre-Quaternary Sea-Level Changes. Annu. Rev. Earth Planet. Sci. 12, 205–243. doi:10.1146/annurev.
ea.12.050184.001225
Hallam, A., 2001. A review of the broad pattern of Jurassic sea-level changes and their possible causes in the light of current
knowledge. Palaeogeogr. Palaeoclimatol. Palaeoecol. 167, 23–37. doi:10.1016/S0031-0182(00)00229-7
Handy, M.R., Schmid, S.M., Bousquet, R., Kissling, E., Bernoulli, D., 2010. Reconciling plate-tectonic reconstructions of
Alpine Tethys with the geological–geophysical record of spreading and subduction in the Alps. Earth-Science Reviews
102, 121–158.
Handy, M.R., Ustaszewski, K., Kissling, E., 2014. Reconstructing the Alps–Carpathians–Dinarides as a key to understanding
switches in subduction polarity, slab gaps and surface motion. International Journal of Earth Sciences 104, 1–26.
Haq, B.U., 2014. Cretaceous eustasy revisited. Glob. Planet. Change 113, 44–58. doi:10.1016/j.gloplacha.2013.12.007
Haq, B.U., Hardenbol, J., Vail, P.R., 1987. Chronology of Fluctuating Sea Levels Since the Triassic. Science 235, 1156–1167.
doi:10.1126/science.235.4793.1156
Haq, B.U., Hardenbol, J., Vail, P.R., 1988. Mesozoic and Cenozoic Chronostratigraphy and Cycles of Sea-Level Change.
SEPM Spec. Publ. 42, 71–108.
Haq, B.U., Schutter, S.R., 2008. A Chronology of Paleozoic Sea-Level Changes. Science 322, 64–68. doi:10.1126/
science.1161648
Harbert, W., Sokolov, S., Alexutin, M., Krylov, K., Grigoriev, V., Heiphetz, A., 2003. Reconnaissance paleomagnetism of
Late Triassic blocks, Kuyul region, Northern Kamchatka Peninsula, Russia. Tectonophysics 361, 215–227. doi:doi: DOI:
10.1016/S0040-1951 (02)00591-7
Hardenbol, J., Thierry, J., Farley, M.B., Jacquin, T., de Graciansky, P.-C., Vail, P.R., 1998. Mesozoic and Cenozoic Sequence
Stratigraphy of European Basins. SEPM Spec. Publ. 60, 3–13.
Hauff, G., Hoernle, K., Tilton, G., Graham, D.W., Kerr, A.C., 2000. Large volume recycling of oceanic lithosphere over short
time scales: geochemical contraints from the Caribbean Large Igneous Province. Earth and Planetary Science Letters
174, 247–263.
He, Y., Wen, L., 2009. Structural features and shear-velocity structure of the “Pacific Anomaly”. Journal of Geophysical
Research 114, B02309–17, doi: 10.1029/2008JB005814.

189
Hébert, R., Bezard, R., Guilmette, C., Dostal, J., Wang, C., Liu, Z.F., 2012. The Indus–Yarlung Zangbo ophiolites from
Nanga Parbat to Namche Barwa syntaxes, southern Tibet: First synthesis of petrology, geochemistry, and geochronology
with incidences on geodynamic reconstructions of Neo-Tethys. Gondwana Research 22, 377–397.
Heine, C., Yeo, L.G., Müller, R.D., 2015. Evaluating global paleoshoreline models for the Cretaceous and Cenozoic. Aust. J.
Earth Sci. 62, 275–287. doi:10.1080/08120099.2015.1018321
Herold, N., Seton, M., Müller, R.D., You, Y., Huber, M., 2008. Middle Miocene tectonic boundary conditions for use in
climate models. Geochem. Geophys. Geosystems 9, Q10009. doi:10.1029/2008GC002046
Hodell, D.A., Mead, G.A., Mueller, P.A., 1990. Variation in the strontium isotopic composition of seawater (8 Ma to
present) : Implications for chemical weathering rates and dissolved fluxes to the oceans. Chem. Geol. Isot. Geosci. Sect.
80, 291–307. doi:10.1016/0168-9622(90)90011-Z
Hoisch, T.D., Wells, M.L., Beyene, M.A., Styger, S., Vervoort, J.D., 2014. Jurassic Barrovian metamorphism in a western U.S.
Cordilleran metamorphic core complex, Funeral Mountains, California. Geology 42, 399–402.
Holm, R.J., Spandler, C., Richards, S.W., 2013. Melanesian arc far-field response to collision of the Ontong Java Plateau:
Geochronology and petrogenesis of the Simuku Igneous Complex, New Britain, Papua New Guinea. Tectonophysics
603, 189–212.
Homke, S., Verges, J., Serra-Kiel, J., Bernaola, G., Sharp, I.R., Garcés, M., Monetro-Verdu, I., Karpuz, R., H, G.M., 2009.
Late Cretaceous–Paleocene formation of the proto-Zagros foreland basin, Lurestan Province, SW Iran. Geological
Society of America Bulletin 121, 963–978.
Honda, S., 2016. Slab stagnation and detachment under northeast China. Tectonophysics 671, pp.127-138.
Horvath, F., Bada, G., Szafian, P., Tari, G., Adam, A., Cloetingh, S., 2006. Formation and deformation of the Pannonian
Basin: constraints from observational data. Geological Society, London, Memoirs 32, 191–206.
Hourigan, J.K., Brandon, M.T., Soloviev, A.V., Kirmasov, A.B., Garver, J.I., Stevenson, J., Reiners, W., 2009. Eocene arc-
continent collision and crustal consolidation in Kamchatka, Russian Far East. American Journal of Science 309,
333–396.
Houser, C., Masters, G., Shearer, P., Laske, G., 2008. Shear and compressional velocity models of the mantle from
cluster analysis of long-period waveforms. Geophysical Journal International 174, 195–212. doi:10.1111/j.1365-
246X.2008.03763.x
Huang, H.-H., Wu, Y.-M., Song, X., Chang, C.-H., Lee, S.-J., Chang, T.-M., Hsieh, H.-H., 2014. Joint Vp and Vs
tomography of Taiwan: Implications for subduction-collision orogeny. Earth and Planetary Science Letters 392, 177–
191.
Huang, J., Zhao, D., 2006. High-resolution mantle tomography of China and surrounding regions. Journal of Geophysical
Research 111, B09305–doi:10.1029–2005JB004066.
Huang, W., van Hinsbergen, D.J.J., Maffione, M., Orme, D.A., Dupont-Nivet, G., Guilmette, C., Ding, L., Guo, Z., Kapp, P.,
2015. Lower Cretaceous Xigaze ophiolites formed in the Gangdese forearc: Evidence from paleomagnetism, sediment
provenance, and stratigraphy. Earth and Planetary Science Letters 415, 142–153.
Hubbard, R.J., 1988. Age and Significance of Sequence Boundaries on Jurassic and Early Cretaceous Rifted Continental
Margins. AAPG Bull. 72, 49–72.
Humphreys, E. 2009. Relation of flat subduction to magmatism and deformation in the western United States. Geological
Society of America Memoirs 204, 85-98.
Hüsing, S.K., Zachariasse, W.J., van Hinsbergen, D.J.J., Krijgsman, W., Inceoz, M., Harzhauser, M., Mandic, O., Kroh, A.,
2009. Oligocene-Miocene basin evolution in SE Anatolia, Turkey: constraints on the closure of the eastern Tethys
gateway. Geological Society, London, Special Publications 311, 107–132.
Hutko, A.R., Lay, T., Garnero, E.J., Revenaugh, J., 2006. Seismic detection of folded subducted lithosphere at the core-mantle
boundary. Nature 441, 333–336.

190
Ishizuka, O., Tani, K., Reagan, M.K., Kanayama, K., Umino, S., Harigane, Y., Sakamoto, I., Miyajima, Y., Yuasa, M., Dunkley,
D.J., 2011. The timescales of subduction initiation and subsequent evolution of an oceanic island arc. Earth and Planetary
Science Letters 306, 229–240.
Isozaki, Y., Aoki, K., Nakama, T., 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, 82–105.
Isozaki, Y., Maruyama, S., Furuoka, F., 1990. Accreted oceanic materials in Japan. Tectonophysics 181, 179–205. doi:doi:
10.1016/0040-1951 (90)90016-2
Iturralde-Vinent, M.A., Díaz-Otero, C., García-Casco, A., van Hinsbergen, D.J.J., 2008. Paleogene foreland basin deposits
of North-Central Cuba: A record of arc-continent collision between the Caribbean and North American plates.
International Geology Review 50, 863–884.
Jackson, J., Priestley, K., Allen, M.B., Berberian, M., 2002. Active tectonics of the South Caspian Basin. Geophysical Journal
International 148, 214–245.
Jagoutz, O., Royden, L., Holt, A.F., and Becker, T.W., 2015. Anomalously fast convergence of India and Eurasia caused by
double subduction. Nature Geoscience 8, 475–478.
Jahn-Awe, S., Froitzheim, N., Nagel, T.J., Frei, D., Georgiev, N., Pleuger, J., 2010. Structural and geochronological evidence
for Paleogene thrusting in the Western Rhodopes (SW Bulgaria): Elements for a new tectonic model of the Rhodope
Metamorphic Province. Tectonics 29, TC3008–doi:10.1029–2009TC002558.
Jarvis, G.T., Lowman, J.P., 2007. Survival times of subducted slab remnants in numerical models of mantle flow. Earth and
Planetary Science Letters 260, 23–36.
Jaxybulatov, K., Koulakov, I., Dobretsov, N.L., 2013. Segmentation of the Izu-Bonin and Mariana slabs based on the analysis
of the Benioff seismicity distribution and regional tomography results. Solid Earth 4, 59–73.
Jerolmack, D.J., Paola, C., 2010. Shredding of environmental signals by sediment transport. Geophys. Res. Lett. 37, L19401.
doi:10.1029/2010GL044638
Jervey, M.T., 1988. Quantitative Geological Modeling of Siliciclastic Rock Sequences and Their Seismic Expression. SEPM
Spec. Publ. 42, 47–69.
Ji, W.-Q., Wu, F.-Y., Chung, S.-L., Li, J.-X., Liu, C.-Z., 2009. Chemical Geology. Chemical Geology 262, 229–245.
Jiang, G., Zhao, D., Zhang, G., 2009. Seismic tomography of the Pacific slab edge under Kamchatka. Tectonophysics 465,
190–203.
Jicha, B.R., Scholl, D.W., Singer, B.S., Yogodzinski, G.M., Kay, S.M., 2006. Revised age of Aleutian Island Arc formation
implies high rate of magma production. Geology 34, 661-664.
John, T., Gussone, N., Podladchikov, Y.Y., Bebout, G.E., Dohmen, R., Halama, R., Klemd, R., Magna, T., Seitz, H.M., 2012.
Volcanic arcs fed by rapid pulsed fluid flow through subducting slabs. Nature Geoscience 5, 489–492.
Johnston, F.K.B., Turchyn, A.V., Edmonds, M., 2011. Decarbonation efficiency in subduction zones: Implications for warm
Cretaceous climates. Earth and Planetary Science Letters 303, 143–152.
Johnston, S.T., Borel, G.D., 2007. The odyssey of the Cache Creek terrane, Canadian Cordillera: Implications for accretionary
orogens, tectonic setting of Panthalassa, the Pacific superwell, and break-up of Pangea. Earth and Planetary Science
Letters 253, 415–428.
Jolivet, L., Brun, J.-P., 2010. Cenozoic geodynamic evolution of the Aegean. International Journal of Earth Sciences 99,
109–138.
Jolivet, L., Faccenna, C., Piromallo, C., 2009. From mantle to crust: Stretching the Mediterranean. Earth and Planetary
Science Letters 285, 198–209.
Jones, R.E., De Hoog, J.C.M., Kirstein, L.A., Kasemann, S.A., Hinton, R., Elliott, T., Litvak, D., EIMF, 2014. Temporal
variations in the influence of the subducting slab on Central Andean arc magmas: Evidence from boron isotope
systematics. Earth and Planetary Science Letters 408, 390–401.
Jordan, T.E., Flemings, P.B., 1991. Large-scale stratigraphic architecture, eustatic variation, and unsteady tectonism: A
theoretical evaluation. J. Geophys. Res. Solid Earth 96, 6681–6699. doi:10.1029/90JB01399

191
Jourdan, S., Bernet, M., Tricart, P., Hardwick, E., Paquette, J.L., Guillot, S., Dumont, T., Schwartz, S., 2013. Short-
lived, fast erosional exhumation of the internal western Alps during the late early Oligocene: Constraints from
geothermochronology of pro-and retro-side foreland basin sediments. Lithosphere 5, 211–225.
Kakar, M.I., Collins, A.S., Mahmood, K., Foden, J.D., Khan, M., 2012. U-Pb zircon crystallization age of the Muslim Bagh
ophiolite: Enigmatic remains of an extensive pre-Himalayan arc. Geology 40, 1099–1102.
Kalyoncuoğlu, Ü.Y., Elitok, Ö., Dolmaz, M.N., Anadolu, N.C., 2011. Geophysical and geological imprints of southern
Neotethyan subduction between Cyprus and the Isparta Angle, SW Turkey. Journal of Geodynamics 52, 70–82.
Kaneshima, S., 2013. Lower mantle seismic scatterers below the subducting Tonga slab: Evidence for slab entrainment of
transition zone materials. Physics of The Earth and Planetary Interiors 222, 35–46.
Kaneshima, S., Helffrich, G., 2010. Small scale heterogeneity in the mid-lower mantle beneath the circum-Pacific area.
Physics of The Earth and Planetary Interiors 183, 91–103.
Kapp, P., DeCelles, G., Gehrels, G.E., Heizler, M., Ding, L., 2007. Geological records of the Cretaceous Lhasa-Qiangtang
and Indo-Asian collisions in the Nima basin area, central Tibet. Geological Society of America Bulletin 119, 917–933.
Kapp, P., Murphy, M.A., Yin, A., Harrison, T.M., Ding, L., Guo, J., 2003. Mesozoic and Cenozoic tectonic evolution of the
Shiquanhe area of western Tibet. Tectonics 22, doi: 10.1029/2001TC001332.
Karamata, S., 2006. The geological development of the Balkan Peninsula related to the approach, collision and compression of
Gondwanan and Eurasian units. Geological Society, London, Special Publications 260, 155–178.
Karaoğlan, F., Parlak, O., Klötzli, U., Thoni, M., Koller, F., 2012. U–Pb and Sm–Nd geochronology of the Kızıldağ (Hatay,
Turkey) ophiolite: implications for the timing and duration of suprasubduction zone type oceanic crust formation in the
southern Neotethys. Geological Magazine 150, 283–299.
Karaoğlan, F., Parlak, O., Klötzli, U., Thoni, M., Koller, F., 2014. U–Pb and Sm–Nd geochronology of the ophiolites from the
SE Turkey: implications for the Neotethyan evolution. Geodinamica Acta 25, 146–161.
Karato, S.I., 2010. Rheology of the Earth's mantle: A historical review. Gondwana Research, 18(1), pp.17-45.
Kashiwagi, H., 2016. Atmospheric carbon dioxide and climate change since the Late Jurassic (150 Ma) derived from a global
carbon cycle model. Palaeogeography, Palaeoclimatology, Palaeoecology 454, 82–90, doi: 10.1016/j.palaeo.2016.04.002
Kaymakci, N., Inceöz, M., Ertepinar, P., Koç, A., 2010. Late Cretaceous to Recent kinematics of SE Anatolia (Turkey).
Geological Society, London, Special Publications 340, 409–435.
Kellogg, J.N., Bonini, W.E., 1982. Subduction of the Caribbean plate and basement uplifts in the overriding South American
plate. Tectonics 1, 251–276.
Kennett, B.L.N., Engdahl, E.R., Buland, R., 1995. Constraints on seismic velocities in the Earth from travel times.
Geophysical Journal International 122, 108–124.
Kent, D.V., Irving, E., 2010. Influence of inclination error in sedimentary rocks on the Triassic and Jurassic apparent pole
wander path for North America and implications for Cordilleran tectonics. J Geophys Res 115, B10103.
Kent, D.V., Muttoni, G., 2008. Equatorial convergence of India and early Cenozoic climate trends. Proc. Natl. Acad. Sci. 105,
16065–16070. doi:10.1073/pnas.0805382105
Kerr, R.C. and Lister, J.R., 1991. The effects of shape on crystal settling and on the rheology of magmas. The Journal of
Geology, 99(3), 457-467.
Kerrick, D.M., 2001. Present and past nonanthropogenic CO2 degassing from the solid earth. Reviews of Geophysics, v. 39,
565–585.
Keskin, M., 2003. Magma generation by slab steepening and breakoff beneath a subduction-accretion complex: An alternative
model for collision-related volcanism in Eastern Anatolia, Turkey. Geophysical Research Letters 30, 8046–doi: 10.1029–
2003GL018019.
Khan, I., Clyde, W., 2013. Lower Paleogene Tectonostratigraphy of Balochistan: Evidence for Time-Transgressive Late
Paleocene-Early Eocene Uplift. Geosciences 3, 466–501.
King, S.D., 2016a. An evolving view of transition zone and midmantle viscosity. Geochemistry, Geophysics, Geosystems, 17,
pp.1234-1237.

192
King, S.D., 2016b. Reconciling laboratory and observational models of mantle rheology in geodynamic modelling. Journal of
Geodynamics, 100, pp.33-50.
Kissling, E., 1993. Deep structure of the Alps—what do we really know? Physics of The Earth and Planetary Interiors 79,
87–112,
Kissling, E., 2008. Deep structure and tectonics of the Valais—and the rest of the Alps. Bulletin Angewandte Geologie 13,
3–10.
Kissling, E., Spakman, W., 1996. Interpretation of tomographic images of uppermost mantle structure: examples from the
Western and Central Alps. Journal of Geodynamics 21, 97–111.
Kito, T., Rost, S., Thomas, C., Garnero, E.J., 2007. New insights into the P- and S-wave velocity structure of the D"
discontinuity beneath the Cocos plate. Geophysical Journal International 169, 631–645.
Kito, T., Thomas, C., Rietbrock, A., Garnero, E.J., Nippress, S.E.J., Heath, A.E., 2008. Seismic evidence for a sharp
lithospheric base persisting to the lowermost mantle beneath the Caribbean. Geophysical Journal International 174,
1019–1028.
Knesel, K.M., Cohen, B.E., Vasconcelos, M., Thiede, D.S., 2008. Rapid change in drift of the Australian plate records collision
with Ontong Java plateau. Nature 454, 754–757.
Ko, J.Y.T., Hung, S.H., Kuo, B.Y. and Zhao, L., 2017. Seismic evidence for the depression of the D″ discontinuity beneath the
Caribbean: Implication for slab heating from the Earth's core. Earth and Planetary Science Letters 467, 128-137.
Koç, A., Kaymakci, N., van Hinsbergen, D.J.J., Vissers, R.L.M., 2016. A Miocene onset of the modern extensional regime in
the Isparta Angle: constraints from the Yalvaç Basin (southwest Turkey). International Journal of Earth Sciences 434,
75-90.
Koglin, N., Kostopoulos, D., Reischmann, T., 2009. Geochemistry, petrogenesis and tectonic setting of the Samothraki mafic
suite, NE Greece: Trace-element, isotopic and zircon age constraints. Tectonophysics 473, 53-68.
König, M., Jokat, W., 2006. The Mesozoic breakup of the Weddell Sea. Journal of Geophysical Research 111, B12102.
Konstantinovskaia, E.A., 2001. Arc-continent collision and subduction reversal in the Cenozoic evolution of the Northwest
Pacifc: an example from Kamchatka (NE Russia). Tectonophysics 333, 75–94.
Koop, W., Stoneley, R., 1982. Subsidence history of the Middle East Zagros Basin: Permian to Recent. Philosophical
Transactions of the Royal Society of London ser. A 305, 149–168.
Koulakov, I., 2011. High-frequency Pand Svelocity anomalies in the upper mantle beneath Asia from inversion of worldwide
traveltime data. Journal of Geophysical Research 116, B04301, doi: 10.1029/2010JB007938.
Koulakov, I., 2013. Studying deep sources of volcanism using multiscale seismic tomography. Journal of Volcanology and
Geothermal Research 257, 205–226.
Koulakov, I., Wu, Y.-M., Huang, H.-H., Dobretsov, N., Jakovlev, A., Zabelina, I., Jaxybulatov, K., Chervov, V., 2014. Slab
interactions in the Taiwan region based on the P- and S-velocity distributions in the upper mantle. Journal of Asian
Earth Sciences 79, 53–64.
Koulakov, I., Zabelina, I., Amanatashvili, I., Meskhia, V., 2012. Nature of orogenesis and volcanism in the Caucasus region
based on results of regional tomography. Solid Earth 3, 327–337.
Koulakov, I.Y., Dobretsov, N., Bushenkova, N., Yakovlev, A.V., 2011. Slab shape in subduction zones beneath the Kurile–
Kamchatka and Aleutian arcs based on regional tomography results. Russian Geology and Geophysics 52, 650–667.
Kovalenko, D.V., 2003. Paleomagnetism of Geological Complexes of Kamchatka and Southern Koryakia. Tectonic and
Geophysical Interpretation [in Russian]. Nauchnyi Mir, Moscow.
Kufner, S.K., Schurr, B., Sippl, C., Schneider, F., Yuan, X., Ischuk, A., Arib, A., Murodkulov, S., Haberland, C., Mechie, J.,
Bianchi, M., Tilmann, F., 2014. Evidence for deeply subducting Asian lithosphere beneath the Pamir-Hindu Kush
region from lithospheric imaging. Geophysical Research Abstracts 16, EGU2014–3412–3.
Kufner, S.-K., Schurr, B., Sippl, C., Yuan, X., Ratschbacher, L., Akbar, A.S.O.M., Ischuk, A., Murodkulov, S., Schneider, F.,
Mechie, J., Tilmann, F., 2016. Deep India meets deep Asia: Lithospheric indentation, delamination and break-off under
Pamir and Hindu Kush (Central Asia). Earth and Planetary Science Letters, 435, 171–184.

193
Kurtz, A.C., Kump, L.R., Arthur, M.A., Zachos, J.C., Paytan, A., 2003. Early Cenozoic decoupling of the global carbon and
sulfur cycles. Paleoceanography 18, 1–14.
Lallemand, S., Font, Y., Bijwaard, H., Kao, H., 2001. New insights on 3-D plates interaction near Taiwan from tomography
and tectonic implications. Tectonophysics 335, 229–253.
Lambeck, K., Rouby, H., Purcell, A., Sun, Y. and Sambridge, M., 2014. Sea level and global ice volumes from the Last Glacial
Maximum to the Holocene. Proceedings of the National Academy of Sciences, 111(43), 15296-15303.
Lambeck, K., Smither, C. and Johnston, P., 1998. Sea-level change, glacial rebound and mantle viscosity for northern Europe.
Geophysical Journal International, 134(1), 102-144.
Lau, H.C., Mitrovica, J.X., Austermann, J., Crawford, O., Al‐Attar, D. and Latychev, K., 2016. Inferences of mantle viscosity
based on ice age data sets: Radial structure. Journal of Geophysical Research: Solid Earth, 121(10), 6991-7012.
Layer, W., Newberry, R., Fujita, K., Parfenov, L., Trunilina, V., Bakharev, A., 2001. Tectonic setting of the plutonic belts of
Yakutia, northeast Russia, based on 40A/39Ar geochronology and trace element geochemistry. Geology 29, 167–170.
Lee, C.-T.A., Shen, B., Slotnick, B.S., Liao, K., Dickens, G.R., Yokoyama, Y., Lenardic, A., Dasgupta, R., Jellinek, M., Lackey,
J.S., Schneider, T., Tice, M.M., 2012. Continental arc–island arc fluctuations, growth of crustal carbonates, and long-term
climate change. Geosphere GES00822.1. doi:10.1130/GES00822.1
Lefebvre, C., Meijers, M.J.M., Kaymakci, N., Peynircioğlu, A., Langereis, C.G., van Hinsbergen, D.J.J., 2013. Reconstructing
the geometry of central Anatolia during the late Cretaceous: Large-scale Cenozoic rotations and deformation between
the Pontides and Taurides. Earth and Planetary Science Letters 366, 83–98.
Lei, J., 2012. Upper-mantle tomography and dynamics beneath the North China Craton. Journal of Geophysical Research
117, B06313–doi:10.1029–2012JB009212.
Lei, J., Zhao, D., 2007. Teleseismic evidence for a break-off subducting slab under Eastern Turkey. Earth and Planetary
Science Letters 257, 14–28.
Lekic, V., Cottaar, S., Dziewonski, A., Romanowicz, B., 2012. Cluster analysis of global lower mantle tomography: A new
class of structure and implications for chemical heterogeneity. Earth and Planetary Science Letters 357–358, doi:
10.1016/j.epsl.2012.09.014
Leng, W. and Gurnis, M., 2012. Shape of thermal plumes in a compressible mantle with depth‐dependent viscosity.
Geophysical Research Letters, 39(5).
Levashova, N.M., 1999. Kinematics of Late Cretaceous and Cretaceous–Paleogene Oceanic Island Arcs [in Russian],
Author's Abstract, Candidate Thesis. GIN RAN.
Levashova, N.M., Shapiro, M.N., Beniamovsky, V.N. and Bazhenov, M.L., 2000. Paleomagnetism and geochronology of the
Late Cretaceous‐Paleogene island arc complex of the Kronotsky Peninsula, Kamchatka, Russia: Kinematic implications.
Tectonics, 19(5), pp.834-851.
Li, C., van der Hilst, R.D., 2010. Structure of the upper mantle and transition zone beneath Southeast Asia from traveltime
tomography. Journal of Geophysical Research 115, B07308–doi:10.1029–2009JB006882.
Li, C., van der Hilst, R.D., Engdahl, E.R., Burdick, S., 2008. A new global model for 3-D variations of P-wave velocity in the
Earth's mantle. Geochemistry Geophysics Geosystems, v. 9, Q05018.
Li, C., van der Hilst, R.D., Meltzer, A.S., Engdahl, E.R., 2008. Subduction of the Indian lithosphere beneath the Tibetan
Plateau and Burma. Earth and Planetary Science Letters 274, 157–168.
Li, J., Zhang, Y., Dong, S., Johnston, S.T., 2014a. Cretaceous tectonic evolution of South China: A preliminary synthesis.
Earth-Science Reviews 134, 98–136.
Li, S.-M., Zhu, D.-C., Wang, Q., Zhao, Z.-D., Sui, Q.-L., Liu, S.-A., Liu, D., Mo, X.-X., 2014b. Northward subduction of
Bangong–Nujiang Tethys: Insight from Late Jurassic intrusive rocks from Bangong Tso in western Tibet. Lithos 205,
284–297.
Li, Z., Ding, L., Lippert, C., Song, P., Yue, Y., van Hinsbergen, D.J.J., 2016. Paleomagnetic constraints on the Mesozoic drift
of the Lhasa terrane (Tibet) from Gondwana to Eurasia. Geology 44, 727-730.

194
Li, Z., Qiu, J.-S., Yang, X.-M., 2014c. A review of the geochronology and geochemistry of Late Yanshanian (Cretaceous)
plutons along the Fujian coastal area of southeastern China: Implications for magma evolution related to slab break-off
and rollback in the Cretaceous. Earth-Science Reviews 128, 232–248.
Li, Z.-X., Evans, D.A.D., Halverson, G.P., 2013. Neoproterozoic glaciations in a revised global palaeogeography from the
breakup of Rodinia to the assembly of Gondwanaland. Sediment. Geol. 294, 219–232. doi:10.1016/j.sedgeo.2013.05.016
Li, Z.-X., Li, X.-H., 2007. Formation of the 1300-km-wide intracontinental orogen and postorogenic magmatic province in
Mesozoic South China: A flat-slab subduction model. Geology 35, 179–182.
Li, Z.-X., Li, X.-H., Chung, S.-L., Lo, C.-H., Xu, X., Li, W.-X., 2012. Magmatic switch-on and switch-off along the South
China continental margin since the Permian: Transition from an Andean-type to a Western Pacific-type plate boundary.
Tectonophysics 532-535, 271–290.
Liati, A., Gebauer, D., Fanning, C.M., 2004. The age of ophiolitic rocks of the Hellenides (Vourinos, Pindos, Crete): first U–
Pb ion microprobe (SHRIMP) zircon ages. Chemical Geology 207, 171–188.
Lippitsch, R., Kissling, E., Ansorge, J., 2003. Upper mantle structure beneath the Alpine orogen from high-resolution
teleseismic tomography. Journal of Geophysical Research 108, 2376–doi:10.1029–2002Jb002016.
Lister, G., Kennett, B., Richards, S., Forster, M., 2008. Boudinage of a stretching slablet implicated in earthquakes beneath the
Hindu Kush. Nature Geoscience 1, 196–201. doi:10.1038/ngeo132
Lister, G.S., White, L.T., Hart, S., Forster, M.A., 2012. Ripping and tearing the rolling-back New Hebrides slab. Australian
Journal of Earth Sciences 59-6, 899-911. doi: 10.1080/08120099.2012.686454
Lithgow‐Bertelloni, C. and Richards, M.A., 1998. The dynamics of Cenozoic and Mesozoic plate motions. Reviews of
Geophysics, 36(1), pp.27-78.
Lithgow-Bertollini, C., Silver, P.G., 1998. Dynamic topography, plate driving forces and the African superswell. Nature 395,
269–272.
Liu, D., Huang, Q., Fan, S., Zhang, L., Shi, R., Ding, L., 2013. Subduction of the Bangong-Nujiang Ocean: constraints from
granites in the Bangong Co area, Tibet. Geological Journal 49, 188–206.
Liu, L., 2014. Constraining Cretaceous subduction polarity in eastern Pacific from seismic tomography and geodynamic
modeling. Geophysical Research Letters 41, 8029–8036.
Liu, L., Spasojevic, S., Gurnis, M., 2008. Reconstructing Farallon Plate subduction beneath North America back to the Late
Cretaceous. Science 322, 934–938.
Liu, L., Stegman, D.R., 2011. Segmentation of the Farallon slab. Earth and Planetary Science Letters 311, 1–10.
Livermore, R., Hillenbrand, C.-D., Meredith, M., Eagles, G., 2007. Drake Passage and Cenozoic climate: An open and shut
case?. Geochemistry, Geophysics, Geosystems 8, doi: 10.1029/2005GC001224.
Lodolo, E., Donda, F., Tassone, A., 2006. Western Scotia Sea margins: Improved constraints on the opening of the Drake
Passage. Journal of Geophysical Research 111, B06101–14, doi: 10.1029/2006JB004361.
Lonergan, L., White, N., 1997. Origin of the Betic‐Rif mountain belt. Tectonics 16, 504–522.
Long, M.D., 2016. The Cascadia Paradox: Mantle flow and slab fragmentation in the Cascadia subduction system. Journal of
Geodynamics, 102, pp.151-170.
Lucente, F.P., Chiarabba, C., Cimini, G.B., Giardini, D., 1999. Tomographic constraints on the geodynamic evolution of the
Italian region. Journal of Geophysical Research 104, 20307–20327.
Lucente, F.P., Speranza, F., 2001. Belt bending driven by lateral bending of subducting lithospheric slab: geophysical evidences
from the northern Apennines (Italy). Tectonophysics 337, 53–64.
Lustrino, M., Morra, V., Fedele, L., Franciosi, L., 2009. Beginning of the Apennine subduction system in central western
Mediterranean: Constraints from Cenozoic “orogenic” magmatic activity of Sardinia, Italy. Tectonics 28, TC5016–
doi:10.1029–2008TC002419.
Luyendyk, B.P., 1995. Hypothesis for Cretaceous rifting of east Gondwana caused by subducted slab capture. Geology 23,
373–376.

195
Ma, X., Sun, X., Wiens, D.A., Wen, L., Nyblade, A., Anandakrishnan, S., Aster, R., Huerta, A., Wilson, T., 2016. Strong
seismic scatterers near the core–mantle boundary north of the Pacific Anomaly. Physics of The Earth and Planetary
Interiors 253, 21–30.
Maffione, M., Thieulot, C., van Hinsbergen, D.J.J., Morris, A., Plumper, O., Spakman, W., 2015a. Dynamics of intraoceanic
subduction initiation: 1. Oceanic detachment fault inversion and the formation of supra-subduction zone ophiolites.
Geochemistry, Geophysics, Geosystems 16, 1753–1770.
Maffione, M., van Hinsbergen, D., de Gelder, G.I.N.O., van der Goes, F., Morris, A., 2017. Kinematics of subduction
initiation in the Neo-Tethys Ocean during the Late Cretaceous reconstructed from ophiolites of Turkey, Cyprus, and
Syria. Journal of Geophysical Research in press.
Maffione, M., van Hinsbergen, D.J.J., Koornneef, L.M.T., Guilmette, C., Hodges, K., Borneman, N., Huang, W., Ding, L.,
and Kapp, P., 2015b, Forearc hyperextension dismembered the south Tibetan ophiolites: Geology 43, 475–478.
Magganas, A., Sideris, C., Kokkinakis, A., 1991. Marginal basin—volcanic arc origin of metabasic rocks of the Circum-
Rhodope Belt, Thrace, Greece. Mineralogy and Petrology 44, 235–252.
Maggi, A., Priestley, K., 2005. Surface waveform tomography of the Turkish-Iranian plateau. Geophysical Journal
International 160, 1068–1080.
Mahmood, K., Boudier, F., Gnos, E., Monié, P., Nicolas, A., 1995. 40Ar/39Ar dating of the emplacement of the Muslim Bagh
ophiolite, Pakistan. Tectonophysics 250, 169–181.
Marquardt, H., Miyagi, L., 2015. Slab stagnation in the shallow lower mantle linked to an increase in mantle viscosity. Nature
Geoscience 8, 311–314. doi:10.1038/ngeo2393
Martin, A.K., 2007. Gondwana breakup via double-saloon-door rifting and seafloor spreading in a backarc basin during
subduction rollback. Tectonophysics 445, 245–272.
Martini, M., Mori, L., Solari, L., Centeno-García, E., 2011. Sandstone Provenance of the Arperos Basin (Sierra de
Guanajuato, Central Mexico): Late Jurassic–Early Cretaceous Back-Arc Spreading as the Foundation of the Guerrero
Terrane. The Journal of Geology 119, 597–617.
Martini, M., Solari, L., Camprubi, A., 2013. Kinematics of the Guerrero terrane accretion in the Sierra de Guanajuato, central
Mexico: new insights for the structural evolution of arc–continent collisional zones. International Geology Review 55,
574–589.
Martin-Short, R., Allen, R.M., Bastow, I.D., 2016. Subduction geometry beneath south central Alaska and its relationship to
volcanism. Geophysical Research Letters 43, doi: 10.1002/2016GL070580.
Marty, B., Tolstikhin, I.N., 1998. CO2 fluxes from mid-ocean ridges, arcs and plumes. Chemical Geology 145, 233–248.
Matenco, L., Bertotti, G., 2000. Tertiary tectonic evolution of the external East Carpathians (Romania). Tectonophysics 316,
255–286.
Matenco, L., Bertotti, G., Cloetingh, S., Dinu, C., 2003. Subsidence analysis and tectonic evolution of the external
Carpathian–Moesian Platform region during Neogene times. Sedimentary Geology 156, 71–94.
Matenco, L., Radivojević, D., 2012. On the formation and evolution of the Pannonian Basin: Constraints derived from the
structure of the junction area between the Carpathians and Dinarides. Tectonics 31, doi:10.1029/2012TC003206.
Matsuoka, A., 1996. Late Jurassic tropical Radiolaria: Vallupus and its related forms. Paleogeogr. Paleoclimatol. Paleoecol.
119, 359–369. doi:doi: 10.1016/0031-0182 (95)00018-6
Matthews, K.J., Williams, S.E., Whittaker, J.M., Müller, R.D., Seton, M., Clarke, G.L., 2015. Geologic and kinematic
constraints on Late Cretaceous to mid Eocene plate boundaries in the southwest Pacific. Earth-Science Reviews 140,
72–107.
Maurin, T., Rangin, C., 2009. Structure and kinematics of the Indo-Burmese Wedge: Recent and fast growth of the outer
wedge. Tectonics 28, TC2010–doi:10.1029–2008TC002276.
McArthur, J.M., Howarth, R.J., Shields, G.A., 2012. Strontium Isotope Stratigraphy, in: Gradstein, F.M., Ogg, J.G., Schmotz,
M.D., Ogg, G.M. (Eds.), The Geological Time Scale 2012. Elsevier, pp. 127–144.

196
Meffre, S., Falloon, T.J., Crawford, T.J., Hoernle, K., Hauff, F., Duncan, R.A., Bloomer, S.H., Wright, D.J., 2012. Basalts
erupted along the Tongan fore arc during subduction initiation: Evidence from geochronology of dredged rocks from the
Tonga fore arc and trench. Geochemistry, Geophysics, Geosystems 13, doi: 10.1029/2012GC004335.
Meijers, M.J.M, Hamers, M.F., van Hinsbergen, D.J.J., van der Meer, D.G., Kitchka, A., Langereis, C.G., and Stephenson,
R.A., 2010, New late Paleozoic paleopoles from the Donbas Foldbelt (Ukraine): implications for the Pangea A vs. B
controversy, Earth and Planetary Science Letters 297, p. 18-33
Meijers, M.J.M., Kaymakci, N., van Hinsbergen, D.J.J., Langereis, C.G., Stephenson, R.A., and Hippolyte, J.-C., 2010.
Late Cretaceous to Paleocene oroclinal bending in the central Pontides (Turkey): Tectonics 29, TC4016–doi:10.1029–
2009TC002620.
Meschede, M., Frisch, W., 1998. A plate-tectonic model for the Mesozoic and Early Cenozoic history of the Caribbean plate.
Tectonophysics 296, 269–291.
Miall, A.D., 1995. Whither stratigraphy? Sediment. Geol. 100, 5–20. doi:10.1016/0037-0738(95)00100-X
Miall, A.D., 2010. The Geology of Stratigraphic Sequences. Springer-Verlag, Berlin, Heidelberg.
Michard, A., Negro, F., Saddiqi, O., Bouybaouene, M.L., Chalouan, A., Montigny, R., Goffé, B., 2006. Pressure–temperature–
time constraints on the Maghrebide mountain building: evidence from the Rif–Betic transect (Morocco, Spain),
Algerian correlations, and geodynamic implications. Comptes Rendus Geoscience 338, 92–114.
Milidragovic, D., Joyce, N.L., Zagorevski, A., and Chapman, J.B., 2016. Petrology of explosive Middle-Upper Triassic
ultramafic rocks in the Mess Creek area, northern Stikine terrane. In: Geological Fieldwork 2015, British Columbia
Ministry of Energy and Mines, British Columbia Geological Survey Paper 2016-1, 95-111.
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., 2005. The Phanerozoic Record of Global Sea-Level Change. Science 310, 1293–1298.
doi:10.1126/science.1116412
Miller, M.S., Gorbatov, A., Kennett, B.L.N., 2005. Heterogeneity within the subducting Pacific slab beneath the Izu–Bonin–
Mariana arc: Evidence from tomography using 3D ray tracing inversion techniques. Earth and Planetary Science Letters
235, 331–342.
Miller, M.S., Gorbatov, A., Kennett, B.L.N., 2006. Three-dimensional visualization of a near-vertical slab tear beneath the
southern Mariana arc. Geochemistry, Geophysics, Geosystems 7, doi: 10.1029/2005GC001110.
Miller, M.S., Kennett, B.L.N., 2006. Evolution of mantle structure beneath the northwest Pacific: Evidence from seismic
tomography and paleogeographic reconstructions. Tectonics 25, 1–14.
Miller, M.S., Kennett, B.L.N., Lister, G.S., 2004. Imaging changes in morphology, geometry, physical properties of the
subducting Pacific plate along the Izu–Bonin–Mariana arc. Earth and Planetary Science Letters 224, 363–370.
Mills, B., Daines, S.J., Lenton, T.M., 2014. Changing tectonic controls on the long-term carbon cycle from Mesozoic to
present. Geochemistry, Geophysics, Geosystems, 15 doi: 10.1002/2014GC005530.
Mitrovica, J.X. and Forte, A.M., 2004. A new inference of mantle viscosity based upon joint inversion of convection and
glacial isostatic adjustment data. Earth and Planetary Science Letters, 225(1), 177-189.
Mohammadi, A., Burg, J.-P., Bouilhol, P., Ruh, J., 2016. U–Pb geochronology and geochemistry of Zahedan and Shah Kuh
plutons, southeast Iran: Implication for closure of the South Sistan suture zone. Lithos 248-251, 293–308.
Moix, P., Beccaletto, L., Kozur, H.W., Hochard, C., Rosselet, F., Stampfli, G.M., 2008. A new classification of the Turkish
terranes and sutures and its implication for the paleotectonic history of the region. Tectonophysics 451, 7–39.
Morley, C.K., 2002. A tectonic model for the Tertiary evolution of strike–slip faults and rift basins in SE Asia. Tectonophysics
347, 189–215.
Morra, G., Yuen, D.A., Boschi, L., Chatelain, P., Koumoutsakos, P. and Tackley, P.J., 2010. The fate of the slabs interacting
with a density/viscosity hill in the mid-mantle. Physics of the Earth and Planetary Interiors, 180(3), 271-282.
Morris, A.,erson, M.W., Robertson, A.H., Al-Riyami, K., 2002. Extreme tectonic rotations within an eastern Mediterranean
ophiolite (Baër–Bassit, Syria). Earth and Planetary Science Letters 202, 247–261.

197
Mortensen, J.K., Hall, B.V., Bissig, T., Friedman, R.M., Danielson, T., Oliver, J., Rhys, D.A., Ross, K.V., Gabites, J.E., 2008.
Age and Paleotectonic Setting of Volcanogenic Massive Sulfide Deposits in the Guerrero Terrane of Central Mexico:
Constraints from U-Pb Age and Pb Isotope Studies. Economic Geology 103, 117–140.
Moucha, R., Forte, A.M., Mitrovica, J.X., Rowley, D.B., Quéré, S., Simmons, N.A., Grand, S.P., 2008. Dynamic topography
and long-term sea-level variations: There is no such thing as a stable continental platform. Earth Planet. Sci. Lett. 271,
101–108. doi:10.1016/j.epsl.2008.03.056
Müller, R.D., Dutkiewicz, A., Seton, M., Gaina, C., 2013. Seawater chemistry driven by supercontinent assembly, breakup,
and dispersal. Geology 20, 907–910. doi:10.1130/G34405.1
Müller, R.D., Royden, J.-Y., Lawver, L.A., 1993. Revised plate motions relative to the hotspots from combined Atlantic and
Indian Ocean hotspot tracks. Geology 21, 275–278.
Müller, R.D., Sdrolias, M., Gaina, C., Roest, W.R., 2008. Age, spreading rates, and spreading asymmetry of the world's ocean
crust. Geochemistry, Geophysics, Geosystems 9, Q04006–doi: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. doi:10.1126/science.1151540
Müller, R.D., Seton, M., and Zahirovic, S., Williams, S.E., Matthews, K.J., Wright, N.M., Shephard, G.E., Maloney, K.T.,
Barnett-Moore, N., Hosseinpour, M., Bower, D.J., and Cannon, J.S., 2016. Ocean Basin Evolution and Global-Scale
Plate Reorganization Events Since Pangea Breakup: Annual Review of Earth and Planetary Sciences 44, 107–138.
Mumladze, T., Forte, A.M., Cowgill, E.S., Trexler, C.C., Niemi, N.A., Yıkılmaz, M.B., Kellogg, L.H., 2015. Subducted,
detached, torn slabs beneath the Greater Caucasus. GEORESJ 5, 36–46.
Munizaga, F., Hervé, F., Drake, R.E., Pankhurst, R.J., Brook, M., Snelling, N., 2002. Geochronology of the Lake Region of
south-central Chile (39°-42°S): Preliminary results. Journal of South American Earth Sciences 1, 309–316.
Muto, T., 2001. Shoreline Autoretreat Substantiated in Flume Experiments. J. Sediment. Res. 71, 246–254.
doi:10.1306/091400710246
Muto, T., Steel, R.J., 1997. Principles of Regression and Transgression: The Nature of the Interplay Between
Accommodation and Sediment Supply: PERSPECTIVES. J. Sediment. Res. 67. doi:10.1306/D42686A8-2B26-11D7-
8648000102C1865D
Muto, T., Steel, R.J., 2014. The autostratigraphic view of responses of river deltas to external forcing, in: Martinius, A.W.,
Ravnås, R., Howell, J.A., Steel, R.J., Wonham, J.P. (Eds.), From Depositional Systems to Sedimentary Successions on the
Norwegian Continental Margin. John Wiley & Sons, Ltd, pp. 139–148.
Muto, T., Steel, R.J., Swenson, J.B., 2007. Autostratigraphy: A Framework Norm for Genetic Stratigraphy. J. Sediment. Res.
77, 2–12. doi:10.2110/jsr.2007.005
Muttoni, G., Mattei, M., Balini, M., Zanchi, A., Gaetani, M., Berra, F., 2009. The drift history of Iran from the Ordovician to
the Triassic. Geological Society, London, Special Publications 312, 7–29.
Myron G Best, E.H.C., 1991. Limited extension during peak Tertiary volcanism, Great Basin of Nevada and Utah. Journal of
Geophysical Research 96, 13509–13528.
Nance, R.D., Murphy, J.B., Santosh, M., 2014. The supercontinent cycle: A retrospective essay. Gondwana Res. 25, 4–29.
doi:10.1016/j.gr.2012.12.026
Natal'in, B.A., Sunal, G., Satir, M., Toraman, E., 2012. Tectonics of the Strandja Massif, NW Turkey: History of a long-lived
arc at the northern margin of Paleo-Tethys. Turkish Journal of Earth Sciences 21, 755–798.
Negredo, A.M., Replumaz, A., Villaseñor, A., Guillot, S., 2007. Modeling the evolution of continental subduction processes in
the Pamir–Hindu Kush region. Earth and Planetary Science Letters 259, 212–225.
Neill, I., Kerr, A.C., Hastie, A.R., Pindell, J.L., Millar, I.L., Atkinson, N., 2012. Age and Petrogenesis of the Lower
Cretaceous North Coast Schist of Tobago, a Fragment of the Proto–Greater Antilles Inter-American Arc System. The
Journal of Geology 120, 367–384.

198
Neill, I., Kerr, A.C., Hastie, A.R., Stanek, K.P., Millar, I.L., 2011. Origin of the Aves Ridge and Dutch-Venezuelan Antilles:
interaction of the Cretaceous “Great Arc” and Caribbean-Colombian Oceanic Plateau?. Journal of the Geological
Society 168, 333–348.
Nerlich, R., Clark, S.R., Bunge, H.-P., 2013. The Scotia Sea gateway: No outlet for Pacific mantle . Tectonophysics 604,
41–50.
Niu, F., Wen, L., 2001. Strong seismic scatterers near the core-mantle boundary west of Mexico. Geophysical Research
Letters 28, 3557–3560.
Nokleberg, W.J., Parfenov, L.M., Monger, J.W.H., Norton, I.O., Khanchuk, A., Stone, D.B., Scotese, C.R., Scholl, D.W.
Fujita, K., 2000. Phanerozoic tectonic evolition of the circum-north Pacific. USGS Professional Paper 1626, 1-122.
Norton, I.O., 1996. Global hotspot reference frames and plate motion, in: The History of Global Plate Motions, Geophysical
Monograph. American Geophysical Union, Washington D.C., pp. 339–357.
Nur, A., Ben-Avraham, Z., 1977. Lost Pacifica continent. Nature 270, 41–43. doi:10.1038/270041a0
Obayashi, M., Yoshimitsu, J., Fukao, Y., 2009. Tearing of Stagnant Slab. Science 324, 1173–1175.
Obayashi, M., Yoshimitsu, J., Nolet, G., Fukao, Y., Shiobara, H., Sugioka, H., Miyamachi, H., Gao, Y., 2013. Finite frequency
whole mantle Pwave tomography: Improvement of subducted slab images. Geophysical Research Letters 40, 5652–5657.
Obrebski, M., Allen, R.M., Xue, M., Hung, S.-H., 2010. Slab-plume interaction beneath the Pacific Northwest. Geophysical
Research Letters 37, doi: 10.1029/2010GL043489.
Oda, H., Suzuki, H., 2000. Paleomagnetism of Triassic and Jurassic red bedded chert of the Inuyama area, central Japan. J
Geophys Res 105, 25743–25767.
Okay, A.I., Monié, P., 1997. Early Mesozoic subduction in the Eastern Mediterranean. Evidence from Triassic eclogite in
northwest Turkey. Geology 25, 595–598.
Okay, A.I., Monod, O., Monié, P., 2002. Triassic blueschists and eclogites from northwest Turkey: vestiges of the Paleo-
Tethyan subduction. Lithos 64, 155–178.
Okay, A.I., Nikishin, A.M., 2015. Tectonic evolution of the southern margin of Laurasia in the Black Sea region.
International Geology Review 57, 1051-1076.
Okay, A.I., Noble, J., Tekin, U.K., 2011. Devonian radiolarian ribbon cherts from the Karakaya Complex, Northwest Turkey:
Implications for the Paleo-Tethyan evolution . Comptes rendus - Palevol 10, 1–10.
Okay, A.I., Zattin, M., Cavazza, W., 2010. Apatite fission-track data for the Miocene Arabia-Eurasia collision. Geology 38,
35–38.
Oliva, P., Viers, J., Dupré, B., 2003. Chemical weathering in granitic environments. Chem. Geol., Controls on Chemical
Weathering 202, 225–256. doi:10.1016/j.chemgeo.2002.08.001
Otto-Bliesner, B.L., 1995. Continental drift, runoff, and weathering feedbacks: Implications from climate model experiments.
J. Geophys. Res. Atmospheres 100, 11537–11548. doi:10.1029/95JD00591
Oxman, V.S., 2003. Tectonic evolution of the Mesozoic Verkhoyansk-Kolyma belt (NE Asia). Tectonophysics 365, 45–76.
doi:doi: DOI: 10.1016/S0040-1951 (03)00064-7
Pang, K.-N., Chung, S.-L., Zarrinkoub, M.H., Khatib, M.M., Mohammadi, S.S., Chiu, H.-Y., Chu, C.-H., Lee, H.-Y., Lo,
C.-H., 2013. Eocene–Oligocene post-collisional magmatism in the Lut–Sistan region, eastern Iran: Magma genesis and
tectonic implications. Lithos 180-181, 234–251.
Pankhurst, R.J., Rapela, C.W., Fanning, C.M., Márquez, M., 2006. Gondwanide continental collision and the origin of
Patagonia. Earth-Science Reviews 76, 235–257.
Paola, C., Heller, P.L., Angevine, C.L., 1992. The large-scale dynamics of grain-size variation in alluvial basins, 1: Theory.
Basin Res. 4, 73–90. doi:10.1111/j.1365-2117.1992.tb00145.x
Parfenov, L.M., Badarch, G., A, B.N., Khanchuk, A., Kuzmin, M.I., Nokleberg, W.J., Prokopiev, A.V., Ogasawara, M., Yan,
H., 2009. Summary of Northeast Asia geodynamics and tectonics. Stephan Mueller Special Publication Series 4, 11–33.
Park, J., Royer, D.L., 2011. Geologic constraints on the glacial amplification of Phanerozoic climate sensitivity. American
Journal of Science 211, 1–26.

199
Parkinson, N., Summerhayes, C., 1985. Synchronous Global Sequence Boundaries. AAPG Bull. 69, 685–687.
Parsons, B., Sclater, J.G., 1977. An analysis of the variation of ocean floor bathymetry and heat flow with age. J. Geophys. Res.
82, 803–827. doi:10.1029/JB082i005p00803
Patzelt, A., Li, H., Wang, J., Appel, E., 1996. Palaeomagnetism of Cretaceous to Tertiary sediments from southern Tibet:
evidence for the extent of the northern margin of India prior to the collision with Eurasia. Tectonophysics 259, 259–284.
Pavlis, G.L., Das, S., 2000. The Pamir-Hindu Kush seismic zone as a strain marker for flow in the upper mantle. Tectonics 19,
103–115.
Payton, C.E., 1977. Seismic Stratigraphy: Applications to Hydrocarbon exploration, AAPG Memoir. American Association
of Petroleum Geologists, Tulsa, Oklahoma.
Pegler, G., Das, S., 1998. An enhanced image of the Pamir-Hindu Kush seismic zone from relocated earthquake hypocentres.
Geophysical Journal International 134, 573–595.
Peltier, W.R. and Drummond, R., 2010. Deepest mantle viscosity: Constraints from Earth rotation anomalies. Geophysical
Research Letters, 37(12).
Peltier, W.R., 1998. Postglacial variations in the level of the sea: Implications for climate dynamics and solid‐earth geophysics.
Reviews of Geophysics, 36(4), 603-689.
Pesicek, J.D., Thurber, C.H., Widiyantoro, S., Zhang, H., DeShon, H.R., Engdahl, E.R., 2010. Sharpening the tomographic
image of the subducting slab below Sumatra, the Andaman Islands and Burma. Geophysical Journal International 182,
433-453.
Pindell, J., Maresch, W.V., Martens, U., Stanek, K., 2012. The Greater Antillean Arc: Early Cretaceous origin and proposed
relationship to Central American subduction mélanges: implications for models of Caribbean evolution. International
Geology Review 54, 131–143.
Pindell, J.L., Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and northern South America in the
mantle reference frame: an update. Geological Society, London, Special Publications 328, 1–55.
Piromallo, C., 2003. Pwave tomography of the mantle under the Alpine-Mediterranean area. Journal of Geophysical Research
108, 2065, doi: 10.1029/2002JB001757.
Piromallo, C., Morelli, A., 1997. Imaging the Mediterranean upper mantle by P-wave travel time tomography. Annali di
Geofisica 40, 963–979.
Piromallo, C., Morelli, A., 2003. P wave tomography of the mantle under the Alpine‐Mediterranean area. Journal of
Geophysical Research 108, 2065, doi: 10.1029/2002JB001757.
Pitman, W.C., 1978. Relationship between eustacy and stratigraphic sequences of passive margins. Geol. Soc. Am. Bull. 89,
1389–1403. doi:10.1130/0016-7606(1978)89<1389:RBEASS>2.0.CO;2
Platt, J.P., Allerton, S., Kirker, A., Mandeville, C., Mayfield, A., Platzman, E.S., Rimi, A., 2003. The ultimate arc: Differential
displacement, oroclinal bending, and vertical axis rotation in the External Betic-Rif arc. Tectonics 22, 1017–doi:10.1029–
2001TC001321.
Posamentier, H.W., Allen, G.P., 1993. Variability of the sequence stratigraphic model: effects of local basin factors. Sediment.
Geol., Basin Analysis and Dynamics of Sedimentary Basin Evolution 86, 91–109. doi:10.1016/0037-0738(93)90135-R
Posamentier, H.W., James, D.P., 2009. An Overview of Sequence-Stratigraphic Concepts: Uses and Abuses, in: Posamentier,
H.W., Summerhayes, C.P., Haq, B.U., Allen, G.P. (Eds.), Sequence Stratigraphy and Facies Associations. Blackwell
Publishing Ltd., Oxford, UK, pp. 1–18.
Posamentier, H.W., Vail, P.R., 1988. Sequence Stratigraphy: Sequences and Systems Tract Development. CSPG Spec. Publ.
15, 571–572.
Postma, G., 2014. Generic autogenic behaviour in fluvial systems, in: Martinius, A.W., Ravnås, R., Howell, J.A., Steel, R.J.,
Wonham, J.P. (Eds.), From Depositional Systems to Sedimentary Successions on the Norwegian Continental Margin.
John Wiley & Sons, Ltd, pp. 1–18.
Pownall, J.M., Hall, R., Armstrong, R.A., Forster, M.A., 2014. Earth's youngest known ultrahigh-temperature granulites
discovered on Seram, eastern Indonesia. Geology 42, 279–282.

200
Prokoph, A., Puetz, S.J., 2015. Period-Tripling and Fractal Features in Multi-Billion Year Geological Records. Math. Geosci.
47, 501–520. doi:10.1007/s11004-015-9593-y
Prokoph, A., Shields, G.A., Veizer, J., 2008. Compilation and time-series analysis of a marine carbonate [delta]18O,
[delta]13C, 87Sr/86Sr and [delta]34S database through Earth history. Earth-Sci. Rev. 87, 113–133. doi:doi: DOI:
10.1016/j.earscirev.2007.12.003
Pullen, A., Kapp, P., 2014. Mesozoic tectonic history and lithospheric structure of the Qiangtang terrane: Insights from the
Qiangtang metamorphic belt, central Tibet, in Geological Society of America Special Papers 507, 71–87.
Puspito, N.T., Shimazaki, K., 1995. Mantle structure and seismotectonics of the Sunda and Banda arcs. Tectonophysics 251,
215–228.
Puspito, N.T., Yamanaka, Y., Miyatake, T., Shimazaki, K., Hirahara, K., 1993. Three-dimensional P-wave velocity structure
beneath the Indonesian region. Tectonophysics 220, 175–192.
Qi, C., Zhao, D., Chen, Y., 2007. Search for deep slab segments under Alaska. Physics of The Earth and Planetary Interiors
165, 68–82.
Quarles van Ufford, A., Cloos, M., 2005. Cenozoic tectonics of New Guinea. AAPG Bulletin 89, 119–140.
Ramos, A., 2008. Journal of South American Earth Sciences. Journal of South American Earth Sciences 26, 235–251.
Rangin, C., Bijwaard, H., Pubellier, M., Spakman, W., 1999. Tomographic and geological constraints on subduction along the
eastern Sundaland continental margin (South-East Asia) . Bulletin de la Société Géologiques de France 170, 775–788.
Rawlinson, N., W. Spakman, 2016. On the use of sensitivity tests in seismic tomography, Geophysical Journal International
205, 1221–1243, doi: 10.1093/gji/ggw084
Raymo, M.E., Ruddiman, W.F., 1992. Tectonic forcing of late Cenozoic climate. Nature 359, 117–122. doi:10.1038/359117a0
Reagan, M.K., McClelland, W.C., Girard, G., Goff, K.R., Peate, D.W., Ohara, Y., Stern, R.J., 2013. The geology of the
southern Mariana fore-arc crust: Implications for the scale of Eocene volcanism in the western Pacific. Earth and
Planetary Science Letters 380, 41–51.
Ren, J., Tamaki, K., Li, S., Junxia, Z., 2002. Late Mesozoic and Cenozoic rifting and its dynamic setting in Eastern China and
adjacent areas. Tectonophysics 344, 175–205.
Ren, Y., Strutzmann, E., van der Hilst, R.D., Besse, J., 2007, Understanding seismic heterogeneities in the lower mantle
beneath the Americas from seismic tomography and plate tectonic history. Journal of Geophysical Research 112,
B01302–doi:10.1029–2005JB004154.
Ren, Y., Stuart, G.W., Houseman, G.A., Dando, B., Ionescu, C., Hegedus, E., Radovanovic, S., Shen, Y., Group, S.C.P.W.,
2012. Upper mantle structures beneath the Carpathian–Pannonian region Implications for the geodynamics of
continental collision. Earth and Planetary Science Letters 349-350, 139–152.
Replumaz, A., Kárason, H., van der Hilst, R.D., Besse, J., Tapponnier, P., 2004. 4-D evolution of SE Asia’s mantle from
geological reconstructions and seismic tomography. Earth and Planetary Science Letters 221, 103–115.
Replumaz, A., Negredo, A.M., Guillot, S., van der Beek, P., Villaseñor, A., 2010b. Crustal mass budget and recycling during
the India/Asia collision. Tectonophysics 492, 99–107.
Replumaz, A., Negredo, A.M., Guillot, S., Villasenor, A., 2010a. Multiple episodes of continental subduction during India/
Asia convergence: Insight from seismic tomography and tectonic reconstruction. Tectonophysics 483, 125–134.
Replumaz, A., Negredo, A.M., Villaseñor, A., Guillot, S., 2010c. Indian continental subduction and slab break-off during
Tertiary collision. Terra Nova 22, 290–296.
Resing, J.A., Lupton, J.E., Feely, R.A., Lilley, M.D., 2004. CO2 and 3He in hydrothermal plumes: implications for mid-ocean
ridge CO2 flux. Earth and Planetary Science Letters 226, 449–464.
Restrepo, J.J., Ordóñez-Carmona, O., Armstrong, R., Pimentel, M.M., 2011. Triassic metamorphism in the northern part of
the TahamI Terrane of the central cordillera of Colombia. Journal of South American Earth Sciences 32, 497–507.
Rezaei-Kahkhaei, M., Kananian, A., Esmaeily, D., Asiabanha, A., 2010. Geochemistry of the Zargoli granite: Implications for
development of the Sistan Suture Zone, southeastern Iran. Island Arc 19, 259–276.

201
Ricard, Y., Richards, M., Lithgow‐Bertelloni, C. and Le Stunff, Y., 1993. A geodynamic model of mantle density
heterogeneity. Journal of Geophysical Research: Solid Earth, 98(B12), pp.21895-21909.
Richards, M.A., Engebretson, D.C., 1992 Large-scale mantle convection and the history of subduction. Nature 355, 437 -
440. doi:10.1038/355437a0
Richter, F.M., Rowley, D.B., DePaolo, D.J., 1992. Sr isotope evolution of seawater: the role of tectonics. Earth Planet. Sci.
Lett. 109, 11–23. doi:doi: DOI: 10.1016/0012-821X(92)90070-C
Ricou, L.E., Burg, J.P., Godfriaux, I., Ivanov, Z., 1998. Rhodope and Vardar: the metamorphic and the olistostromic paired
belts related to the Cretaceous subduction under Europe. Geodinamica Acta 11, 285–309.
Rioux, M., Hacker, B., Mattinson, J., Kelemen, P., Blusztajn, J. and Gehrels, G., 2007. Magmatic development of an intra-
oceanic arc: High-precision U-Pb zircon and whole-rock isotopic analyses from the accreted Talkeetna arc, south-central
Alaska. Geological Society of America Bulletin, 119(9-10), 1168-1184.
Ritsema, J., Deuss, A., van Heijst, H.J., Woodhouse, J.H., 2011. S40RTS: a degree-40 shear-velocity model for the mantle
from new Rayleigh wave dispersion, teleseismic traveltime and normal-mode splitting function measurements.
Geophysical Journal International 184, 1223–1236.
Robertson, A.H.F., 2002. Overview of the genesis and emplacement of Mesozoic ophiolites in the Eastern Mediterranean
Tethyan region. Lithos 65, 1–67.
Rogers, R.D., Kárason, H., van der Hilst, R.D., 2002. Epeirogenic uplift above a detached slab in northern Central America.
Geology 30, 1031–1034.
Romans, B.W., Castelltort, S., Covault, J.A., Fildani, A., Walsh, J.P., 2016. Environmental signal propagation in sedimentary
systems across timescales. Earth-Sci. Rev., Source-to-Sink Systems: Sediment & Solute Transfer on the Earth Surface
153, 7–29. doi:10.1016/j.earscirev.2015.07.012
Ronov, A.B., 1994. Phanerozoic transgressions and regressions on the continents; a quantitative approach based on areas
flooded by the sea and areas of marine and continental deposition. Am. J. Sci. 294, 777–801. doi:10.2475/ajs.294.7.777
Ronov, A.B., Khain, V.E.., Seslavinsky, K.B., 1984. Atlas of Lithological-Paleogeographical Maps of the World: Precambrian
and Paleozoic of Continents and Oceans. Academy of Sciences Press, Leningrad.
Ronov, A.B., Khain, V.E., Balukhovsky, A.N., 1989. Atlas of Lithological-Paleogeographical Maps of the World: Mesozoic
and Cenozoic of Continents and Oceans. Editorial Publishing Group, Moscow.
Rosenbaum, G., Avigad, D., Sánchez-Gómez, M., 2002a. Coaxial flattening at deep levels of orogenic belts: evidence from
blueschists and eclogites on Syros and Sifnos (Cyclades, Greece). Journal of Structural Geology 24, 1451–1462.
Rosenbaum, G., Gasparon, M., Lucente, F.P., Peccerillo, A., Miller, M.S., 2008. Kinematics of slab tear faults during
subduction segmentation and implications for Italian magmatism. Tectonics 27, TC2008. doi:10.1029–2007TC002143.
Rosenbaum, G., Lister, G.S., 2004. Formation of arcuate orogenic belts in the western Mediterranean region. Geological
Society of America Special Paper 383, 41–56.
Rosenbaum, G., Lister, G.S., Duboz, C., 2002b. Reconstruction of the tectonic evolution of the western Mediterranean since
the Oligocene. Journal of the virtual explorer 8, 107–130.
Rost, S., Garnero, E.J., Williams, Q., 2008. Seismic array detection of subducted oceanic crust in the lower mantle. Journal of
Geophysical Research 113, B06303–11. doi: 10.1029/2007JB005263.
Roth, J.B., Fouch, M.J., James, D.E., Carlson, R.W., 2008. Three-dimensional seismic velocity structure of the northwestern
United States. Geophysical Research Letters 35, L15304–6, doi: 10.1029/2008GL034669.
Rowley, D.B., 2002. Rate of plate creation and destruction: 180 Ma to present. Geol. Soc. Am. Bull. 114, 927–933.
doi:10.1130/0016-7606(2002)114<0927:ROPCAD>2.0.CO;2
Rudolph, M.L., Lekić, V. and Lithgow-Bertelloni, C., 2015. Viscosity jump in Earth’s mid-mantle. Science, 350(6266), 1349-
1352.
Ruiz-Martinez, V.C., Torsvik, T.H., van Hinsbergen, D.J.J., Gaina, C., 2012. Earth at 200 Ma: Global palaeogeography
refined from CAMP palaeomagnetic data. Earth and Planetary Science Letters 331-332, 67–79.

202
Sagong, H., Kwon, S.T., Ree, J.H., 2005. Mesozoic episodic magmatism in South Korea and its tectonic implication. Tectonics
24, TC5002. doi:10.1029–2004TC001720.
Sano, Y., Williams, S.N., 1996. Fluxes of mantle and subducted carbon along convergent plate boundaries. Geophysical
Research Letters 23, 2749–2752.
Sayıt, K., Göncüoglu, M.C., 2013. Geodynamic evolution of the Karakaya Mélange Complex, Turkey: A review of geological
and petrological constraints. Journal of Geodynamics 65, 56–65.
Schaefer, B.F., Turner, S., Parkinson, I., Rogers, N., Hawkesworth, C., 2002. Evidence for recycled Archaean oceanic mantle
lithosphere in the Azores plume. Nature 420, 304–307.
Schellart, W.P., Freeman, J., Stegman, D.R., Moresi, L., May, D., 2007. Evolution and diversity of subduction zones controlled
by slab width. Nature 446, 308–311. doi:10.1038/nature05615
Schellart, W.P., Kennett, B.L.N., Spakman, W., Amaru, M.L., 2009. Plate reconstructions and tomography reveal a fossil
lower mantle slab below the Tasman Sea 278, 143–151.
Schellart, W.P., Lister, G.S., Toy, V.G., 2006. A Late Cretaceous and Cenozoic reconstruction of the Southwest Pacific region:
Tectonics controlled by subduction and slab rollback processes. Earth-Science Reviews 76, 191–233.
Schellart, W.P., Spakman, W., 2012. Mantle constraints on the plate tectonic evolution of the Tonga–Kermadec–Hikurangi
subduction zone and the South Fiji Basin region. Australian Journal of Earth Sciences 59, 933–952.
Schellart, W.P., Spakman, W., 2015. Australian plate motion and topography linked to fossil New Guinea slab below Lake
Eyre. Earth and Planetary Science Letters 421, 107-116.
Scherreiks, R., Meléndez, G., Boudagher-Fadel, M., Fermeli, G., Bosence, D., 2014. Stratigraphy and tectonics of a time-
transgressive ophiolite obduction onto the eastern margin of the Pelagonian platform from Late Bathonian until
Valanginian time, exemplified in northern Evvoia, Greece. International Journal of Earth Sciences 103, 2191–2216.
Schildgen, T.F., Cosentino, D., Caruso, A., Buchwaldt, R., Yıldırım, C., Bowring, S.A., Rojay, B., Echtler, H., Strecker, M.R.,
2012. Surface expression of eastern Mediterranean slab dynamics: Neogene topographic and structural evolution of the
southwest margin of the Central Anatolian Plateau, Turkey. Tectonics 31. doi: 10.1029/2011TC003021.
Schlager, W., 1993. Accommodation and supply—a dual control on stratigraphic sequences. Sediment. Geol., Basin Analysis
and Dynamics of Sedimentary Basin Evolution 86, 111–136. doi:10.1016/0037-0738(93)90136-S
Schmandt, B., Humphreys, E., 2011. Seismically imaged relict slab from the 55 Ma Siletzia accretion to the northwest United
States. Geology 39, 175–178.
Schmid, S.M., Bernoulli, D., Fügenschuh, B., Maţenco, L., Schefer, S., Schuster, R., Tischler, M., Ustaszewski, K., 2008. The
Alpine-Carpathian-Dinaridic orogenic system: correlation and evolution of tectonic units. Swiss Journal of Geosciences
101, 139–183.
Schmid, S.M., Fugenschuh, B., Kissling, E., Schuster, R., 2004. Tectonic map and overall architecture of the Alpine orogen.
Eclogae Geologicae Helvetiae 97, 93–117.
Schneider, F.M., Yuan, X., Schurr, B., Mechie, J., Sippl, C., Haberland, C., Minaev, V., Oimahmadov, I., Gadoev, M.,
Radjabov, N., Abdybachaev, U., Orunbaev, S., Negmatullaev, S., 2013. Seismic imaging of subducting continental lower
crust beneath the Pamir. Earth and Planetary Science Letters 375, 101–112.
Scholl, D.W., Vallier, T.L., Stevenson, A.J., 1989. Geologic evolution and petroleum geology of the Aleutian Ridge, in Scholl,
D.W., Grantz, A., Vedder, J.G. eds., Houston, Texas, 122–155.
Schumacher, L., Thomas, C., 2016. Detecting lower-mantle slabs beneath Asia and the Aleutians. Geophysical Journal
International 205, 1512–1524.
Scotese, C.R., 1998. Animation: Continental Drift, 0-750 million years (paleogeography.mov, Quicktime format),
PALEOMAP Project, www.scotese.com, Evanston, IL. Researchgate.
Scotese, C.R., 2014a. The PALEOMAP Project PaleoAtlas for ArcGIS, version 1, Volume 1, Cenozoic Paleogeographic and
Plate Tectonic Reconstructions, PALEOMAP Project, Evanston, IL. DOI:10.13140/RG.2.1.2535.7041.
Scotese, C.R., 2014b. The PALEOMAP Project PaleoAtlas for ArcGIS, version 1, Volume 2, Cretaceous Paleogeographic and
Plate Tectonic Reconstructions, PALEOMAP Project, Evanston, IL. DOI:10.13140/RG.2.1.2535.7041.

203
Scotese, C.R., 2014c. The PALEOMAP Project PaleoAtlas for ArcGIS, version 1, Volume 3, Triassic and Jurassic
Paleogeographic and Plate Tectonic Reconstructions, PALEOMAP Project, Evanston, IL. DOI:10.13140/
RG.2.1.2535.7041.
Scotese, C.R., 2014d. The PALEOMAP Project PaleoAtlas for ArcGIS, version 1, Volume 4, Late Paleozoic Paleogeographic
and Plate Tectonic Reconstructions, PALEOMAP Project, Evanston, IL. DOI:10.13140/RG.2.1.2535.7041.
Scotese, C.R., 2014e. The PALEOMAP Project PaleoAtlas for ArcGIS, version 1, Volume 5, Early Paleozoic Paleogeographic
and Plate Tectonic Reconstructions, PALEOMAP Project, Evanston, IL. DOI:10.13140/RG.2.1.2535.7041.
Scotese, C.R., 2014f. The PALEOMAP Project PaleoAtlas for ArcGIS, version 1, Volume 6, Late Precambrian
Paleogeographic and Plate Tectonic Reconstructions, PALEOMAP Project, Evanston, IL. DOI:10.13140/
RG.2.1.2535.7041.
Scotese, C.R., 2016. PALEOMAP PaleoAtlas for GPlates and the PaleoData Plotter Program, PALEOMAP Project. http://
www.earthbyte.org/paleomap-paleoatlas-for-gplates/.
Scotese, C.R., Golonka, J., 1992. Paleogeographic Atlas, PALEOMAP Progress Report 20-0692, Department of Geology,
University of Texas at Arlington, Texas, 34 p. Researchgate.
Sdrolias, M., Müller, R.D., Gaina, C., 2004. Tectonic evolution of the southwest Pacific using constraints from backarc basins.
Geological Society of Australia Special Publication 22, 343–359.
Searle, M.P., Corfield, R.I., Stephenson, B., McCarron, J., 1997. Structure of the north Indian continental margin in the
Ladakh-Zanskar Himalayas: Implications for the timing of obduction of the Spontang ophiolite, India-Asia collision
and deformation events in the Himalaya. Geological Magazine 134, 297–316.
Searle, M.P., Cox, J.C., 2009. Tectonic setting, origin and obduction of the Oman Ophiolite. Geological Society of America
Bulletin 111, 104-122.
Sengör, A.M.C., Özeren, S., Genç, T., Zor, E., 2003. East Anatolian high plateau as a mantle-supported, north-south
shortened domal structure. Geophysical Research Letters 30, 8045–doi: 10.1029–2003GL017858.
Sengör, A.M.C., Yilmaz, Y., 1981. Tethyan evolution of Turkey: A plate tectonic approach. Tectonophysics 75, 181–241.
Seno, T., Maruyama, S., 1984. Paleogeographic reconstruction and origin of the Philippine Sea. Tectonophysics 102, 53–84.
Seton, M., Flament, N., Whittaker, J., Müller, R.D., Gurnis, M., and Bower, D.J., 2015. Ridge subduction sparked
reorganization of the Pacific plate-mantle system 60–50 million years ago: Geophysical Research Letters 42, 1732–1740.
Seton, M., Müller, R.D., Zahirovic, S., Gaina, C., Torsvik, T.H., Shephard, G., Talsma, A., Gurnis, M., Turner, M., Maus,
S., Chandler, M., 2012. Global continental and ocean basin reconstructions since 200 Ma. Earth-Science Reviews 113,
212–270.
Sewall, J.O., Wal, R.S.W.V. de, Zwan, K.V.D., Oosterhout, C.V., Dijkstra, H.A., Scotese, C.R., 2007. Climate model
boundary conditions for four Cretaceous time slices. Clim. Past 3, 647–657.
Shahnas, M.H, Yuen, D.A, Pysklywec, R.N., 2017. Mid-mantle heterogeneities and iron spin transition in the lower mantle:
Implications for mid-mantle slab stagnation, Earth and Planetary Science Letters 458, 293-304
Shapiro, M.N., Solov’ev, A.V., 2009. Formation of the Olyutorsky–Kamchatka foldbelt: a kinematic model. Russian Geology
and Geophysics 50, 668–681.
Shephard, G.E., Flament, N., Williams, S., Seton, M., Gurnis, M., Müller, R.D., 2014. Circum-Arctic mantle structure and
long-wavelength topography since the Jurassic. Journal of Geophysical Research 119, doi: 10.1002/2014JB011078.
Shephard, G.E., Müller, R.D., Seton, M., 2013. The tectonic evolution of the Arctic since Pangea breakup: Integrating
constraints from surface geology and geophysics with mantle structure. Earth-Science Reviews 124, 148–183.
Shephard, G.E., Trønnes, R.G., Spakman, W., Panet, I., Gaina, C., 2016. Evidence for slab material under Greenland and
links to Cretaceous High Arctic magmatism. Geophysical Research Letters 43, 3717-3726.
Shi, G.R., 2006. The marine Permian of East and Northeast Asia: an overview of biostratigraphy, paleobiogeography and
paleogeographical implications. J. Asian Earth Sci. 26, 175–206. doi:doi: DOI: 10.1016/j.jseaes.2005.11.004
Sibuet, J.-C., Hsu, S.-K., 2004. How was Taiwan created? Tectonophysics 379, 159–181.

204
Sibuet, J.C., Srivastava, S.P., Spakman, W., 2004. Pyrenean orogeny and plate kinematics. Journal of Geophysical Research
109, B08104–doi: 10.1029–2003JB002514.
Sigloch, K., 2011. Mantle provinces under North America from multifrequency P wave tomography. Geochemistry,
Geophysics, Geosystems 12, Q02W08–doi:10.1029–2010GC003421.
Sigloch, K., McQuarrie, N., Nolet, G., 2008. Two-stage subduction history under North America inferred from multiple-
frequency tomography. Nature Geoscience 1, 458–462.
Sigloch, K., Mihalynuk, M.G., 2013. Intra-oceanic subduction shaped the assembly of Cordilleran North America. Nature
496, 50–56.
Silver, P.G., Behn, M.D., 2008. Intermittent Plate Tectonics? Science 319, 85–88.
Simmons, N.A., Myers, S.C., Johannesson, G., Matzel, E., 2012. LLNL-G3Dv3: Global P wave tomography model
for improved regional and teleseismic travel time prediction. Journal of Geophysical Research 117, B10302, doi:
10.1029/2012JB009525.
Simmons, N.A., Myers, S.C., Johannesson, G., Matzel, E., Grand, S.P., 2015. Evidence for long-lived subduction
of an ancient tectonic plate beneath the southern Indian Ocean. Geophys. Res. Lett., 42, 9270–9278,
doi:10.1002/2015GL066237
Sippl, C., Schurr, B., Tympel, J., Angiboust, S., Mechie, J., Yuan, X., Schneider, F.M., Sobolev, S.V., Ratschbacher, L.,
Haberland, C., TIPAGE-Team, 2013. Deep burial of Asian continental crust beneath the Pamir imaged with local
earthquake tomography. Earth and Planetary Science Letters 384, 165–177.
Skolbeltsyn, G., Mellors, R., Gok, R., Turkelli, N., 2014. Upper Mantle S wave Velocity Structure of the East Anatolian‐
Caucasus Region. Tectonics 33, 207-221.
Sluijs, A., Brinkhuis, H., Crouch, E.M., John, C.M., Handley, L., Munsterman, D., Bohaty, S., Zachos, J.C., Reichart, G.J.,
Schouten, S., Pancost, R.D., Sinninghe Damsté, J.S., Welters, N.L.D., Lotter, A.D., Dickens, G.R., 2008. Eustatic
variations during the Paleocene-Eocene greenhouse world. Paleoceanography 23, PA4216.
Smith, A.G., Smith, D.G., Funnell, B.M., 2004. Atlas of Mesozoic and Cenozoic Coastlines. Cambridge University Press,
Cambridge.
Snedden, J.W., Liu, C., 2010. A Compilation of Phanerozoic Sea-Level Change, Coastal Onlaps and Recommended
Sequence Designations. Search Discov. Artic.
Snedden, J.W., Liu, C., 2011. Recommendations for a uniform chronostratigraphic designation system for Phanerozoic
depositional sequences. AAPG Bull. 95, 1095–1122. doi:10.1306/01031110138
Sobel, E.R., Chen, J., Schoenbohm, L.M., Thiede, R., Stockli, D.F., Sudo, M., Strecker, M.R., 2013. Oceanic-style subduction
controls late Cenozoic deformation of the Northern Pamir orogen. Earth and Planetary Science Letters 363, 204–218.
Sokolov, S.D., 1992. Accretionary Tectonics of the Koryak–Chukchi Segment of the Pacific Belt [in Russian]. Nauka,
Moscow.
Sokolov, S.D., 2010. Tectonics of Northeast Asia: An overview. Geotectonics 44, 493–509.
Sokolov, S.D., Bondarenko, G.Y., Layer, W., Kravchenko-Berezhnoy, I.R., 2009. South Anyui suture: tectono-stratigraphy,
deformations, and principal tectonic events. Stephan Mueller Special Publication Series 4, 201–221.
Šoštarić, S.B., Palinkaš, A.L., Neubauer, F., Cvetkovic, V., Bernroider, M., Genser, J., 2014. The origin and age of the
metamorphic sole from the Rogozna Mts., Western Vardar Belt: New evidence for the one-ocean model for the Balkan
ophiolites. Lithos 192-195, 39–55.
Spakman, W. and Nolet, G., 1988. Imaging algorithms, accuracy and resolution in delay time tomography. In Mathematical
geophysics (155-187). Springer Netherlands.
Spakman, W., 1986a. Subduction beneath Eurasia in connection with the Mesozoic Tethys. Geologie en Mijnbouw 65,
145–153.
Spakman, W., 1986b. The upper mantle structure in the Central European-Mediterranean region, in Freeman, R., Mueller, S.,
Giese, eds., Strassbourg, 215–222.

205
Spakman, W., Hall, R., 2010. Surface deformation and slab–mantle interaction during Banda arc subduction rollback. Nature
Geoscience 3, 562–566.
Spakman, W., Stein, S., van der Hilst, R.D., Wortel, M.J.R., 1989. Resolution experiments for NW Pacific subduction zone
tomography. Geophysical Research Letters 16, 1097–1100.
Spakman, W., van der Lee, S., van der Hilst, R., 1993. Travel-time tomography of the European-Mediterranean mantle down
to 1400 km. Physics of The Earth and Planetary Interiors 79, 3–74.
Spakman, W., Wortel, M., Vlaar, N.J., 1988. The Hellenic subduction zone: a tomographic image and its geodynamic
implications. Geophysical Research Letters 15, 60–63.
Spakman, W., Wortel, M.J.R., 2004. A tomographic view on western Mediterranean geodynamics. in: Cavazza, W., Roure, F.,
Spakman, W., Stampfli, G.M., Ziegler, P. The TRANSMED Atlas, The Mediterranean Region from Crust to Mantle,
31-52.
Spasojevic, S., Gurnis, M., 2012. Sea level and vertical motion of continents from dynamic earth models since the Late
Cretaceous. AAPG Bull. 96, 2037–2064. doi:10.1306/03261211121
Spooner, E.T.C., 1976. The strontium isotopic composition of seawater, and seawater-oceanic crust interaction. Earth Planet.
Sci. Lett. 31, 167–174. doi:10.1016/0012-821X(76)90108-4
Stampfli GM, 2003. Stratigraphic and Structural Evolution on the Late Carboniferous to Triassic Continental and Marine
Successions in Tuscany (Italy). Regional Reports and General Correlation (Bollettino della Società Geologica Italiana,
Roma) Vol speciale 2, 1-23
Stampfli, G.M., Borel, G.D., 2002. A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate
boundaries and restored synthetic oceanic isochrons. Earth and Planetary Science Letters 196, 17–33.
Stampfli, G.M., Borel, G.D., 2004. The TRANSMED Transects in Space and Time: Constraints on the Paleotectonic
Evolution of the Mediterranean Domain. In: The TRANSMED Atlas. The Mediterranean Region from Crust to
Mantle, Springer Berlin Heidelberg, Berlin, Heidelberg, 53–80.
Stanek, K.P., Maresch, W.V., Pindell, J.L., 2009. The geotectonic story of the northwestern branch of the Caribbean Arc:
implications from structural and geochronological data of Cuba (K. P. Stanek, W. V. Maresch, & J. Pindell, Eds.).
Geological Society, London, Special Publications 328, 361–398.
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. doi:10.1038/359123a0
Steinberger, B. and Calderwood, A.R., 2006. Models of large-scale viscous flow in the Earth's mantle with constraints
from mineral physics and surface observations. Geophysical Journal International, 167(3), 1461-1481.doi:10.1038/
nature06824
Steinberger, B., 2000. Slabs in the lower mantle - results of dynamic modelling compared with tomographic images and the
geoid. Physics of the Earth and Planetary Interiors 118, 241–257.
Steinberger, B., Torsvik, T.H., 2008. Absolute plate motions and true polar wander in the absence of hotspot tracks. Nature
452, 620-623.
Steinberger, B., Torsvik, T.H., 2010. Toward an explanation for the present and past locations of the poles. Geochemistry
Geophysics Geosystems, v. 11, n/a–n/a. doi:10.1029/2009GC002889
Stern, R.J., 2002. Subduction zones. Reviews of Geophysics, v. 40, doi: 10.1029–2001RJ000108. doi:10.1029/2001RG000108
Stern, R.J., Bloomer, S.H., 1992. Subduction zone infancy: Examples from the Eocene Izu-Bonin-Mariana and Jurassic
California arcs. Geological Society of America Bulletin 104, 1621–1636.
Stern, R.J., Reagan, M., Ishizuka, O., Ohara, Y., Whattam, S.A., 2012. To understand subduction initiation, study forearc
crust: To understand forearc crust, study ophiolites. Lithosphere 4, 469–483.
Stocchi, P., Escutia, C., Houben, A.J.P., Vermeersen, B.L.A., Bijl, P.K., Brinkhuis, H., DeConto, R.M., Galeotti, S., Passchier,
S., Pollart, D., Scientists, E.3., 2013. Relative sea-level rise around East Antarctica during Oligocene glaciation. Nature
Geoscience 6, 380–384.

206
Stone, D.B., Layer, W., Raikevich, M.I., 2009. Age and paleomagnetism of the Okhotsk-Chukotka Volcanic Belt (OCVB)
near Lake El'gygytgyn, Chukotka, Russia. Stephan Mueller Special Publication Series 4, 243–260.
Stone, D.B., Minyuk, P., Kolosev, E., 2003. New paleomagnetic paleolatitudes for the Omulevka terrane of northeast Russia:
a comparison with the Omolon terrane and the eastern Siberian platform. Tectonophysics 377, 55–82. doi:doi: DOI:
10.1016/j.tecto.2003.08.025
Stoykov, S., Peytcheva, I., Quadt, von, A., Moritz, R., Frank, M., Fontignie, D., 2004. Timing and magma evolution of the
Chelopech volcanic complex (Bulgaria). Swiss Bulletin of Mineralogy and Petrology 84, 101–117.
Sugioka, H., Suetsugu, D., Obayashi, M., Fukao, Y., Gao, Y., 2010, Fast P- and S-wave velocities associated with the “cold”
stagnant slab beneath the northern Philippine Sea. Physics of The Earth and Planetary Interiors 179, 1–6.
Suzuki, Y., Kawai, K., Geller, R.J., Borgeaud, A.F.E., Konishi, K., 2016. Waveform inversion for 3-D S-velocity structure of
D" beneath the Northern Pacific: possible evidence for a remnant slab and a passive plume. Earth, Planets and Space, 68,
198, doi: 10.1186/s40623-016-0576-0.
Svensen, H., Planke, S., Malthe-Sørenssen, A., Jamtveit, B., Myklebust, R., Rasmussen Eidem, T., Rey, S.S., 2007. Release of
methane from a volcanic basin as a mechanism for initial Eocene global warming. Nature 429, 542–545.
Swarbrick, R.E., Naylor, M.A., 1980. The Kathikas melange, SW Cyprus: late Cretaceous submarine debris flows.
Sedimentology 27, 63–78.
Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J., Rivera, C., 2000. Geodynamics of
the northern Andes: Subductions and intracontinental deformation (Colombia). Tectonics 19, 787–813.
Talavera-Mendoza, O., Ruiz, J., Gehrels, G.E., ValenciaA., Centeno-Garcia, E., 2007. Detrital zircon U/Pb geochronology of
southern Guerrero and western Mixteca arc successions (southern Mexico): New insights for the tectonic evolution of
southwestern North America during the late Mesozoic. Geological Society of America Bulletin 119, 1052–1065.
Tankard, A., Welsink, H., Aukes, P., Newton, R., Stettler, E., 2009. Tectonic evolution of the Cape and Karoo basins of South
Africa. Marine and Petroleum Geology 26, 1379–1412.
Tate, G.W., McQuarrie, N., Tiranda, H., van Hinsbergen, D.J.J., Harris, R., Zachariasse, W.J., Fellin, M.G., Reiners, W.,
Willett, S.D., 2017. First-order orogenic cylindricity amid locally high variability in uplift, exhumation and geometry:
insights on orogenic controls from West Timor. Gondwana Research, under review
Tate, G.W., McQuarrie, N., van Hinsbergen, D.J., Bakker, R.R., Harris, R.A., Jiang, H., 2015, Australia going down under:
Quantifying continental subduction during arc-continent accretion in Timor-Leste. Geosphere 11, 1860-1883.
Tate, G.W., McQuarrie, N., van Hinsbergen, D.J.J., Bakker, R.R., Harris, R., Willett, S., Reiners, W., Fellin, M.G., Ganerød,
M., Zachariasse, W.J., 2014. Resolving spatial heterogeneities in exhumation and surface uplift in Timor‐Leste:
Constraints on deformation processes in young orogens. Tectonics 33, 1089–1112.
Tekin, U.K., Göncüoğlu, M.C., 2007. Discovery of the oldest (Upper Ladinian to Middle Carnian) radiolarian assemblages
from the Bornova Flysch zone in Western Turkey: implications for the evolution of the Neotethyan Izmir-Ankara ocean.
Ofioliti 32, 131–150.
Tekin, U.K., Göncüoglu, M.C., Turhan, N., 2002. First evidence of Late Carnian radiolarians from the Izmir–Ankara suture
complex, central Sakarya, Turkey: implications for the opening age of the Izmir–Ankara branch of Neo-Tethys. Geobios
35, 127–135.
Thomas, C., Garnero, E.,J., Lay, T., 2004. High-resolution imaging of lowermost mantle structure under the Cocos plate.
Journal of Geophysical Research 109. B08307–doi:10.1029–2004JB003013.
Tirrul, R., Bell, I.R., Griffis, R.J., Camp, E., 1983. The Sistan suture zone of eastern Iran. Geological Society of America
Bulletin 94, 134–150.
Tomurtogoo, O., Windley, B.F., Kröner, A., Badarch, G., Liu, D.Y., 2005. Zircon age and occurrence of the Adaatsag ophiolite
and Muron shear zone, central Mongolia: constraints on the evolution of the Mongol-Okhotsk ocean, suture and orogen.
Journal of the Geological Society 162, 125–134.
Torsvik, T.H., Burke, K., Steinberger, B., Webb, S.J., Ashwal, L.D., 2010. Diamonds sampled by plumes from the core–mantle
boundary. Nature 466, 352–355.

207
Torsvik, T.H., Cocks, L.R.M., 2009. The Lower Palaeozoic palaeogeographical evolution of the northeastern and eastern peri-
Gondwanan margin from Turkey to New Zealand. Geological Society, London, Special Publications 325, 3–21.
Torsvik, T.H., Doubrovine, P.V., Steinberger, B., Gaina, C., Spakman, W., and Domeier, M., 2017. Pacific plate motion change
caused the Hawaiian-Emperor Bend : Nature Communications in press.
Torsvik, T.H., Müller, R.D., Van der Voo, R., Steinberger, B., Gaina, C., 2008. Global plate motion frames: Toward a unified
model. Reviews of Geophysics 46, RG3004–doi:10.1029–2007RG000227.
Torsvik, T.H., Steinberger, B., Cocks, L.R.M., Burke, K., 2008. Longitude: Linking Earth's ancient surface to its deep interior.
Earth and Planetary Science Letters 276, 273–282.
Torsvik, T.H., Steinberger, B., Gurnis, M., Gaina, C., 2010. Plate tectonics and net lithosphere rotation over the past 150 My.
Earth Planet. Sci. Lett. 291, 106–112. doi:10.1016/j.epsl.2009.12.055
Torsvik, T.H., Van der Voo, R., Doubrovine, V., Burke, K., Steinberger, B., Ashwal, L.D., Tronnes, R.G., Webb, S.J., Bull,
A.L., 2014. Deep mantle structure as a reference frame for movements in and on the Earth. Proceedings of the National
Academy of Sciences of the United States of America 111, 8735–8740.
Torsvik, T.H., Van der Voo, R., Preeden, U., MacNiocaill, C., Steinberger, B., Doubrovine, V., van Hinsbergen, D.J.J.,
Domeier, M., Gaina, C., Tohver, E., Meert, J.G., McCausland, J.A., Cocks, L.R.M., 2012. Phanerozoic polar wander,
palaeogeography and dynamics. Earth-Science Reviews 114, 325–368.
Tremblay, A., Meshi, A., Deschamps, T., Goulet, F., Goulet, N., 2015. The Vardar Zone as a suture for the Mirdita
ophiolites, Albania : constraints from the structural analysis of the Korabi-Pelagonia Zone. Tectonics 34, doi:
10.1002/2014TC003807.
Trop, J.M., Ridgway, K.D., 2007. Mesozoic and Cenozoic tectonic growth of southern Alaska: A sedimentary basin
perspective. Geological Society of America Special Paper 432, 55–94.
Trop, J.M., Szuch, S.A., Rioux, M., Blodgett, R.B., 2005. Sedimentology and provenance of the Upper Jurassic Naknek
Formation, Talkeetna Mountains, Alaska; bearings on the accretionary tectonic history of the Wrangellia composite
terrane. Geological Society of America Bulletin 117, 570–588.
Ueda, H., Miyashita, S., 2005. Tectonic accretion of a subducted intraoceanic remnant arc in Cretaceous Hokkaido, Japan, and
implications for evolution of the Pacific northwest. Isl. Arc 14, 582–598. doi:10.1111/j.1440-1738.2005.00486.x
Umbgrove, J.H.F., 1940. Periodicity in terrestrial processes. Am. J. Sci. 238, 573–576. doi:10.2475/ajs.238.8.573
Umbgrove, J.H.F., 1947. The pulse of the earth. Martinus Nijhoff, The Hague, Netherlands.
Uno, K., Furukawa, K., Hada, S., 2011. Margin-parallel translation in the western Pacific: Paleomagnetic evidence from an
allochthonous terrane in Japan. Earth and Planetary Science Letters 303, 153–161. doi:doi: 10.1016/j.epsl.2011.01.002
Ustaszewski, K., Schmid, S.M., Fügenschuh, B., Tischler, M., Kissling, E., Spakman, W., 2008. A map-view restoration of the
Alpine-Carpathian-Dinaridic system for the Early Miocene. Swiss Journal of Geosciences 101, 273–294.
Ustaszewski, K., Wu, Y.-M., Suppe, J., Huang, H.-H., Chang, C.-H., Carena, S., 2012. Crust–mantle boundaries in
the Taiwan–Luzon arc-continent collision system determined from local earthquake tomography and 1D models:
Implications for the mode of subduction polarity reversal. Tectonophysics 578, 31–49.
Vail, P.R., Mitchum, R.M., Thompson, S.I., 1977. Seismic Stratigraphy and Global Changes of Sea Level: Part 4. Global
Cycles of Relative Changes of Sea Level 165, 83–97.
van Benthem, S., Govers, R., Spakman, W., Wortel, M.J.R., 2013. Tectonic evolution and mantle structure of the Caribbean
118, 3019–3036.
van de Lagemaat, S.H.A., van Hinsbergen, D.J.J., Boschman, L.M., Kamp, J.J., 2017. Southwest Pacific absolute plate
kinematic reconstruction reveals major mid – Late Cenozoic Tonga-Kermadec slab-dragging, Earth-Science Reviews
submitted.
Van den Berg van Saparoea, A.P., Postma, G., 2008. Control of climate change on the yield of river systems, in: Recent
Advances in Models of Siliciclastic Shallow-Marine Stratigraphy, SEPM (Society for Sedimentary Geology) Special
Publication,. pp. 15–33.

208
van der Hilst, R. and Seno, T., 1993. Effects of relative plate motion on the deep structure and penetration depth of slabs
below the Izu-Bonin and Mariana island arcs. Earth and Planetary Science Letters, 120(3-4), 395-407.
Van der Hilst, R., Engdahl, R., Spakman, W., Nolet, G., 1991. Tomographic imaging of subducted lithosphere below
northwest Pacific island arcs. Nature 353, 37–43.
van der Hilst, R.D. and Mann, P., 1994. Tectonic implications of tomographic images of subducted lithosphere beneath
northwestern South America. Geology 22, 451–454.
van der Hilst, R.D. and Spakman, W., 1989. Importance of reference model in linearized tomography and images of
subduction below the Caribbean plate. Geophysical Research Letters 16, 1093–1096.
van der Hilst, R.D., 1995. Complex morphology of subducted lithosphere in the mantle beneath the Tonga trench. Nature
374, 154–157.
van der Hilst, R.D., de Hoop, M.V., Wang, P., Shim, S.H., Ma, P., Tenorio, L., 2007. Seismostratigraphy and Thermal
Structure of Earth's Core-Mantle Boundary Region. Science 315, 1813–1817.
van der Hilst, R.D., Widiyantoro, S., Engdahl, E.R., 1997. Evidence for deep mantle circulation from global tomography.
Nature 386, 578-584. doi:10.1038/386578a0
van der Lee, S., Nolet, G., 1997. Seismic image of the subducted trailing fragments of the Farallon plate. Nature 386, 266–
269.
van der Meer, D.G., Spakman, W., van Hinsbergen, D.J.J., Amaru, M.L., Torsvik, T.H., 2010. Towards absolute plate motions
constrained by lower-mantle slab remnants. Nature Geoscience 3, 36–40.
van der Meer, D.G., Torsvik, T.H., Spakman, W., van Hinsbergen, D.J.J., Amaru, M.L., 2012. Intra-Panthalassa Ocean
subduction zones revealed by fossil arcs and mantle structure. Nature Geoscience 5, 215–219.
van der Meer, D.G., van den Berg van Saparoea, A.P.H., van Hinsbergen, D.J.J., van de Weg, R.M.B., Goddéris, Y., Le Hir,
G., Donnadieu, Y., 2017. Reconstructing first-order changes in sea level during the Phanerozoic and Neoproterozoic
using strontium isotopes. Gondwana Research 44, 22-34.
van der Meer, D.G., van Hinsbergen, D.J.J., Spakman, W., submitted, The Atlas of the Underworld: a catalogue of slab
remnants in the mantle imaged by seismic tomography and their geological interpretation, Tectonophysics
van der Meer, D.G., Zeebe, R.E., van Hinsbergen, D.J.J., Sluijs, A., Spakman, W., Torsvik, T.H., 2014. Plate tectonic controls
on atmospheric CO2 levels since the Triassic. Proceedings of the National Academy of Sciences 111, 4380–4385.
Van der Voo, R., Spakman, W., Bijwaard, H., 1999a. Mesozoic subducted slabs under Siberia. Nature 397, 246–249.
Van der Voo, R., Spakman, W., Bijwaard, H., 1999b. Tethyan subducted slabs under India. Earth and Planetary Science
Letters 171, 7–20.
Van der Voo, R., van Hinsbergen, D.J.J., Domeier, M., Spakman, W., Torsvik, T.H., 2015. Latest Jurassic-earliest Cretaceous
closure of the Mongol-Okhotsk Ocean: A paleomagnetic and seismological-tomographic analysis, Geological Society of
America Special Paper 513, 589-606.
van Hinsbergen, D.J., Lippert, C., Dupont-Nivet, G., McQuarrie, N., Doubrovine, V., Spakman, W., Torsvik, T.H., 2012.
Greater India Basin hypothesis and a two-stage Cenozoic collision between India and Asia. Proceedings of the National
Academy of Sciences of the United States of America 109, 7659–7664.
van Hinsbergen, D.J., Maffione, M., Plunder, A., Kaymakci, N., Ganerød, M., Hendriks, B.W.H., Corfu, F., Gürer, D., de
Gelder, G.I.N.O., Peters, K., McPhee, J., Brouwer, F.M., Advokaat, E.L., Vissers, R.L.M., 2016a. Tectonic evolution and
paleogeography of the Kırşehir Block and the Central Anatolian Ophiolites, Turkey. Tectonics 35, 983–1014.
van Hinsbergen, D.J.J., 2010. A key extensional metamorphic complex reviewed and restored: The Menderes Massif of
western Turkey. Earth-Science Reviews 102, 60–76.
van Hinsbergen, D.J.J., Abels, H.A., Bosch, W., Boekhout, F., Kitchka, A., Hamers, M.F., van der Meer, D.G., Geluk, M.,
and Stephenson, R.A., 2015, Sedimentary geology of the Middle Carboniferous of the Donbas region (Dniepr-Donets
Basin, Ukraine), Scientific Reports 5, 9099, doi: 10.1038/srep09099
van Hinsbergen, D.J.J., Hafkenscheid, E., Spakman, W., Meulenkamp, J.E., Wortel, M.J.R., 2005. Nappe stacking resulting
from subduction of oceanic and continental lithosphere below Greece. Geology 33, 325–328.

209
van Hinsbergen, D.J.J., Kapp, P., Dupont-Nivet, G., Lippert, C., DeCelles, G., Torsvik, T.H., 2011a. Restoration of Cenozoic
deformation in Asia, and the size of Greater India. Tectonics 30, TC5003–doi:10.1029/2011TC002908.
van Hinsbergen, D.J.J., Kaymakci, N., Spakman, W., Torsvik, T.H., 2010. Reconciling the geological history of western Turkey
with plate circuits and mantle tomography. Earth and Planetary Science Letters 297, 674–686.
van Hinsbergen, D.J.J., Li., S., Lippert, P.C., Huang, W., Advokaat, E.L., and Spakman, W., 2017, Reconstructing the
paleogeography and subduction geodynamics of Greater India: how to apply Ockham’s Razor? Tectonophysics,under
review
van Hinsbergen, D.J.J., Peters, K., Maffione, M., Spakman, W., Guilmette, C., Thieulot, C., Plumper, O., Gürer, D., Brouwer,
F.M., Aldanmaz, E., Kaymakci, N., 2015. Dynamics of intra-oceanic subduction initiation, part 2: supra-subduction zone
ophiolite formation and metamorphic sole exhumation in context of absolute plate motions. Geochemistry, Geophysics,
Geosystems 16, 1771-1785.
van Hinsbergen, D.J.J., Spakman, W., Vissers, R.L.M., van der Meer, D.G., 2017. Comment on “Assessing discrepancies
between previous plate kinematic models of Mesozoic Iberia and their constraints” by Barnett-Moore et al.Tectonics,
submitted
van Hinsbergen, D.J.J., Steinberger, B., Doubrovine, V., Gassmöller, R., 2011b. Acceleration and deceleration of India-Asia
convergence since the Cretaceous: Roles of mantle plumes and continental collision. Journal of Geophysical Research
116, B06101–doi: 10.1029/2010JB008051.
van Hinsbergen, D.J.J., van der Meer, D.G., Zachariasse, W.J., and Meulenkamp, J.E., 2006, Deformation of western Greece
during Neogene clockwise rotation and collision with Apulia: International Journal of Earth Sciences 95, p. 463-490.
van Hinsbergen, D.J.J., Vissers, R., Spakman, W., 2014. Origin and consequences of western Mediterranean subduction,
rollback, and slab segmentation. Tectonics 33, 393–419.
Van Kranendonk, M.J., Kirkland, C.L., 2016. Conditioned duality of the Earth system: Geochemical tracing of the
supercontinent cycle through Earth history. Earth-Sci. Rev. 160, 171–187. doi:10.1016/j.earscirev.2016.05.009
Vaughan, A.P.M., Pankhurst, R.J., Fanning, C.M., 2002. A mid-Cretaceous age for the Palmer Land event, Antarctic
Peninsula: implications for terrane accretion timing and Gondwana palaeolatitudes. Journal of the Geological Society of
London 159, 113–116.
Veevers, J.J., 1990. Tectonic-climatic supercycle in the billion-year plate-tectonic eon: Permian Pangean icehouse alternates
with Cretaceous dispersed-continents greenhouse. Sediment. Geol. 68, 1–16. doi:10.1016/0037-0738(90)90116-B
Verard, C., Hochard, C., Baumgartner, P.O., Stampfli, G.M., 2015. 3D palaeogeographic reconstructions of the Phanerozoic
versus sea-level and Sr-ratio variations. J. Paleogeography 4, 64–84. doi:10.3724/SP.J.1261.2015.00068
Villagomez, D., Spikings, R., Magna, T., Kammer, A., Winkler, W., Beltrán, A., 2011. Geochronology, geochemistry and
tectonic evolution of the Western and Central cordilleras of Colombia. Lithos 125, 875–896.
Vincent, S.J., Braham, W., LavrishchevA., Maynard, J.R., Harland, M., 2017. The formation and inversion of the western
Greater Caucasus Basin and the uplift of the western Greater Caucasus: Implications for the wider Black Sea region.
Tectonics, 36 doi: 10.1002/2016TC004204.
Vincent, S.J., Carter, A., LavrishchevA., Rice, S.P., Barabadze, T.G., Hovius, N., 2010. The exhumation of the western Greater
Caucasus: a thermochronometric study. Geological Magazine 148, 1–21.
Vincent, S.J., Morton, A.C., Carter, A., Gibbs, S., Barabadze, T.G., 2007. Oligocene uplift of the Western Greater Caucasus:
an effect of initial Arabia?Eurasia collision. Terra Nova 19, p. 160–166.
Vishnevskaya, S., Filatova, N.I., 2012. Allochthonous Mesozoic marine sequences of northeastern Asia and western North
America: correlation of stratigraphic and geodynamic depositional settings. Russian Journal of Pacific Geology, 6,
189–208.
Vissers, R., van Hinsbergen, D., Meijer, T., Piccardo, G.B., 2013. Kinematics of Jurassic ultra-slow spreading in the Piemonte
Ligurian ocean. Earth and Planetary Science Letters 380, 138–150.
Vissers, R.L.M., Meijer, T., 2012a. Iberian plate kinematics and Alpine collision in the Pyrenees. Earth-Science Reviews 114,
61–83.

210
Vissers, R.L.M., Meijer, T., 2012b. Mesozoic rotation of Iberia: Subduction in the Pyrenees?. Earth-Science Reviews 110,
93–110.
Vissers, R.L.M., van Hinsbergen, D.J.J., van der Meer, D.G., Spakman, W., 2016. Cretaceous slab break-off in the Pyrenees:
Iberian plate-kinematics in paleomagnetic and mantle reference frames. Gondwana Research 34, 49-59
Vollstaedt, H., Eisenhauer, A., Wallmann, K., Böhm, F., Fietzke, J., Liebetrau, V., Krabbenhöft, A., Farkaš, J., Tomašových, A.,
Raddatz, J., Veizer, J., 2014. The Phanerozoic δ88/86Sr record of seawater: New constraints on past changes in oceanic
carbonate fluxes. Geochim. Cosmochim. Acta 128, 249–265. doi:10.1016/j.gca.2013.10.006
Von Bertalanffy, L., 1968. General Systems Theory. George Brazillier: New York.
von Blanckenburg, F., Davies, J.H., 1995. Slab breakoff: A model for syncollisional magmatism and tectonics in the Alps.
Tectonics 14, 120–131.
von Quadt, A., Moritz, R., Peytcheva, I., Heinrich, C.A., 2005. 3: Geochronology and geodynamics of Late Cretaceous
magmatism and Cu–Au mineralization in the Panagyurishte region of the Apuseni–Banat–Timok–Srednogorie belt,
Bulgaria. Ore Geology Reviews 27, 95–126.
Vrielynck, B., Bouysse, P., 2003. The Changing Face of the Earth; The break-up of Pangaea and continental drift over the past
250 million years in ten steps, Earth Sciences series. UNESCO Publishing / Commission for the Geological Map of the
World (CGMW).
Wakabayashi, J., 2015. Anatomy of a subduction complex: architecture of the Franciscan Complex, California, at multiple
length and time scales. International Geology Review 57, 669-746.
Wakabayashi, J., Dilek, Y., 2000. Spatial and temporal relationships between ophiolites and their metamorphic soles: a test of
models of forearc ophiolite genesis. Geological Society of America Special Paper 349, 53–64.
Wakabayashi, J., Dumitru, T.A., 2007. 40Ar/39Ar Ages from Coherent, High- Pressure Metamorphic Rocks of the
Franciscan Complex, California: Revisiting the Timing of Metamorphism of the World's Type Subduction .
International Geology Review 49, 1–36.
Walker, J.C.G., Lohmann, K.C., 1989. Why the oxygen isotopic composition of sea water changes with time. Geophys. Res.
Lett. 16, 323–326. doi:10.1029/GL016i004p00323
Walker, L.J., Wilkinson, B.H., Ivany, L.C., 2002. Continental Drift and Phanerozoic Carbonate Accumulation in Shallow
Shelf and Deep Marine Settings. J. Geol. 110, 75–87. doi:10.1086/324318
Wanke, M., Portnyagin, M., Hoernle, K., Werner, R., Hauff, F., van den Bogaard, P., Garbe-Schonberg, D., 2012. Bowers
Ridge (Bering Sea): An Oligocene-Early Miocene island arc. Geology 40, 687–690.
Ward, L., 1995. Subduction cycles under western North America during the Mesozoic and Cenozoic eras. GSA Special Paper
299, 1–40.
Warren, C.J., Parrish, R.R., Searle, M.P., Waters, D.J., 2003. Dating the subduction of the Arabian continental margin
beneath the Semail ophiolite, Oman. Geology 31, 889–892.
Wegner, W., Worner, G., Harmon, R.S., Jicha, B.R., 2011. Magmatic history and evolution of the Central American Land
Bridge in Panama since Cretaceous times. Geological Society of America Bulletin 123, 703–724.
Wei, W., Xu, J., Zhao, D., Shi, Y., 2012. Journal of Asian Earth Sciences. Journal of Asian Earth Sciences 60, 88–103.
Wells, R., Bukry, D., Friedman, R., Pyle, D., Duncan, R., Haeussler, P. and Wooden, J., 2014. Geologic history of Siletzia, a
large igneous province in the Oregon and Washington Coast Range: Correlation to the geomagnetic polarity time scale
and implications for a long-lived Yellowstone hotspot. Geosphere, 10(4), 692-719.
Wells, R.E., McCaffrey, R., 2013. Steady rotation of the Cascade arc. Geology 41, 1027–1030.
West, J.D., Fouch, M.J., Roth, J.B., Elkins-Tanton, L.T., 2009. Vertical mantle flow associated with a lithospheric drip beneath
the Great Basin. Nature Geoscience 2, 439–444.
Widiyantoro, S., Gorbatov, A., Kennett, B. L. N., Fukao, Y., 2000. Improving global shear wave traveltime tomography
using three-dimensional ray tracing and iterative inversion. Geophys Journal International 141-3, 747-758. doi:
10.1046/j.1365-246x.2000.00112.x

211
Widiyantoro, S., Kennett, B.L.N., van der Hilst, R.D., 1999. Seismic tomography with P and S data reveals lateral variations
in the rigidity of deep slabs. Earth and Planetary Science Letters 173, 91–100.
Widiyantoro, S., Pesicek, J.D., Thurber, C.H., 2011. Subducting slab structure below the eastern Sunda arc inferred from non-
linear seismic tomographic imaging. Geological Society, London, Special Publications 355, 139–155.
Widiyantoro, S., van der Hilst, R., 1996a. Structure and Evolution of Lithospheric Slab Beneath the Sunda Arc, Indonesia.
Science 271, 1566–1570.
Widiyantoro, S., van der Hilst, R.D., 1996b. Structure and evolution of subducted lithosphere beneath the Sunda arc,
Indonesia. Science 271, 1566–1570.
Widiyantoro, S., van der Hilst, R.D., 1997. Mantle structure beneath Indonesia inferred from high-resolution tomographic
imaging. Geophysical Journal International 130, 167–182.
Williams, S., Flament, N., Müller, R.D., and Butterworth, N., 2015, Absolute plate motions since 130 Ma constrained by
subduction zone kinematics: Earth and Planetary Science Letters, v. 418, no. C, p. 66–77, doi: 10.1016/j.epsl.2015.02.026
Wilson, J.T., 1966. Did the Atlantic Close and then Re-Open? Nature 211, 676–681. doi:10.1038/211676a0
Worsley, T.R., Moody, J.B., Nance, R.D., 1985. Proterozoic to Recent Tectonic Tuning of Biogeochemical Cycles, in:
Sundquist, E.T., Broecker, W.S. (Eds.), The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to
Present. American Geophysical Union, pp. 561–572.
Worsley, T.R., Moore, T.L., Fraticelli, C.M., Scotese, C.R., 1994. Phanerozoic CO2 levels and global temperatures inferred
from changing paleogeography. Geol. Soc. Am. Spec. Pap. 288, 57–74. doi:10.1130/SPE288-p57
Worsley, T.R., Nance, D., Moody, J.B., 1984. Global tectonics and eustasy for the past 2 billion years. Mar. Geol. 58, 373–400.
doi:10.1016/0025-3227(84)90209-3
Wortel, M.J.R., Spakman, W., 1992. Structure and dynamics of subducted lithosphere in the Mediterranean region.
Proceedings van de Koninklijke Nederlandse Academie van Wetenschappen 95, 325–347.
Wortel, M.J.R., Spakman, W., 2000. Subduction and slab detachment in the Mediterranean-Carpathian region. Science 290,
1910–1917.
Wu, H., Li, C., Hu, P., Li, X., 2014. Early Cretaceous (100–105 Ma) Adakitic magmatism in the Dachagou area, northern
Lhasa terrane, Tibet: implications for the Bangong–Nujiang Ocean subduction and slab break-off. International Geology
Review 57, 1172-1188.
Wu, H., Li, C., Xu, M., Li, X., 2015. Early Cretaceous adakitic magmatism in the Dachagou area, northern Lhasa terrane,
Tibet: Implications for slab roll-back and subsequent slab break-off of the lithosphere of the Bangong-Nujiang Ocean.
Journal of Asian Earth Sciences 97, 51–66.
Wu, J., Suppe, J., Lu, R., Kanda, R., 2016. Philippine Sea and East Asian plate tectonics since 52 Ma constrained by new
subducted slab reconstruction methods: Journal of Geophysical Research 121, 4670–4741.
Xiao, W.J., Huang, B., Han, C., Sun, S., Li, J., 2010. A review of the western part of the Altaids: A key to understanding the
architecture of accretionary orogens. Gondwana Research 18, 253–273.
Xu, M., LI, C., Zhang, X., Wu, Y., 2014. Nature and evolution of the Neo-Tethys in central Tibet: synthesis of ophiolitic
petrology, geochemistry, geochronology. International Geology Review 56, 1072–1096.
Yin, A., Harrison, T.M., 2000. Geologic evolution of the Himalayan-Tibetan orogen. Annual Review of Earth and Planetary
Sciences 28, 211–280.
Zahirovic, S., Matthews, K.J., Flament, N., Müller, R.D., Hill, K.C., Seton, M., Gurnis, M., 2016. Tectonic evolution and
deep mantle structure of the eastern Tethys since the latest Jurassic. Earth-Science Reviews 162, 293–337.
Zahirovic, S., Müller, R.D., Seton, M., Flament, N., Gurnis, M., Whittaker, J., 2012. Insights on the kinematics of
the India-Eurasia collision from global geodynamic models. Geochemistry, Geophysics, Geosystems 13, doi:
10.1029/2011GC003883.
Zahirovic, S., Seton, M., Müller, R.D., 2014. The Cretaceous and Cenozoic tectonic evolution of Southeast Asia. Solid Earth
5, 227–273.

212
Zeng, M., Zhang, X., Cao, H., Ettensohn, F.R., Cheng, W., Lang, X., 2015. Late Triassic initial subduction of the Bangong-
Nujiang Ocean beneath Qiangtang revealed: stratigraphic and geochronological evidence from Gaize, Tibet. Basin
Research 28, 147-157.
Zhang, J., Shan, X., Huang, X., 2011. Seismotectonics in the Pamir: An oblique transpressional shear and south-directed
deep-subduction model. Geoscience Frontiers 2, 1–15.
Zhao, D. and Lei, J., 2004. Seismic ray path variations in a 3D global velocity model. Physics of the Earth and Planetary
Interiors, 141(3), pp.153-166.
Zhao, D., Hasegawa, A. and Horiuchi, S., 1992. Tomographic imaging of P and S wave velocity structure beneath
northeastern Japan. Journal of Geophysical Research: Solid Earth, 97(B13), pp.19909-19928.
Zhao, D., Ohtani, E., 2009. Deep slab subduction and dehydration and their geodynamic consequences: Evidence from
seismology and mineral physics. Gondwana Research 16, 401–413.
Zhao, D., Pirajno, F., Dobretsov, N.L., Liu, L., 2010. Mantle structure and dynamics under East Russia and adjacent regions.
RGG 51, 925–938.
Zheng, H.-W., Gao, R., Li, T.-D., Li, Q.-S., He, R.-Z., 2013. Collisional tectonics between the Eurasian and Philippine Sea
plates from tomography evidences in Southeast China 606, 14–23.
Zhu, D.-C., Zhao, Z.-D., Niu, Y., Dilek, Y., Hou, Z.-Q., Mo, X.-X., 2013. The origin and pre-Cenozoic evolution of the
Tibetan Plateau. Gondwana Research 23, 1429–1454.
Zhu, D.-C., Zhao, Z.-D., Niu, Y., Mo, X.-X., Chung, S.-L., Hou, Z.-Q., Wang, L.-Q., Wu, F.-Y., 2011. The Lhasa Terrane:
Record of a microcontinent and its histories of drift and growth. Earth and Planetary Science Letters 301, 241–255.
Zhu, H., Bozdağ, E., Peter, D., Tromp, J., 2012. Structure of the European upper mantle revealed by adjoint tomography 5,
493–498.
Ziegler, A., 1989. Evolution of Laurussia. Kluwer Academic Publications, Dordrecht, 102p.
Zimmerman, A., Stein, H.J., Hannah, J.L., Kozelj, D., Bogdanov, K., Berza, T., 2007. Tectonic configuration of the Apuseni–
Banat—Timok–Srednogorie belt, Balkans-South Carpathians, constrained by high precision Re–Os molybdenite ages.
Mineralium Deposita 43, 1–21.
Zor, E., 2008. Tomographic evidence of slab detachment beneath eastern Turkey and the Caucasus. Geophysical Journal
International 175, 1273–1282.

213
Samenvatting in het Nederlands
De aarde kan worden onderverdeeld in drie delen, een harde korst en onderliggende lithosfeer, de
meer vloeibare mantel en de kern. De harde korst is verdeeld in een tiental grote tektonische platen
die onafhankelijk van elkaar bewegen met snelheden in de orde van grootte van centimeters per jaar.
Waar ze van elkaar spreiden leidt dit tot opbreking van de korst en de vorming van continentale
en mid-oceanische ruggen. Waar platen naar elkaar toe bewegen botsen de platen met als gevolg
gebergtevorming aan de oppervlakte. Vervolgens duikt een van de platen in de mantel, subductie
genoemd.
Het proces van spreiding en botsingen van platen wordt platentektoniek genoemd en heeft directe
gevolgen op de geografie van de aarde. Met behulp van geologische informatie trachten paleo-
geografische en plaat-tektonische studies te reconstructuren hoe de wereld er mijoenen jaren
geleden er uit zag. Deze reconstructies worden meer en meer gebruikt als input voor geodynamische
en paleo-klimatologisch onderzoek.

In geodynamische studies vormen plaat-tektonische reconstructies randvoorwaarden voor plaat-


tektonische en mantelconvectie modellering, terwijl lithosfeer subductie zoals voorspeld uit
plaattektonische reconstructie, kan worden gebruikt als inzicht voor het bepalen van neerwaartse
mantelbeweging. Paleogeografische reconstructies leveren de basis voor paleoklimaat en paleo-
mileu wat betreft land-zee ratios en eerste order bathymetrie. Plaat tektonische reconstructiees
kunnen ook worden gebruikt om geochemische cycli van mantel inputs (bijvoorbeeld vulkanisme)
in de atmosfeer en zeewater van chemische componenten zoals CO2 en strontium te kwantificeren.
Tevens, door middel van schatting van de gemiddelde ouderdom van oceanen door korstspreiding,
zijn reconstructies nuttig om de invloed van platentektoniek op zeeniveau te bepalen.

Het wetenschappelijk inzicht in de aarde neemt af naarmate men verder terug in de tijd kijkt. Dit
komt voor een groot deel omdat de oceanische korst is gesubduceerd en is vervangen door nieuwe
korst door spreiding. Dit gebrek aan informatie neemt toe verder terug in de tijd en de platen
tektoniek vóór 180 miljoen jaar geleden (vroeg Jura) is daarmee voor ~70% onbekend. Daarom is
de oceaanspreidingsgeschiedenis (belangrijk voor zee-niveau reconstructie) en plaattektonische
activiteit in het algemeen (belangrijk voor vulkanische ontgassing van CO2) steeds minder goed
bepaald verder terug in de tijd. De locatie en leeftijden van intra-oceanische subductiezones in de
Panthalassa en Tethys Oceanen, zijn andere voorbeelden waar onderzoek nadeel ondervindt door
een gebrek aan observationele data. Deze subductiezones zijn daarom moeilijk te reconstructueren
en worden daarom vaak weggelaten uit wereldwijde paleogeografische reconstructies.

Verder terug in de tijd, komen er extra onzekerheden, wanneer relatieve plaat tektonische
reconstructies in zogenaamde mantelreferentiekaders worden geplaatst. In het kader van de
diepe mantel, wordt dit ‘absolute‘ positie en ‘absolute’ plaatbeweging genoemd. Plaat-tektonische
reconstructies in een mantelreferentiekader zijn belangrijk voor het correleren van mantelprocessen
(bijv. mantel convectie, mantelpluimen, opsmelten en subductie) met plaat-tektonische evolutie en
zijn daarom een voorwaarde voor dynamische modellering van de planeet. Klassieke technieken
op basis van hotspots om de absolute locaties en beweging van platen te bepalen kunnen
niet worden toegepast voor het Krijt (65-145 miljoen jaar geleden) vanwege het gebrek aan
overgebleven sporen van hotspot-vulkanisme. Daarbij zijn laterale bewegingen of afbuigingen van

214
mantelpluimen onzeker. Alhoewel er moderne hypotheses bestaan wat betreft de diepe oorsprong
van mantelplumen, grote vulkanische gebieden en kimberlieten, en daarmee absolute plaatbeweging
kan worden gereconstrueerd, is de hoeveelheid data hiervoor beperkt.

Dit proefschrift richt zich er op om te zoeken naar nieuwe manieren om plaat-tektonische, paleo-
geografie, paleo-atmosfeer en paleo-zeeniveau reconstructies beter te kunnen bepalen, met behulp
van paleo-subductie. Hiervoor gebruik ik drie-dimensionale aardbevingsgolven-snelheidsmodellen
van de mantel. Deze ‘tomografische’ modellen zijn gebaseerd op de reistijden van duizenden
aardbevingsgolven die door de Aarde zijn gereisd. In de bovenmantel (<660 km diepte), waar
aardbevingensgolven sneller reisden, komen deze veelal overeen waar subductie momenteel nog
plaatsvindt onder huidige gebergteketens. In de ondermantel (660-2890 km diepte) zijn de snellere
delen van de mantel mogelijk ook gesubduceerde platen, maar slechts weinig studies hadden
deze snelheidsanomalieeen gecorreleerd met een geologische geschiedenis. De kern van mijn
onderzoek was het in kaart brengen van deze snelle gebieden en correleren met de geologie aan
het oppervlakte. Hiermee werden cruciale nieuwe inzichten verkregen op diverse vlakken, hieronder
nader beschreven per hoofdstuk:

Opbouw van het proefschrift:

Hoofdstuk 1: Definitie van de methode


Dit hoofdstuk gaat over de eerste interpretatie van de ondermantel snelheidsstructuur gebaseerd op
het tomografie model UU-P07. De focus was het identificeren van gesubduceerde platen, ‘slabs’, in
ruimte en tijd, op een consistente manier door het correleren van de locatie, orientatie en helling
van snelheidsanomalieeen met de geologische geschiedenis van subductie, zoals de start en eind
van gebergtevorming of vulkanisme. Drie ondermantel slabs waren voorheen beschreven (Farallon,
Aegean Tethys, Mongol-Okhotsk) en werden geselecteerd als referentiekader om de interpretatie
de weg te wijzen. De interpretatiewerd was gebaseerd op het startpunt dat in het algemeen, diepere
slabs eerder waren gesubduceerd dan oudere slabs, leidend tot een subductieleeftijd-diepte relatie.
In eerste instantie werden 28 slabs in kaart gebracht (Hoofdstuk 1) en zeven jaar later resulteerde
dit in de Atlas van de Onderwereld met ~100 slabs (Hoofdstuk 5). Hoofdstuk 1 is de ruggegraat
van de filosofie en methode hoe tomografische informatie van de overblijfselen van voormalige
subductie kan worden gebruikt om paleo-longitude in relatieve plaat-tektonische reconstructies te
bepalen: subductiezones in plaattektonische reconstructies worden geplaatst boven de plekken van
de corresponderende gesubduceerde slabs in de mantel. Dit wordt ook wel slab-fitting genoemd en
leidt tot een nieuw mantelreferentiekader, hier het subductiereferentiekader genoemd.

Hoofdstuk 2: Toepassing van de methode om intra-oceanische subductie te bepalen in een


paleo-geografische reconstructie
De methode van Hoofdstuk 1 biedt een nieuwe manier om de absolute locatie van plaat-
tektonische reconstructies ten opzichte van de mantel te bepalen door middel van ‘slab-fitting’.
Gebruik makend van de tijd-diepte relatie van Hoofdstuk 1, onderzoek ik in Hoofdstuk 2 in
hoeverre intra-oceanische subductie zones kunnen worden gereconstrueerd in de Panthalassa
Oceaan. Dit doe ik door het correleren van allochtone tektonische blokken, samengegroeid met de
Aziatische en Amerikaanse continentale randen, met locaties in de (centrale) Panthalassa Oceaan
op basis van mantelstructuur en plaat-tektonische transport schattingen.

215
Hoofdstuk 3: Mantelstructuur gebruiken om CO2 in de atmosfeer terug in de tijd te bepalen
Op basis van de wereldwijde identificatie van slabs in Hoofstukken 1 en 2, wordt een schatting
gemaakt van de totale laterale slab lengte varieërend met diepte. Met de bepaalde tijd-diepte relatie,
wordt dit geconverteerd naar subductie-zone lengte terug in de tijd.
Dit wordt gecalibreerd met een korst-productie curve voor de laaste 140 miljoen jaar. Met de
aanname dat de hoeveelheid koolstof en convergentiesnelheiden van subducerende platen niet
significant veranderen, kan subductielengte door de tijd een eerste orde schatting geven van
vulkanische ontgassing in de atmosfeer. Tevens geeft dit een schatting van paleo-strontium in
de oceanen en zeeën. De genormaliseerde subductielengte door de tijd heen, werd gebruikt om
paleo-strontium te berekenen en als input voor het GEOCARBSULF koolstof-cyclus model om
te berekenen in hoeverre plaat-tektoniek CO2 in de atmosfeer gedurende het Mesozoicum en
Cenozoicum bepaalt.

Hoofdstuk 4: Een paleo-zeeniveau reconstructie gebaseerd op plaattektonische activiteit


schattingen
Paleo-strontium informatie (Hoofdstuk 3), gecorrigeerd voor strontium input afkomstig van
verwering, levert een schatting van plaattektonische activiteit op, dat kan worden gecorreleerd met
zowel oceanische spreiding en subductie. Variaties in de gecorrigeerde strontium data, representeren
oceaan spreiding variaties, die korstouderdom en oceandiepte bepaalt en daarmee hoeveel
water oceanen bevatten. Dit leverde een berekening van de oceanische opbergruimte. Wanneer
opbergruimte verminderde, werd het surplus aan water herverdeeld over het totale oppervlakte van
oceanen en het overstroomde continentale plat. De resulterende absolute zeeniveau curves werd
vervolgens vergeleken met ‘conventionele’ zeeniveau curves op basis van sequentie-stratigrafie en
plaatbewegingmodellen.

Hoofdstuk 5: De Atlas van de Onderwereld, een catalogus van slab overblijfselen in de


mantel zichtbaar in seismische tomografie en hun geologische interpretatie
Dit laatste hoofdstuk is de afsluiting van 15 jaar (part-time) onderzoek, gedurende dit proefschrift
werd ontwikkeld: De Atlas van de Onderwereld, documenteert de identificatie en associatie van
~100 slabs en slab overblijfselen in de boven- en ondermantel met hun geologische oorsprong. Dit
is een uitbreiding van de slab catalogus van Hoofdstuk 1. Deze uitgebreide catalogus levert een
basis voor verdere studies in het subductie-gebaseerde mantelreferentiekader, terwijl gemiddelde
zinksnelheden nieuwe informatie levert voor mantelviscositeit. Als vervolgstappen kunnen
plaattektonische activiteit, vulkanische ontgassing van CO2 en paleo-zeeniveau door de tijd weer
verder worden uitgewerkt.

216
Dankwoord (Acknowledgements)
Ik ben vele mensen direct of indirect dank verschuldigd voor het welslagen van dit promotieproject.
Na 15 jaar semi-continue part-time wetenschap, is de vraag geoorloofd hoe dit alles begon. Ik dank
mijn promotor Wim Spakman voor de onmogelijke opdracht van het tomografisch begrijpen van
het Caribisch gebied voor mijn Drs. afstudeerproject, wat uiteindelijk leidde tot een begrip van een
derde van de mantel. Ik dank hem ook voor de argwaan over zijn eigen data en het temperen van
mijn wilde speculaties. Het resultaat was dat ‘borrelpraat’ plaats maakte voor goed onderbouwde
argumentatie en foutenberekeningen. Ik dank het KNGMG voor de M.C. Escher prijs in 2002,
wat mij stimuleerde om koste wat kost te publiceren, al duurde dat 8 jaar.
Mijn co-promoter Douwe van Hinsbergen dank ik voor de jarenlange vriendschap. tevens
voor het Griekse veldwerk en bijbehorende avontuur met kalashnikovs/handboeien/cel/rechtszaak.
Dit leverde een levenslang blijvende herinnering op en droeg bij tot het verkrijgen van een baan
in de olieindustrie en een fantastische vrouw. Daarnaast dank ik hem voor het enthousiasme
en ondersteuning op vele vlakken gedurende de vele jaren, inclusief het ophalen van de NAC
posterprijs in 2009. Dit was wederom een stimulans om Terra Incognita in de diepte gepubliceerd
te krijgen.
Ik dank mijn voormalige en huidige collega’s van Shell en Nexen en met name Frank
van Bergen voor de geologische discussies over de tectono-stratigrafische ontwikkeling van
bekkens en petroleum geologie over de gehele wereld. Dit leverde toepassingen op van mijn
onderzoeksinteresses en leidde tot een verbreding naar zeespiegel en paleo-klimaat. Tevens dank
ik Arjen Crince, Ruben van de Weg, Peggy Aggreh en Jon Webb voor GIS support. Tenslotte
dank mijn verschillende bazen en mentors gedurende de jaren voor de erkenning van een soort
van expertise en het mij laten gaan naar diverse congressen, opnemen van studieverlof en financiële
ondersteuning.
I would like to thank Trond Torsvik and NGU colleagues for not hiring me but hosting me
at their institute for three months in 2009. This provided a welcome break of finding oil, enabling
focus of brainpower and resulting in several articles in the years to come. Exercising along
the Norwegian fjords and in the highlands in combination with a healthy diet of caught fish
contributed to a wonderful daughter. Most stimulating summer ever!
NewScientist, der Spiegel, Nu.nl, Scientas, BNR Nieuwsradio, diverse kranten en blogs dank
ik voor de publieke aandacht voor mijn onderzoek door de jaren heen. Ik dank het Geologisch
College Miölnir en met name de Mythologische Commissie voor de inspiratie voor werknamen
voor vele slabs, paleo-oceanen en verloren micro-continenten. Ik dank Aart-Peter van den Berg van
Saparoea en Appy Sluijs voor het me verklaren waarom paleo-zeespiegel en vulkanische ontgassing
problemen zijn. Het hielp enorm om een probleem te definiëren, vooral als de oplossing al was
uitgewerkt.
I would like to thank my other co-authors, including Maisha Amaru, Richard Zeebe, Yves
Godderis, Andy Biggin, Reinoud Vissers, Guillaume Le Hir, Yannick Donnadieu, some who were
willing to cooperate with a stranger known only to them by email. Maybe we should actually meet?
Ik dank Thomas van der Linden voor het ontwerpen van de Atlas-of-the-underworld.org website.
Tevens dank ik de leescommissie en reviewers van voorgaande artikelen voor de tijd en feedback die
de manuscripten aanscherpten. Ik dank mijn ouders, familie, vrienden voor het luisteren naar mijn
gewauwel over de afgelopen 15 jaar. Maar bovenal dank ik Yvonne en Ariane voor het begrip als ik
weer eens avonden en weekenden aan het zombieën was voor een volgend artikel.

217
Curriculum Vitae
2 Augustus 1977 Born in Sluiskil, The Netherlands

1988-1995 Secondary school, Petrus Hondius, Terneuzen, The Netherlands

1995-1996 Propaedeutics Geophysics, Faculty of Geosciences, Utrecht University,


The Netherlands

1996-2002 Drs. Geophysics, Faculty of Geosciences, Utrecht University, The Netherlands

1997-2001 Operations Geophysicist (part-time), Geo Resources Consultants, Rotterdam,


The Netherlands

2002-2003 Seismic Acquisition Geophysicist, Nederlandse Aardolie Maatschappij,


Assen, The Netherlands

2003-2005 Exploration Geoscientist, New Ventures Europe, Nederlandse Aardolie


Maatschappij, Assen, The Netherlands

2005-2011 Regional Geologist, New Ventures, Shell International Exploration &


Production, Rijswijk & The Hague, The Netherlands

2009 Sabbatical (3 months), Norges Geologiske Undersøkelse, Trondheim, Norway

2011-2013 Team Lead New Ventures Europe Unconventionals, Nexen Petroleum UK Ltd.,
Uxbridge, United Kingdom

2014-2016 Manager New Ventures North Atlantic Margins, Nexen Petroleum UK Ltd.,
Uxbridge, United Kingdom

2016-Present Senior Manager New Ventures Europe & Africa, Nexen Petroleum UK Ltd.
Uxbridge, United Kingdom

218
Publication list
Publications incorporated in this thesis:

van der Meer, D.G., Spakman, W., van Hinsbergen, D.J.J., Amaru, M.L., Torsvik, T.H., 2010.
Towards absolute plate motions constrained by lower-mantle slab remnants. Nature
Geoscience 3, 36–40. doi:10.1038/ngeo708.
van der Meer, D.G., Torsvik, T.H., Spakman, W., van Hinsbergen, D.J.J., Amaru, M.L., 2012.
Intra-Panthalassa Ocean subduction zones revealed by fossil arcs and mantle structure.
Nature Geoscience 5, 215–219. doi:10.1038/ngeo1401.
van der Meer, D.G., Zeebe, R.E., van Hinsbergen, D.J.J., Sluijs, A., Spakman, W., Torsvik, T.H.,
2014. Plate tectonic controls on atmospheric CO2 levels since the Triassic. Proceedings of
the National Academy of Sciences 111, 4380–4385. doi:10.1073/pnas.1315657111.
van der Meer, D.G., van den Berg van Saparoea, A.P., van Hinsbergen, D.J.J., van de Weg, R.,
Godderis, Y., Le Hir, G., and Donnadieu, Y., 2017, Reconstructing first-order changes in
sea level during the Phanerozoic and Neoproterozoic using strontium isotopes, Gondwana
Research 44, p. 22-34, doi:10.1016/j.gr.2016.11.002.
van der Meer, D.G., van Hinsbergen, D.J.J., Spakman, W., submitted, Atlas of the Underworld:
slab remnants in the mantle, their sinking history, and a new outlook on lower mantle
viscosity, Tectonophysics.

Other publications:

van Hinsbergen, D.J.J., van der Meer, D.G., Zachariasse, W.J., and Meulenkamp, J.E., 2006,
Deformation of western Greece during Neogene clockwise rotation and collision with
Apulia: International Journal of Earth Sciences 95, p. 463-490.
Meijers, M.J.M, Hamers, M.F., van Hinsbergen, D.J.J., van der Meer, D.G., Kitchka, A., Langereis,
C.G., and Stephenson, R.A., 2010, New late Paleozoic paleopoles from the Donbas
Foldbelt (Ukraine): implications for the Pangea A vs. B controversy, Earth and Planetary
Science Letters 297, p. 18-33.
Biggin, A.J., Steinberger, B., Aubert, J., Suttie, N., Holme, R., Torsvik, T.H., van der Meer, D.G.,
and van Hinsbergen, D.J.J., 2012, Possible links between long term geomagnetic variations
and whole-mantle convection processes, Nature Geoscience 5, 526-533.
van Hinsbergen, D.J.J., Abels, H.A., Bosch, W., Boekhout, F., Kitchka, A., Hamers, M.F., van der
Meer, D.G., Geluk, M., and Stephenson, R.A., 2015, Sedimentary geology of the Middle
Carboniferous of the Donbas region (Dniepr-Donets Basin, Ukraine), Scientific Reports 5,
9099, doi: 10.1038/srep09099.
Vissers, R.L.M., van Hinsbergen, D.J.J., van der Meer, D.G., and Spakman, W., 2016, Cretaceous
slab break-off in the Pyrenees: Iberian plate-kinematics in paleomagnetic and mantle
reference frames, Gondwana Research, v. 34, p. 49-59.
van Hinsbergen, D.J.J., Spakman, W., Vissers, R.L.M., van der Meer, D.G., in press, Comment on
“Assessing discrepancies between previous plate kinematic models of Mesozoic Iberia and
their constraints” by Barnett-Moore et al., Tectonics.

219
Appendices

220
Table A1. Table of interpreted slabs and their orogens

221
222
223
224
Figure A1. Global paleo-subduction zone interpretation from seismic tomography. Mollweide
projections centred at 0° longitude. Left) Tomographic slices of UU-P07 model (Amaru, 2007)
projected to earth’s surface. Right) modified reconstruction of Torsvik et al. (2008) at corresponding
averaged time of subduction (van der Meer et al., 2010, Chapter 5). Paleolongitudes of continents
were modified on the basis of slab-fitting (van der Meer et al. 2010). Interpreted subduction zones
in red.

225
226
227
228
Figure A2. Global paleo-subduction zone interpretation from seismic tomography. Mollweide
projections centred at 180° longitude. Left) Tomographic slices of UU-P07 model (Amaru, 2007)
projected to earth’s surface. Right) modified reconstruction of Torsvik et al. (2008) at corresponding
averaged time of subduction (van der Meer et al., 2010, Chapter 5). Paleolongitudes of continents
were modified on the basis of slab-fitting (van der Meer et al. 2010). Interpreted subduction zones
in red.

229

You might also like