Download PALAEOMAGNETISM, PLATE MOTION AND POLAR WANDER

Document related concepts

Spherical Earth wikipedia , lookup

Earth's magnetic field wikipedia , lookup

Geochemistry wikipedia , lookup

Ocean wikipedia , lookup

Post-glacial rebound wikipedia , lookup

Polar ecology wikipedia , lookup

Physical oceanography wikipedia , lookup

Age of the Earth wikipedia , lookup

Geomagnetic reversal wikipedia , lookup

Global Energy and Water Cycle Experiment wikipedia , lookup

Nature wikipedia , lookup

History of Earth wikipedia , lookup

Magnetotellurics wikipedia , lookup

Geology wikipedia , lookup

History of geomagnetism wikipedia , lookup

Earthscope wikipedia , lookup

Large igneous province wikipedia , lookup

History of geology wikipedia , lookup

Geophysics wikipedia , lookup

Supercontinent wikipedia , lookup

Pangaea wikipedia , lookup

Plate tectonics wikipedia , lookup

Geological history of Earth wikipedia , lookup

Transcript
Review Paper
PALAEOMAGNETISM, PLATE MOTION AND POLAR WANDER
David PRATT
Abstract: This article examines the assumptions, uncertainties and problems associated with palaeomagnetic
data and their interpretation, and examines related theories of plate motion and polar wander. It shows that
theories based on palaeomagnetic data are often contradicted by geological and geophysical evidence,
suggesting that palaeomagnetic data are not a reliable guide to past locations of the Earth’s geographic poles.
Moreover, palaeoclimatic and palaeontological data since at least the Late Proterozoic are consistent with
continental and polar stability.
Keywords: palaeomagnetism, plate-tectonic reconstructions, polar wander, age of seafloor, continental roots,
palaeoclimate, palaeobiogeography, surge tectonics, wrench tectonics
Geomagnetism and palaeomagnetism
Earth’s magnetic field
The Earth’s north and south magnetic dip poles – the places on the Earth’s surface where a compass needle
points directly downward or upward – are currently located at 85.9ºN, 147.0ºW, and 64.4ºS, 137.1ºE
respectively; they are therefore not antipodal. About 90% of the Earth’s magnetic field – the ‘main field’ –
corresponds to the field that would be produced by a hypothetical dipole magnet located at the Earth’s
centre, but with its axis offset from the Earth’s rotation axis by about 10°. The two locations where the
dipole axis intersects the Earth’s surface are known as the north and south geomagnetic poles. The dipole
field is currently moving westward around the geographic poles at a rate of about 0.08° per year, whereas the
nondipole field is moving at a faster average rate of 0.18° per year (McElhinny & McFadden, 2000). Fig. 1.1
shows the location of the north and south magnetic and geomagnetic poles and their secular motion since
1900. Fig. 1.2 shows the positions of the north geomagnetic pole over the past 2000 years.
The geomagnetic field is defined at each point on the Earth’s surface by its strength and direction. The two
terms needed to describe its direction are: magnetic declination – the angle between magnetic north (the
direction in which the north-seeking end of the compass needle points) and true north; and magnetic
inclination – the angle between a compass needle and the horizontal plane at any particular location.
The Earth is far from being a uniformly magnetized sphere. This means that a compass needle rarely points
in the direction of the nearest magnetic pole. For instance, to travel direct to the north magnetic pole, a
person setting out from Moscow will need to head 0.5° east of north, whereas the compass points 10.4° east
of north (www.ngdc.noaa.gov; www.movable-type.co.uk). A compass will eventually lead the traveller to
the north magnetic pole, but not by the most direct route. Likewise, magnetic inclination rarely corresponds
exactly to magnetic latitude. Figures 1.3 and 1.4 show the current global variations in declination and
inclination. The present geomagnetic field clearly deviates substantially from that of a perfect geocentric
axial dipole (GAD).
New Concepts in Global Tectonics Journal (www.ncgt.org), vol. 1, no. 1, March 2013, pp. 66-152
2
Fig. 1.1. Motion of magnetic and geomagnetic poles since 1900. (wdc.kugi.kyoto-u.ac.jp)
Fig. 1.2. Positions of the north geomagnetic pole over the
past 2000 years. Each data point is the mean geomagnetic
pole at 100-year intervals; numbers indicate the date in
years AD; circles around geomagnetic poles at 900, 1300,
and 1700 AD are 95% confidence limits; the mean
geomagnetic pole position over the past 2000 years is shown
by the square with stippled region of 95% confidence.
(Butler, 2004, fig. 1.9)
According to Tauxe (2013), our ability to describe the geomagnetic field ‘far exceeds our understanding of
its origin’. The prevailing view is that the field is generated mainly by motions of conducting fluid in the
Earth’s liquid outer core, which create a self-sustaining magnetohydrodynamic dynamo. Other sources of the
overall magnetic field include electrical currents flowing in the ionized upper atmosphere, currents flowing
in the Earth’s crust, and magnetized crustal rocks. The official theory implies that the average geomagnetic
field strength should increase (nearly double) with latitude, but the available data suggest that it weakens
above 65° (fig. 1.5).
3
Fig. 1.3. (http://ngdc.noaa.gov)
Fig. 1.4. (http://ngdc.noaa.gov)
4
Fig. 1.5. (a) Palaeointensity versus latitude for the past 5 Myr. Mean palaeointensity results (diamonds) are calculated
for 15° latitude bins with 95% confidence levels. Southern hemisphere data have been flipped to the northern
hemisphere. The black line represents the mean intensity for today’s field as defined by the 2005 International
Geomagnetic Reference Field model coefficients, while the red dashed line represents the intensity associated with a
geocentric axial dipole with a dipole term of 30 µT (microteslas). (b) Illustration of theoretical outer core flow regimes.
(Tauxe, 2013, fig. 14.14)
Palaeomagnetic assumptions
Study of the fossil magnetism of iron-bearing minerals in ancient rocks and sediments allows the
determination of a virtual magnetic pole relative to the location of the sample in question. Averages of a
number of virtual magnetic poles sufficient to average out secular variation are known as palaeomagnetic
poles, which are assumed to correspond to past geographic (rotation) poles. When palaeomagnetic poles for
a particular continent are plotted on a map, they tend to ‘wander’ away from the spin axis with increasing
age of the rock unit sampled. If the palaeomagnetic data are reliable and the underlying assumptions are
sound, this could mean that the continent in question is wandering while the pole remains fixed, or that the
magnetic/geographic pole is wandering while the continent remains fixed, or a combination of the two.
Because different continents yield different polar wander paths (e.g. the poles determined from Palaeozoic
and Mesozoic rocks of Europe are systematically displaced eastward from poles determined from
corresponding rocks of North America), most palaeomagneticians in the 1950s concluded that it is mainly
the continents that have wandered over the Earth’s surface, rather than the magnetic or geographic poles.
The present consensus is that there has also been a certain amount of true polar wander – i.e. a shift of the
rotation axis relative to the entire Earth (or at least its outer shell). Palaeomagnetic poles are used to specify
palaeolatitude, but palaeolongitude cannot be constrained from palaeopoles alone.
Alfred Wegener’s (1912, 1929) theory was that the continents (consisting of sial, i.e. silica and aluminium)
ploughed slowly through the denser oceanic crust (consisting of sima, i.e. silica and magnesium), whereas
modern plate tectonics postulates that new lithosphere is generated at ‘spreading ridges’ and consumed at
‘subduction zones’, and that moving lithospheric plates carry the continents with them. Apparent polar
wander paths (APWPs) tend to consist of long, gently curved tracks linked by sharp corners or ‘cusps’. The
curved tracks are assumed to correspond to periods of constant plate motion while the cusps are assumed to
correspond to sudden changes in plate motion (‘plate reorganization’). The relative motion between two
plates is described by rotation about an Euler rotation pole. Different tracks are said to represent plate
rotations about different poles.
The conventional interpretation of palaeomagnetic data is therefore founded on two basic assumptions: 1.
5
when rocks are formed, they are magnetized in the direction of the geomagnetic field existing at the time and
place of their formation, and the acquired magnetization is retained in the rocks at least partially over
geologic time; 2. the geomagnetic field averaged for any time period of the order of 105 years (except
magnetic-reversal epochs) is a dipole field oriented along the Earth’s rotation axis. Both these assumptions
are questionable. Palaeomagnetism is in fact plagued with uncertainties. Merrill et al. (1996, p. 69) state:
When a rock forms it usually acquires a magnetization parallel to the ambient magnetic field (usually presumed to be
the Earth’s magnetic field) and this is referred to as a primary magnetization. This primary magnetization then provides
information about the direction and intensity of the magnetic field in which the rock formed. However, there are
numerous pitfalls that await the unwary: first, in sorting out the primary magnetization from secondary magnetizations
(acquired subsequent to formation), and second, in extrapolating the properties of the primary magnetization to those of
the Earth’s magnetic field.
Remagnetization is now accepted as a widespread phenomenon, particularly in sedimentary rocks, and the
problem increases with rock age. Another complicating factor is that it is not always certain whether the
geomagnetic field at any given time in the past was of normal or reversed polarity (i.e. which direction was
north and which was south); this is particularly a problem for rocks of early Palaeozoic and Precambrian
age. Rock magnetism is also subject to modification by weathering, thermal effects, metamorphism,
chemical changes, and tectonic deformation. Inclination shallowing (resulting from sediment compaction),
horizontal and vertical block rotations, and other crustal motions have been identified as potential sources of
error (Butler, 2004). So has magnetostriction – the alteration of the direction of magnetization by directed
stress (Graham et al., 1957; Jeffreys, 1976).
Palaeomagneticians argue that although, at any particular time, the geomagnetic field deviates significantly
from a geocentric axial dipole (GAD), the field’s long-term secular variation (e.g. westward drift) means that
the time-averaged field will closely approximate a GAD. This assumption is fundamental to
palaeomagnetism. The general view is that analyses have demonstrated that the time-averaged field for the
past 5 Myr is approximately a GAD, with a subsidiary geocentric axial quadrupole representing about 2 to
6% of the total field, and a persistent zonal octupole representing up to 5% of the total field (Besse &
Courtillot, 2002; Domeier et al., 2012). These nondipole fields are thought to have been mostly zonal (i.e.
symmetrical about the rotation axis) and are estimated to cause palaeolatitude errors of some 5° (Van der
Voo, 1998).
Fig. 1.6. (http://web.ics.purdue.edu)
There is a persistent tendency for palaeomagnetic poles, when viewed from their sampling sites, to be
6
located to the right-hand side of the dipole. Gordon (1987) suggested that this might be due to polar wander
combined with a sampling bias. There is also a tendency for shallow inclinations to be over-represented in
frequency distributions of palaeomagnetic inclinations (Andrews, 1985; Kent & Smethurst, 1998). Kent &
Smethurst argued that inclination flattening in sedimentary rocks was not the primary cause and that
Precambrian and Palaeozoic palaeomagnetic fields may have included strong zonal quadrupole and octupole
components (estimated at 10% and 25% respectively) that diminished with time. In their view,
palaeolatitudes could be underestimated by up to about 15° in mid-latitudes. Van der Voo (1998) stated that
a palaeolatitude calculated with the GAD model could differ by up to 18° from the real palaeolatitude, and
that the relative palaeolatitude difference between two coeval palaeomagnetic sites could be in error by over
30o.
Tauxe (2013, 16.8) says: ‘There is no compelling evidence that the field has operated in a vastly different
way in ancient times, apart from the puzzling change in reversal frequency.’ During the period from 125 to
84 Ma (the Cretaceous Normal Superchron), for example, the field was predominantly of a single polarity
for 40 Myr, whereas the average reversal frequency in the last 5 Myr is 4.0/Myr (Biggin et al., 2008).
As already noted, the Earth’s magnetic axis is currently tilted by about 10° to the rotation axis. Jupiter’s
magnetic field is tilted by the same amount. On some of the outer planets much greater offsets are found: 47°
in the case of Neptune, and 60° in the case of Uranus (Russell & Dougherty, 2010). It is possible that the
Earth’s magnetic poles, too, have wandered considerably with respect to the geographic poles in former
times. Furthermore, Beloussov (1990) argued that if in past geologic periods there were stable magnetic
anomalies of the same intensity as the present-day East Asian anomaly (or slightly more intensive), this too
would render the GAD hypothesis invalid (see fig. 1.7).
Fig. 1.7. (http://ngdc.noaa.gov)
To test whether the assumed geocentric dipole was aligned with the Earth’s rotation axis in the past,
independent determinations of palaeolatitude are required, and palaeoclimatic indicators are considered the
7
best available (Butler, 2004). We will see later that several major palaeoclimatic indicators, along with
palaeontological data, are fully compatible with an Earth model that excludes large-scale wandering of the
continents and poles.
Palaeopole reliability and scatter
Over 10,000 palaeomagnetic poles have been published since 1925. The IAGA Global Paleomagnetic
Database can be freely accessed at the Norwegian Geological Survey’s website:
www.ngu.no/geodynamics/gpmdb. The palaeomagnetic literature is said to be bursting with inconsistencies
(Storetvedt, 1997, p. 79). According to Tauxe (2013, 16.2): ‘Picking out the meaningful poles from the
published data is part of the art of paleomagnetism.’ Like other art forms, it appears to involve a high degree
of subjectivity.
Based on a study of pole determinations classed as ‘reliable’, Rezanov (1968, p. 772) stated: ‘the larger the
number of palaeomagnetic measurements for any given region, the broader becomes the scattering of the
palaeomagnetic poles as determined from rocks of some particular epoch in one and the same continent,
even in one and the same district.’ He found that the minimum pole scatter from any geologic period was
5000 to 6000 km, and that for geologic periods older than the Carboniferous the scatter was 10,000 km.
Palaeomagnetic data sometimes imply that sizeable landmasses were in the same place at the same time –
e.g. Azerbaijan and Japan during the mid-Cretaceous. He concluded that ‘paleomagnetic data are still so
unreliable and contradictory that they cannot be used as evidence either for or against the hypothesis of the
relative drift of the continents or their parts’ (p. 775).
Rezanov pointed out that the 6000-9000 km discrepancies between the Ordovician palaeomagnetic poles of
the Siberian platform had been attributed to remagnetization. He asks: ‘If the 90º divergence is believed to
be due to an error in the case of Siberia, why is it then that, in the case of Australia, the same divergence
must be taken as evidence of the colossal relocation of the continent at the beginning of the Carboniferous
and again in the Cenozoic?’ (p. 773). N.A. Khramov attributed the differences between the Ordovician poles
of Europe and Siberia to the relative drift of the Siberian platform or its parts with respect to the Russian
platform and the Urals. Rezanov pointed out that, to be consistent, the rotation and drift of parts of Siberia
should also be postulated for post-Permian time as well, but most of Siberia had already consolidated by the
Permian and to believe otherwise ‘means deliberately shutting one’s eyes to reality’. He urged that the
causes of palaeomagnetic inconsistencies be investigated before drawing conclusions about tremendous
horizontal shifts and rotations of continents and their parts.
Like Rezanov, Meyerhoff (1970a) highlighted the fact that if individual palaeomagnetic pole positions are
plotted for a particular age, rather than an averaged pole position, the scatter is typically so great, even from
single localities and geologic provinces, that the circles of error are wider than the Atlantic Ocean, even for
the Pleistocene and Holocene. Consequently, palaeomagnetic data cannot be used to prove either continental
drift or polar wander. Northrop & Meyerhoff (1963) pointed out that if – instead of producing an apparent
polar wander path by averaging all the palaeopoles from a particular continent for specific periods –
different APWPs are determined for different regions, the resulting curves can only be made to coincide by
moving and rotating different parts of geologically contiguous provinces, and the gyrations required are so
complex and erratic as to strain credulity.
Barron et al. (1978) held that palaeomagnetism is a ‘relatively inaccurate tool’ and ‘cannot be used for
making precise reconstructions’. They wrote: ‘the error of palaeomagnetic pole determination, especially for
individual formation results, could mean as much displacement as 1000-1600 km, without even having
regard for the longitudinal indeterminacy’ (p. 437). Errors of up to 16° are common in individual
palaeomagnetic positions from the same continent for the same period of geologic time. The authors also say
that pole positions for the eastern and western parts of Iceland differ by 14.3°, and that strict interpretation of
these data would require ‘an unreasonable amount of rotation’. They add that there is no reason to believe
that these results are in any way atypical of palaeomagnetic results in general.
8
Fig 1.8. Permian palaeomagnetic pole positions plotted for the northern continents. Poles not meeting minimum criteria
of reliability are also shown. (Meyerhoff & Harding, 1971, fig. 1).
Fig 1.9. Palaeomagnetic pole positions plotted for Proterozoic through Triassic time from a single locality on a shield.
This is not an isolated case or ‘exception’. (Meyerhoff & Harding, 1971, fig. 3)
Tarling (1982a) stated that it is only too easy to be sceptical about the palaeomagnetic database. The
uncertainties are immense and require subjective evaluation. ‘Unfortunately,’ he says, ‘this can lead to the
syndrome in which scattered data are examined and the odd points that are near to previous observations are
accepted for publication and the remainder appear, if at all, in the inaccessible appendix of a thesis.’
9
Van der Voo (1990) proposed seven criteria for determining the reliability of palaeomagnetic data. The
highest score (Q=7) is awarded if a palaeopole satisfies all the following criteria: 1. well-determined rock
age; 2. sufficient number of samples; 3. adequate laboratory demagnetization; 4. age of magnetization is
constrained by field tests; 5. structural control and tectonic coherence with craton or block involved; 6.
presence of reversals; 7. palaeopoles should not fall on a younger part of the polar wander path. Van der Voo
pointed out that many of the seven criteria are frequently not satisfied; very few poles satisfy all seven
criteria; most authors use poles with Q>2; some Q=4 results are known to be based on remagnetizations; and
even the maximum score is absolutely no guarantee that a result is a better indication of the geomagnetic
field than poles with lower scores. He also stated that it would be inappropriate to simply reject data
according to the date of publication: ‘some results published in the 1960s are still very valid today and other
results from the most recent decade have already been shown to be inaccurate in, for instance, their age
determination, structural correction or determination of characteristic directions’ (p. 1).
A basic tenet of palaeomagnetic research is that ‘to be considered reliable indicators of the ancient field,
magnetizations from rocks of indistinguishable age should yield well-grouped directions’ (Cottrell &
Tarduno, 2000). There is inevitably a tendency to view inconsistent data or data that do not fit theoretical
preconceptions as being of poor quality, or the result of remagnetizations, inclination shallowing, tectonic
rotations/deformation, nondipole geomagnetic field components, and so on. Any or all of these problems
may well apply, but with so many degrees of freedom, there is great scope for selectivity and subjectivity.
Plate tectonics: a failed revolution
Changes in ancient magnetic polarity at irregular intervals are recorded in the surface rock record, and over some fifty
years, palaeomagnetic data have been used to create the geomagnetic time scale, to firmly document seafloor spreading,
to validate plate tectonics, and to reconstruct vanished supercontinents. (Torsvik et al., 2012, p. 326)
The idea of large-scale continental drift has been around for some 200 years, but the first detailed theory was
proposed by Alfred Wegener in 1912. He postulated that all the continents were once joined together in a
supercontinent called Pangaea, and then drifted apart. The continents, made of ‘soft’ brittle sial, supposedly
ploughed slowly through the denser ocean crust, made of ‘hard’ fluid sima, under the influence of
gravitational and rotational forces. His theory led to vigorous debate and met with widespread rejection (Le
Grand, 1998).
Interest in the hypothesis of continental drift was revived in the 1950s, with the rise of the new science of
palaeomagnetism, which seemed to provide strong support for continental drift and/or polar wander. In the
early 1960s new data from ocean exploration led to the idea of seafloor spreading. A few years later, these
and other concepts were synthesized into the model of plate tectonics, which by the early 1970s had become
the new orthodoxy. Le Grand (1998, p. 229) described it as a ‘rapidly-moving juggernaut’ that quickly
crushed most of the remaining pockets of resistance. Although plate tectonics remains the ruling paradigm
today, all its basic elements have been called into serious doubt (e.g. Meyerhoff et al., 1996a; Storetvedt,
1997; Pratt, 2000, 2001).
Age of the seafloor
According to the seafloor-spreading hypothesis, new oceanic lithosphere is generated at midocean ridges
(‘divergent plate boundaries’) by the upwelling of molten material from the mantle, and as the magma cools
it spreads away from the flanks of the ridges. The horizontally moving plates are said to plunge back into the
mantle at ocean trenches or ‘subduction zones’ (‘convergent plate boundaries’).
Palaeomagnetic data are interpreted to mean that at the end of the Permian, some 250 Myr ago, virtually all
the present continents were part of a vast supercontinent, Pangaea, which began to break up in the Jurassic,
about 180 Myr ago. As a result, virtually all the ocean floor is thought to be no older than this age (see fig.
2.5). Yet for many decades evidence has been accumulating that contradicts this fundamental claim.
10
Fig. 2.1. (a) to (d) Maps of continental reconstructions based on palaeomagnetic data. (e) Poles and APW paths for five
continents for the last 200 Myr, evaluated at 5-Myr intervals. (Tauxe, 2013, fig. 16.8)
There have been numerous finds in the Atlantic, Pacific, and Indian Oceans of rocks far older than 180 Ma,
many of them continental in nature (for an overview, see: Vasiliev & Yano, 2007; Vasiliev & Choi, 2008;
Yano et al., 2009; Yano et al., 2011). This evidence alone is sufficient to refute all the fanciful,
palaeomagnetism-based reconstructions of continents drifting thousands of miles around the Earth’s surface.
It is also fatal to earth-expansionist claims that the areas occupied by the present oceans did not exist 200
Myr ago, when the Earth was supposedly much smaller. Plate tectonicists occasionally make ad-hoc efforts
to explain away all such finds, e.g. as glacial erratics or ship’s ballast (Heezen et al., 1959), or as
‘nonspreading blocks’ left behind during rifting, caused by the spreading axis and related ‘transform faults’
jumping from place to place (e.g. Bonatti & Honnorez, 1971; Bonatti & Crane, 1982; Bonatti, 1990; Pilot et
al., 1998). On the whole, however, this abundant and well-documented evidence is simply ignored – a
damning indictment of the plate-tectonic establishment.
A recent geology textbook (Carlson et al., 2008), for example, simply repeats the mantra that the ocean
lithosphere is ‘very young’ and is silent about any evidence to the contrary. A modern palaeomagnetism
textbook (Tauxe, 2013) assures its readers that ‘the oldest sea floor is about 180 Ma’. But actual age
determinations contradict this. For instance, during legs 37 and 43 of the Deep Sea Drilling Project (DSDP)
Palaeozoic and Proterozoic igneous rocks were recovered in cores on the Mid-Atlantic Ridge and the
Bermuda Rise (Reynolds & Clay, 1977; Houghton et al., 1979). Yet not one of these occurrences of ancient
rocks was mentioned in the Cruise Site Reports or Cruise Synthesis Reports. Reynolds & Clay (1977),
reporting on a Proterozoic date (635±102 Ma) near the crest of the Mid-Atlantic Ridge, wrote that the age
had to be wrong because, on the basis of marine magnetic anomalies, the site could not contain rocks older
than about 10 Ma.
11
Fig. 2.2. Localities of continental and ‘anomalously’ old rocks in the Atlantic Ocean. For comparison, the localities are
superimposed on the theoretical age distribution of the ocean floor according to plate tectonics. (Yano et al., 2009, fig.
1)
Aumento & Loncarevic (1969) reported that 75% of 84 rock samples dredged from the Bald Mountain
region just west of the Mid-Atlantic Ridge crest at 45°N consisted of continental-type rocks, and described
12
this as a ‘remarkable phenomenon’ – so remarkable that they classified these rocks as ‘glacial erratics’ and
gave them no further consideration. The Bald Mountain locality has an estimated volume of 80 km³, so it is
hardly likely to have been rafted out to sea on an iceberg or dumped by a ship. It consists of granitic and
silicic metamorphic rocks ranging in age from 1690 to 1550 Ma, and is intruded by 785-Ma mafic rocks
(Wanless et al., 1968), whereas its predicted age is 10 Ma or less. Zircons with ages of 330 and 1600 Ma
were found in gabbros beneath the Mid-Atlantic Ridge near the Kane fracture zone (Pilot et al., 1998). St
Peter and Paul’s Rocks at the crest of the Mid-Atlantic Ridge just north of the equator have a predicted age
of 35 Ma, but Melson et al. (1972) found an 835-Ma peridotite stock associated with other rocks giving 350-,
450-, and 2000-Ma ages.
Fig. 2.3 Plot of rock age versus
distance from the crest of the East
Pacific Rise, based mainly on DSDP
cores from Legs 1-54 (1969-1980). Leg
9, on which seafloor spreading in the
Pacific is based, is shown.
Radiometrically dated Cenozoic and
Cretaceous basalt dredges (Budinger &
Enbysk, 1967; Ozima et al., 1968) have
been added to show that the alleged
linear relation between rock age and
distance from the ridge crest is illusory.
Note that the Shipboard Scientific
Party omitted site 78, which does not
fit plate-tectonic theory. (Meyerhoff et
al., 1992a, fig. 25)
The spatial distribution of shallow-water sediments in the present oceans and their vertical arrangement in
some of the drilled sections are inconsistent with seafloor spreading (Ruditch, 1990; Orlenok, 1986). The
present oceans have undergone large-amplitude subsidence since the Jurassic, but this occurred mosaically
rather than showing a systematic relationship with distance from the ocean ridges. Younger shallow-water
sediments are often located farther from the axial zones of the ridges than older ones, and some areas of the
oceans appear to have undergone alternating subsidence and elevation.
A major effort should be made to drill the ocean floor to much greater depths to see whether there are more
ancient sediments beneath the basalt layer that is currently labelled ‘basement’. That older sediments are
likely to be found is shown by the fact that some basalts have baked contacts with the overlying sediments,
have chilled margins, alternate with sediments, or show other characteristics indicative of intrusives (dykes
and sills), or extrusives on the seafloor (e.g. pillow structure) (Meyerhoff et al., 1992a; Choi, 2001). The
basalts appear to be magma floods which cover the real ‘oceanic’ basement underneath. This was clearly
shown at drill site 10 on the Mid-Atlantic Ridge, where the lowermost sediments are Cretaceous (about 80
Ma) and the underlying basaltic sill, erroneously termed ‘basement’, had a fission-track age of only 15.9 Ma
(Macdougall, 1971).
Ocean-floor sampling and drilling, seismic data, palaeocurrent and sediment-provenance data, and ocean-
13
bed flora and fauna indicate that there used to be large (now submerged) continental landmasses in the
present oceans (Dickins et al., 1992; Dickins, 1994b; Choi et al., 1992; Choi, 1999, 2001). Many islands and
ocean plateaus with semi-continental crust appear to be the remnants of larger palaeolands, whose former
continental crust has undergone varying degrees of ‘oceanization’.
Fig. 2.4. Former land areas in the present Pacific and Indian Oceans. Only those areas for which substantial evidence
already exists are shown. Their exact outlines and full extent are as yet unknown. G1 – Seychelles area; G2 – Great
Oyashio Palaeoland; G3 – Obruchev Rise; G4 – Lemuria; S1 – area of Ontong-Java Plateau, Magellan Seamounts, and
Mid-Pacific Mountains; S2 – Northeast Pacific; S3 – Southeast Pacific including Chatham Rise and Campbell Plateau;
S4 – Southwest Pacific; S5 – area including South Tasman Rise; S6 – East Tasman Rise and Lord Howe Rise; S7 –
Northeast Indian Ocean; S8 – Northwest Indian Ocean. (Dickins, 1994b, fig. 1)
If the seafloor-spreading hypothesis is wrong, so is the subduction hypothesis, as a wealth of evidence
suggests (see Meyerhoff et al., 1996a; Storetvedt, 1997; Pratt, 2000, 2001; Oard, 2000b; Choi, 2000; Smoot,
1997). An alternative view of Benioff zones is that they are very ancient contraction fractures produced by
the cooling of the Earth.
Marine magnetic anomalies
The plate-tectonic belief in a geologically young ocean crust and vast continental displacements is
supposedly supported by marine magnetic anomalies – alternating bands of slightly higher and lower
magnetic intensity on either side of ocean ridges, which are widely believed to be produced by seafloor
spreading in combination with global magnetic reversals. However, linear magnetic anomalies are found on
only 70% of seismically active midocean ridges, and the diagrams of symmetrical, parallel, linear bands of
anomalies displayed in many plate-tectonic publications bear little resemblance to reality. The anomalies are
symmetrical to the ridge axis in less than 50% of the ridge system where they are present, and in about 21%
of it they are oblique to the ridge trend. Linear anomalies are sometimes present where a ridge system is
completely absent, and not all the charted anomalies are formed of oceanic crustal materials (Meyerhoff &
Meyerhoff, 1974b; Grant, 1980; Choi et al., 1990, 1992).
The initial, highly simplistic seafloor-spreading model for the origin of ocean magnetic anomalies has been
disproved by ocean drilling (Hall & Ryall, 1977; Hall & Robinson, 1979; Pratsch, 1986; Storetvedt, 2010).
The anomalies are not produced in the upper 500 m of oceanic crust, as originally thought. And magnetic
intensities, general polarization directions, and the frequent existence of different polarity zones at different
14
depths suggest that the source of magnetic anomalies lies in deeper levels of ocean crust not yet drilled or
dated. The virtual absence of predicted sheeted dyke complexes along midocean ridges, and the abundance
of continental and dynamo-metamorphic rocks in oceanic settings argue strongly against both seafloor
spreading and earth expansion.
Fig. 2.5. Seafloor age based on marine magnetic anomalies. (Müller et al., 2008; www.ngdc.noaa.gov)
As shown above, there are numerous instances where the theoretical ages of magnetic anomalies and the
seafloor are contradicted by actually measured rock ages. Magnetic-anomaly bands strike into the continents
in 16 places and ‘dive’ beneath Proterozoic or younger rocks; they are also approximately concentric around
Archaean continental shields. This suggests that they are the sites of ancient fractures, which partly formed
during the Proterozoic and have been reactivated since (Meyerhoff & Meyerhoff, 1974b).
Fig. 2.6. Subjective correlations
between two deep-tow profiles
over Jurassic Quiet Zone
magnetic lineations in the western
Pacific. (Sager et al., 1998, fig 5;
reprinted with permission from
the American Geophysical Union)
Correlations between magnetic stripes on either side of a ridge or in different parts of the ocean have been
largely qualitative and subjective, and are therefore highly suspect; virtually no effort has been made to test
them quantitatively by transforming them to the pole (i.e. recalculating each magnetic profile to a common
latitude). The magnetic anomalies of the Reykjanes Ridge are supposed to be a classic example of ridgeparallel symmetry, but Agocs et al. (1992) concluded from a detailed, quantitative study that the correlations
were very poor; the correlation coefficient along strike averaged 0.31 and that across the ridge 0.17, with
limits of +1 to -1. Correlations between the anomalies and bottom topography, on the other hand, averaged
0.42. The magnetic anomalies are better explained by fault-related bands of rock of different magnetic
susceptibilities (Agocs et al., 1992; Choi et al., 1992; Storetvedt, 2010). Reported magnetization values from
15
oceanic rocks are more than adequate to produce the observed anomalies (Luyendyk & Melson, 1967;
Opdyke & Hekinian, 1967).
Side-scanning radar images show that the midocean ridges are cut by thousands of long, linear, ridge-parallel
fissures, fractures, and faults. This strongly suggests that the ridges are underlain at shallow depth by
interconnected magma channels, in which semi-fluid lava moves horizontally and parallel with the ridges
rather than at right-angles to them, as claimed by plate tectonics (Meyerhoff et al., 1992a,b). Magnetic
anomalies are associated with these ridge-parallel fractures. This explanation eliminates the need to postulate
scores of ‘spreading centres’ of different ages in different places and oriented in different directions. For
example, Larson & Chase (1972) speculated that the various anomaly patterns in the western Pacific were
generated by a system of five spreading centres joined at two triple points. They stated that all correlations
among the magnetic profiles were established ‘by eye’. In contrast, if the anomalies were produced by onceactive magma channels, as proposed by surge tectonics, a coherent and internally consistent flow pattern
emerges (fig. 2.7).
Fig. 2.7. The various magnetic lineation sets in the Pacific basin could have originated above different surge channels
active at different times in the past. (Meyerhoff et al., 1992b, fig. 21)
Plate motion
According to plate tectonics, the Earth’s outer shell, or lithosphere (the crust and uppermost mantle), is
divided into 13 major plates, ranging in size from about 400 by 2500 km to 10,000 by 10,000 km. Over time,
hundreds of microplates and ‘exotic’ terranes have been added, often to accommodate discrepant
palaeomagnetic poles. The strong lithosphere is assumed to move in a relatively rigid way over a continuous,
weaker, and hotter (low-velocity) asthenosphere.
16
The lithosphere is believed to average about 70 km thick beneath oceans, and to be 125 to 250 km thick
beneath continents. The asthenosphere is said to extend to a depth of 200 km beneath oceans, while its
thickness, depth and very existence under continents are vigorously debated (Carlson et al., 2008). The
continental lithosphere was originally thought to be no
thicker than about 150 km. But evidence from seismicvelocity, heat-flow, and gravity studies shows that some
parts of the continents – especially ancient, stable
continental shields, e.g. the Fennoscandian, Siberian,
Canadian, Australian, and Antarctic shields and the
Amazonian and northwest African cratons – have very
deep roots and that the asthenosphere is very thin or
absent beneath them (e.g. MacDonald, 1963; Jordan,
1975, 1978, 1988; Pollack & Chapman, 1977;
Pavlenkova & Pavlenkova, 2006; Artemieva & Mooney,
2002). Seismic tomography, which produces threedimensional images of the Earth’s interior, confirms that
Fig. 2.8. Estimated lithospheric thickness of the
the continental lithosphere extends in places to depths of
Canadian Shield. (Shapiro et al., 2004, fig. 15)
up to 400 km or more (Legendre et al., 2012; O’Reilly et
al., 2009, Begg et al., 2009; Kustowski et al., 2008; Priestley & McKenzie, 2006; Zhou et al., 2006; Conrad
& Lithgow-Bertelloni, 2006; Shapiro et al., 2004; Hirth et al., 2000; Rudnick et al., 1998; Masters et al.,
1996). The higher seismic velocities of the continental roots (or keels) are attributed to conductive cooling
and chemical depletion.
Fig. 2.9. Seismotomographic model images (Vs models) for four depth slices beneath Africa. Hot (red-white) colours
indicate higher velocities and cool colours lower velocities. High-velocity domains beneath cratonic crust in Africa
extend to depths of 300-400 km. (O’Reilly, 2009, fig. 3)
This issue remains controversial, as different seismotomographic methods and models give different results.
17
Some plate tectonicists would like to restrict the maximum thickness of the lithosphere to 200-250 km
(Gung et al., 2003; Van Summeren et al., 2012; Yuan & Romanowicz, 2010). Gung et al. (2003) note that
continental roots determined from horizontally polarized shear waves (SH) are 100 km or more deeper than
those determined from vertically polarized shear waves (SV) (fig. 2.10). They attribute this to significant
radial anisotropy under most cratons at depths of 250-400 km, with horizontally polarized shear waves
travelling faster than vertically polarized waves; similar anisotropy is found under ocean basins at depths of
80-250 km. They propose that, in both cases, the anisotropy is related to shear in a low-viscosity, global
asthenosphere. They argue that the ‘Lehmann discontinuity’, observed mostly under continents at about 200250 km, and the ‘Gutenberg discontinuity’, observed under oceans at about 60-80 km, may mark the
lithosphere-asthenosphere transition. This interpretation is challenged by O’Reilly et al. (2009), who say that
‘the high S-wave velocities observed in the deep continental roots are unlikely to be explained simply by
anisotropy’, and that there is no clear evidence that the Lehmann discontinuity represents the boundary
between the cratonic lithosphere and the convecting mantle.
Fig. 2.10. Seismotomographic cross-sections
through three continents (see locations at top)
derived from SH waves (left) and SV waves
(right). The SH sections consistently indicate fast
velocities extending to depths in excess of 220
km, whereas the SV sections do not. (Gung et al.,
2003, fig. 4)
Geophysical data show that, far from the asthenosphere being a continuous layer, it is made up of
disconnected lenses, which are observed only in regions of tectonic activation and high heat flow. Seismic
research reveals a complicated zoning of the upper mantle, with alternating layers of higher and lower
velocities and different strengths. This stratification is particularly obvious in tectonically active regions, but
also exists under old platforms. Individual low-velocity layers, which might be associated with the
asthenosphere, are bedded at different depths in different regions and do not compose a single, continuous
layer. Furthermore, there are close correlations between near-surface geological features, crustal structure,
heat flow, geoid anomalies, and inhomogeneities in both the upper and lower mantle. The fact that such
connections remain stable for long periods of geologic time contradicts the idea of considerable horizontal
displacements of lithospheric plates in relation to deeper mantle structures (Pavlenkova, 1990, 1995, 1996).
Although averaged surface-wave observations suggested that the asthenosphere was universally present
beneath the oceans, detailed seismic studies indicate that here, too, there are only asthenospheric lenses.
Several low-velocity zones occur in the oceanic mantle, but it is difficult to establish any regularity between
their depth and their distance from the midocean ridge. Moreover, the mantle structure was found to be
asymmetrical in relation to the midoceanic ridges and composed of blocks, thereby contradicting the
18
seafloor-spreading process (Pavlenkova, 1990, 1996).
Fig. 2.11. Tomographic model images at three depth slices
for the Atlantic Ocean; the numbered locations are oceanic
island basalt provinces (O’Reilly et al., 2009, fig. 4). Hot
(red-white) colours indicate higher velocities and cool
colours lower velocities. Some high-velocity regions are
continuous with continental regions and some occur as
discrete ‘blobs’ scattered throughout the basin. (Similar
high-velocity regions are irregularly and densely distributed
in the Pacific Ocean.) O’Reilly et al. suggest that ‘these
high-velocity volumes represent remnants of depleted
(buoyant) ancient continental lithosphere, fragmented and
stranded during the rifting process at the opening of the
ocean basin’ – a theory-driven interpretation.
Fig. 2.12. Seismic profile across the old East European platform, the Urals, the young West Siberian platform, and the
Siberian craton, derived from peacetime nuclear explosions. Legend: 1 – isovelocity line (km/s); 2 – seismic boundary
sites from which the high-amplitude reflections were obtained; 3 – low-velocity layer; 4 – high-velocity blocks; 5 –
high-reflectivity zone. The profile shows that the asthenosphere, as a layer of partial melting, does not exist beneath
Northern Eurasia. (Pavlenkova, 2012, fig. 6)
19
Just as the base of lithospheric ‘plates’ is often ill-defined, some ‘plate’ boundaries are also ill-defined or
even nonexistent, e.g. the northwest Pacific boundary of the Pacific, North American, and Eurasian plates,
the boundary between the North and South American plates and Caribbean plate, and the boundary between
the South American, Antarctic and Scotian plates (Oard, 2000a). And whereas plate boundaries were
initially considered to be fairly narrow, their width is now believed to range from a few hundred metres to
thousands of kilometres; ‘diffuse plate boundaries’ are said to cover about 15% of the Earth’s surface
(Gordon & Stein, 1992).
The Earth’s crust is in constant motion. The Earth’s relief currently ranges from 8.8 km above sea level to
10.8 km below it. There is ample evidence that mantle heat flow and material transport can cause significant
changes in crustal thickness, composition, and density, resulting in substantial uplifts and subsidences –
without the need for ‘plate collisions’ and ‘subduction’ (Pratt, 2000). The scale of vertical movements is
indicated by the fact that the thickness of marine sedimentary layers in mountain belts is commonly over 10
km and can reach 23 km (Bucher, 1933). As far as horizontal movements are concerned, field evidence
indicates that crustal strata can be thrust tens if not hundreds of kilometres, that crustal extension or
shortening of up to hundreds of kilometres has occurred, and that motion of over a hundred kilometres has
taken place along some wrench faults. But given the widely varying thickness of the lithosphere, the
existence of deep continental roots, the lack of a continuous asthenosphere, the absence of some ‘plate’
boundaries, and the correlation between near-surface features and deep mantle features, the movement of
lithospheric slabs as relatively rigid bodies over hundreds or thousands of kilometres seems utterly
impossible. Nevertheless, plate tectonicists have convinced themselves that their models of plate motions are
supported by palaeomagnetic data and marine magnetic anomalies, and also by space-geodetic data – despite
some significant discrepancies with plate-tectonic models (Pratt, 2001).
Plate-tectonic reconstructions
Today, geologists and geophysicists tend to treat pieces of the Earth’s crust like a roomful of furniture, objects that can
be pushed around at will into whatever configuration is required to satisfy a particular model. Unfortunately, the Earth’s
crust is not so easily manipulated, particularly if one is faithful to physical laws as well as the geologic data. Instead,
the rigid crust and uppermost mantle form a massive interlocking mosaic, the lithosphere. ...
We know from geological field mapping that objects within the lithosphere mosaic are moved substantial distances,
both vertically and laterally. However, the argument that large lithosphere plates, each 50 to 200 km thick, each
extending for thousands of kilometers in all directions, and each weighing incalculable tons, can be moved freely and
systematically about the earth’s surface defies all physical laws and common sense. (Meyerhoff et al., 1996a. p. 1-2)
Supercontinents
Palaeomagnetic data in particular have led plate tectonicists to believe that during the Earth’s history there
have been at least three supercontinents, consisting of most of the present continental blocks. They have
been named Columbia, Rodinia, and Pangaea (Meert, 2012). Reconstructions of all of them vary
significantly.
The final assembly of Columbia allegedly took place at about 1800 Ma, and it broke up between about 1600
and 1200 Ma (Zhao et al., 2004; Hou et al., 2008; Rogers & Santosh, 2002). The various blocks separated
and rotated, and reassembled to form the supercontinent Rodinia between about 1300 and 900 Ma; it began
to break up at about 750 Ma, and during the process some of the pieces joined together to form
Gondwanaland (Li et al., 2008; Evans, 2009; Dalziel et al., 2000).
20
Fig. 3.1. Reconstructions of Columbia. Left: Rogers & Santosh (2002). Right: Zhao et al. (2004, fig. 17).
Fig. 3.2. Reconstruction of Columbia by Hou et al. (2008, fig. 7).
The Pangaea supercontinent allegedly formed at 300-250 Ma. According to Torsvik et al. (2012, p. 340),
‘Pangaea is the only supercontinent in Earth’s history that can be modelled with some, if any, confidence’.
In plate-tectonic mythology, the formation of Pangaea took place as follows. In the mid-Silurian (430-420
Ma), Laurentia – composed of cratonic North America, Greenland, Ellesmere, and parts of present-day
Europe (e.g. Scotland, NW Ireland, and Svalbard) – collided with Baltica (interior portion of northeastern
Europe) and Avalonia (composed of what is now southwest Great Britain and the eastern coast of North
America), producing the Caledonian orogeny. Laurentia thereby became the western portion of Laurussia. In
the late Cambrian and early Ordovician, Gondwana stretched from the south pole (northern Africa) to the
equator (Australia); it was almost 100 million km2 in size and covered about 20% of the Earth’s surface. The
bulk of Pangaea was formed in the late Carboniferous when Gondwana, Laurussia and intervening terranes
21
collided, producing the Hercynian orogenic belt in Western Europe. By
late Permian time, Siberia had joined Baltica, and along with other
European and Asian elements, the combined northern half of Pangaea is
referred to as Laurasia. The first major breakup occurred in the early
Jurassic with the opening of the Central Atlantic and the separation of
Gondwana and Laurasia.
Fig. 3.3. Reconstruction of Rodinia by
Dalziel et al. (2000). G = Greenland;
RP = Rio de la Plata; SF = Sao
Francisco; WAF = West Africa.
Fig. 3.4. The assembly of Rodinia, reconstructed by Li et al. (2008, fig. 9).
22
Fig. 3.5. Reconstruction
of Rodinia by Evans
(2009, fig. 11): (a) soon
after assembly (1070
Ma); (b) shortly prior to
breakup (780 Ma),
showing incipient
breakup rift margins
(red) and transform
offsets (black).
Fig. 3.6. Laurussia (also known as Euramerica or
the Old Red Continent) in the Devonian
(http://en.wikipedia.org). Plate boundaries are
shown in red. Note how, when Laurentia
supposedly collided with Baltica and Avalonia
(~430 Ma), a simultaneous perfect match was
produced between the northern and southern
halves of present-day Britain and Ireland!
Fig. 3.7. Gondwana and peri-Gondwana terranes in the
Lower Ordovician (480 Ma), based on palaeomagnetic data
from dark green areas. ATA = Amorican Terrane
Assemblage; MBL = Marie Bird Land; FI = Falkland
Islands; DML= Dronning Maud Land; MAD =
Madagascar. (Torsvik et al., 2012, fig. 10)
23
Fig. 3.8. APW spline path for Siberia (with error ellipses),
compared with the spline path for Baltica, showing their
supposed collision. (Torsvik et al., 2012, fig. 12)
Fig. 3.9. Plate-tectonic reconstructions of Palaeozoic and early Mesozoic geography. (Torsvik et al., 2012, figs. 17-19)
24
Drift rates are said to have been rarely higher than 10 cm/yr, and rotation rates were usually lower than
4º/Myr (Torsvik et al., 2012). For instance, during the Cambrian, Central Africa allegedly drifted southward
at rates as high as 10 cm/yr or more and underwent anticlockwise rotations as high as 2°/Myr. Baltica rotated
anticlockwise (1-2°/Myr) from Cambrian to early Devonian times (~160° in total from 500 to 400 Ma).
From late Devonian times, Gondwana drifted northward, accompanied by large clockwise rotations that
peaked at ~360 Ma (>4°/Myr).
According to palaeomagnetic data (Li et al., 1990), Australia rapidly rotated 130º clockwise between the
early Ordovician and mid-Silurian, then 30º anticlockwise between the mid-Silurian and early Devonian,
then a further 30º anticlockwise from early to mid-Devonian, and a further 15º anticlockwise to the late
Devonian. During this period Australia remained at low to equatorial latitudes, and this was succeeded by a
rapid southward movement during mid-Carboniferous times. Laing (1998) concluded that this scenario is so
geologically improbable that palaeomagnetic data are worthless.
Pangaea allegedly covered an area of 160 million km2 at 250 Ma, or about 30% of the Earth’s surface.
Although Pangaea means ‘all land’, plate tectonicists believe that it did not include all continental crust. For
instance, the North and South China blocks were never part of it, and during the early Permian phase of
Pangaea assembly, the Neotethys opened and Cimmerian terranes such as Lut, Helmand, Qiangtang (North
Tibet) and Sibumasu drifted away from the northeast Gondwana margin.
The APW paths published for the combined Gondwana continents differ widely and depend critically on the
authors’ data selection/rejection criteria, resulting in a large variety of shapes and loops (fig. 3.10)
Fig. 3.10. Various Palaeozoic apparent polar wander paths for Gondwana (south poles) (Van der Voo, 1993, fig. 5.15;
reprinted with permission from Cambridge University Press). T = Tertiary, K = Cretaceous, J = Jurassic, Tr = Triassic,
P = Permian, C = Carboniferous, D = Devonian, S = Silurian, O = Ordovician, Є = Cambrian, l = lower, m = middle, u
= upper. (a) Morel & Irving (1978), with a path X and a more complicated path Y. (b) Bachtadse & Briden (1990). (c)
Schmidt et al. (1990). (d) Kent & Van der Voo (1990).
25
Fig. 3.11. Cambrian through early Carboniferous individual palaeopoles (north poles) from each of the Gondwana
continents rotated into West African coordinates, combined with a best estimate of the apparent polar wander path.
(Van der Voo, 1993, fig. 5.14; reprinted with permission from Cambridge University Press)
According to the palaeomagnetically-based reconstruction shown in fig. 3.12, Gondwanaland supposedly
collided with North America in the early Devonian, causing the Acadian orogeny from present-day New
York to Newfoundland and transferring the Avalon terrane to North America (where it is now part of the
Appalachians). Gondwanaland then moved southward again, and finally redocked with North America in the
Carboniferous/Permian, before rifting away again during the breakup of Pangaea. Scotese & Barrett (1990)
accept the rapid northward motion of Gondwana during the latest Ordovician-early Silurian but not the
subsequent southward motion in the Devonian.
Fig. 3.12. (a) The Palaeozoic apparent
polar wander path for Gondwana
contains a loop in Silurian to early
Devonian time; ‘traditional’
interpolation of the Silurian to early
Devonian portion of the APWP is
shown by the dashed line; the
palaeomagnetic south poles are plotted
on the present geographic grid fixed to
Africa. (b) Ordovician
palaeogeography of Gondwana and
North America. (c) Early Devonian
palaeogeography. (d) Late Devonian
palaeogeography. (Butler, 2004, fig.
10.11)
26
Distribution of land and sea
It is interesting to compare the highly lopsided arrangement of continents and oceans during the periods of
the above hypothetical supercontinents with the notable regularities that characterize the distribution of land
and sea today. First, the continents tend to be triangular, with their pointed ends to the south. Second, the
northern polar ocean is almost entirely ringed by land, from which three continents project southward, while
the continental landmass at the south pole is surrounded by water, with three oceans projecting northward.
Third, the oceans and continents are arranged antipodally – i.e. if there is land in one area of the globe, there
tends to be water in the corresponding area on the opposite side of the globe (Gregory, 1899; Bucher, 1933;
Steers, 1950). The Arctic Ocean is antipodal to Antarctica; North America is antipodal to the Indian Ocean;
Europe and Africa are antipodal to the central area of the Pacific Ocean; Australia is antipodal to the small
basin of the North Atlantic; and the South Atlantic corresponds – though less exactly – to the eastern half of
Asia. Only 6% of the Earth’s surface does not obey the antipodal rule. Harrison et al. (1968) calculated that
there is a 1 in 14 probability that this arrangement is the result of chance.
The antipodal arrangement of land and sea is reminiscent of a tetrahedron – a regular polyhedron in which a
face is always opposite an apex (Umbgrove, 1947; Bucher, 1933). If an imaginary tetrahedron is placed
within the Earth with one corner in Antarctica, at the south pole, the other three will lie in three vast blocks
of very ancient, Archaean rocks in the northern hemisphere: the Canadian shield, the Fennoscandian shield,
and the Siberian shield, and the three edges running down to the south pole correspond to the three roughly
meridional lines running through the three pairs of continents: North and South America, Europe and Africa,
Asia and Australia. As Umbgrove pointed out, the antipodal, axisymmetric, and ‘tetrahedral’ arrangement of
land and sea is incompatible with both polar wander and continental drift – unless it is dismissed as pure
coincidence.
Gregory (1899) speculated that in the Upper Palaeozoic the tetrahedron might have been the other way up,
with one corner at the north pole. Instead of a continuous southern ocean belt separating triangular points of
land, there was then a southern land belt, supported by three great equidistant cornerstones: the Archaean
blocks of South America, South Africa, and Australia. Meyerhoff (1995, p. 172), on the other hand, held that
the antipodal arrangement of oceans and continents indicates that the Earth’s continents must have reached
their present positions in a molten Earth very early in its history, and is most likely a response to the Earth’s
rotation, with the surface masses being distributed more or less axisymmetrically for rotational stability. As
more data accumulate, it will be interesting to see whether a roughly antipodal distribution of land and sea
also applied in the geologic past when parts of the present continents were submerged and landmasses
existed in parts of the present oceans.
Fig. 3.13. Antipodal arrangement of land and sea. Overlaps in the northern hemisphere are outlined.
(http://nwhyte.livejournal.com)
27
Fig. 3.14. Map of seismic-wave vertical travel
time delays to a depth of 250 km computed
from the CUB2.0 global tomographic model.
Blue colours correspond to fast regions
(ancient cratons) and red colours to slow
regions. (Poupinet & Shapiro, 2009, fig. 2)
Palaeomagnetism versus geology
Problems with Pangaea
Fig. 4.1 shows four different reconstructions of Pangaea. Pangaea A (or A-1) is the computer-generated
Bullard et al. (1965) fit (based on the 500-fathom depth contour), as extended by Smith & Hallam (1970). It
is widely accepted as the starting point for the opening of the Atlantic in the Jurassic and is supported by a
good match of European and North American palaeopoles. However, there are disagreements among
Carboniferous, Permian and Triassic palaeopoles from Laurussia and Gondwana when the continents are
assembled in this fashion (Kent & May, 1987). To minimize the disagreements, alternative fits rotate
Gondwana clockwise relative to the northern continents, by amounts ranging from about 20º in the Pangaea
A-2 fit (Van der Voo & French, 1974) to 35º in the Pangaea B fit (Irving, 1977; Morel & Irving, 1981), and
even greater in the Pangaea C fit (Smith et al., 1981).
In the Pangaea A-2 fit, northwestern South America is fitted tightly into the Gulf of Mexico, leaving no
space for northern Mexico and its neighbouring continental blocks (Yucatán, Cuba, etc.). Reconstructing
Pangaea according to the longitude constraints of an A-type model would result in ~1000 km of crustal
overlap between West Gondwana and Laurussia (Domeier et al., 2012).
The Pangaea B fit takes advantage of the non-uniqueness of longitude in palaeomagnetic reconstructions and
places northwestern South America adjacent to eastern North America. Although the (lower) Permian poles
from most continents agree with the Pangaea B fit, those from Europe are significantly offset (Tauxe, 2013,
16.5). Morel & Irving proposed that Pangaea B existed during the latest Carboniferous to early Permian, and
that during the late Permian and Triassic Gondwana rotated anticlockwise into the Pangaea A configuration.
This requires a ~3500 km dextral megashear between Gondwana and Laurussia, for which there is no
evidence (Domeier et al., 2012).
In Pangaea C, Gondwana is displaced further east relative to its position in Pangaea B, allowing Gondwana
to be nudged northward to conform to palaeomagnetic data without resulting in overlap between the
continents. Pangaea C faces the same problems as Pangaea B, but exacerbated by the greater offset between
Gondwana and Laurasia. If Pangaea C were to transform to Pangaea A in the Permian or Triassic, a ~6000
km megashear would be required (Domeier et al., 2012). Pangaea B and C reconstructions can be made to
agree perfectly with palaeomagnetic data by letting Gondwana move episodically with respect to Laurussia
during the interval of interest.
28
Fig. 4.1. Four reconstructions of Pangaea. (McElhinny & McFadden. 2000, fig. 7.10)
The preferred plate-tectonic scenario is an initial Carboniferous and Permian Pangaea A-2 configuration that
evolves into the Pangaea A configuration by the late Triassic. Domeier et al. (2012) argue that late
Palaeozoic-early Mesozoic palaeomagnetic data from Laurussia and Gondwana can be reconciled with an Atype Pangaea without invoking alternative reconstructions or nondipole fields, by using an updated
palaeomagnetic dataset, refined Euler parameters, and applying theoretical inclination shallowing
corrections.
Fig. 4.2. Palaeogeographic reconstructions by Domeier et al. (2012, fig. 23).
All the different Pangaea reconstructions are an exercise in futility: they ignore a mass of evidence
indicating that the crust underlying the present oceans is up to several billion years old, and contain many
other glaring flaws. In the Bullard et al. (1965) fit, for example, the whole of Central America and much of
southern Mexico are omitted, despite the fact that extensive areas of Palaeozoic and Precambrian continental
rocks occur there. This region of some 2,100,000 km² overlaps South America in a region consisting of a
29
craton at least 2 billion years old. The entire West Indian archipelago has also been omitted. In fact, much of
the Caribbean is underlain by ancient continental crust, and the total area involved, 300,000 km², overlaps
Africa; the overlap extends 1500 km in an east-west direction. The Cape Verde Islands-Senegal basin, too, is
underlain by ancient continental crust, creating an additional overlap of 800,000 km² (Meyerhoff &
Meyerhoff, 1974a; Meyerhoff & Hatten, 1974). The Central American/Caribbean overlap is even worse in
the A-2 reconstruction. Some post-Bullard models have creatively reconstructed Middle America by placing
the continental blocks Maya (Yucatán) and Chortis (Honduras-Nicaragua-Jamaica) in the Gulf of Mexico,
alongside southwest Mexico, where they rotated 135º and 180º respectively into today’s positions. But
geological data contradict this (James, 2012).
All the Pangaea reconstructions ignore several major submarine structures in the Atlantic that appear to be of
continental origin, including the Faeroe-Iceland-Greenland Ridge, Jan Mayen Ridge, Vøring Plateau, Walvis
Ridge, Rio Grande Rise, and the Falkland Plateau. However, the Rockall Plateau was included for the sole
reason that it could be ‘slotted in’. In the Smith & Hallam (1970) reconstruction of the Gondwanaland
continents, the South Orkneys and South Georgia are omitted, as is Kerguelen Island in the Indian Ocean,
and there is a large gap west of Australia. Fitting India against Australia, as in other fits, leaves a
corresponding gap in the western Indian Ocean (Hallam, 1976). As shown in figure 4.3, even the celebrated
fit of South America and Africa is poor and requires ad hoc adjustment of the African coastline.
Fig. 4.3. Mismatch between Africa (AFR) and South America (SAM)
(solid lines, as presently constituted). Van der Voo says that this mismatch
‘must be corrected by postulating plate boundaries within Africa, so that
Africa’s outline for pre-Cretaceous times is represented by the dashed
line’. FP = Falkland Plateau, AFZ = Aguilhas Fracture Zone. The Euler
rotation pole is shown. (Van der Voo, 1993, fig. 6.7; reprinted with
permission from Cambridge University Press)
Geological connections
The opening of the Atlantic Ocean allegedly began in the Jurassic by the rifting apart of the Eurasian and
American plates. However, on the other side of the globe, northeastern Eurasia is joined to North America
by the Bering-Chukotsk Shelf, which is underlain by Precambrian continental crust that is continuous from
Alaska to Siberia. Geologically these regions constitute a single unit, and it is unrealistic to suppose that they
were formerly divided by an ocean several thousand kilometres wide, which closed to compensate for the
opening of the Atlantic. If a suture is absent there, one ought to be found in Eurasia or North America, but
no such suture appears to exist (Shapiro, 1990). The geologic continuity between the Bering-Chukotsk Shelf
and the Lomonosov Ridge rules out seafloor spreading or continental drift in the Arctic Ocean since
Proterozoic time. The idea that Siberia collided with Europe along the line of the present Urals in the late
Palaeozoic is contradicted by abundant evidence demonstrating that the Siberian and East European
(Russian) platforms formed a single continent during Archaean to early Proterozoic time. Structures and
rock units of the East European Timan Range strike beneath the Urals, reappear on the eastern side, and are
present under the Mesozoic cover of the West Siberian platform (Meyerhoff & Meyerhoff, 1974a).
If Baffin Bay and the Labrador Sea had formed by Greenland and North America drifting apart, this would
30
have produced hundreds of kilometres of lateral offset across the Nares Strait between Greenland and
Ellesmere Island, but geological field studies reveal no such offset (Grant, 1980, 1992). Greenland is
separated from Europe west of Spitsbergen by only 50-75 km at the 1000-fathom depth contour, and it is
joined to Europe by the continental Faeroe-Iceland-Greenland Ridge. In fact, more than 60% of the watercovered areas between 62 and 82°N, in the North Atlantic Ocean, appears to be underlain by continental
crust (Meyerhoff, 1974; Meyerhoff & Meyerhoff, 1974a). All these facts rule out the possibility of east-west
drift in the northern hemisphere.
Plate models require a suture zone running the length of the Mediterranean Sea, despite stratigraphic
continuity between Europe and Africa. There has been a direct tectonic connection between Europe and
Africa across the zones of Gibraltar and Rif on the one hand, and Calabria and Sicily on the other, at least
since the end of the Palaeozoic, contradicting plate-tectonic claims of hundreds of kilometres of lateral
displacement between Europe and Africa during this period (Kent, 1969; King, 1971; Trümpy, 1971;
Beloussov, 1990). On the basis of palaeomagnetic data, plate tectonicists have claimed that the Bay of
Biscay formed by the Iberian Peninsula rotating anticlockwise by up to 40o, but this is contradicted by
geological and geophysical evidence (Kent, 1969; Jones & Ewing, 1969; Bacon & Gray, 1971; Maxwell,
1970).
Plate tectonicists hold widely varying opinions on the Middle East region. Some advocate the former
presence of two or more plates, some postulate several microplates (19, according to Reilinger et al., 2006),
others support island-arc interpretations, and a majority favour the existence of at least one suture zone (and
subduction zone) that marks the location of a continent-continent collision. Kashfi (1992, p. 119, 128)
comments:
Nearly all of these hypotheses are mutually exclusive. Most would cease to exist if the field data were honored. These
data show that there is nothing in the geologic record to support a past separation of Arabia-Africa from the remainder
of the Middle East. …
The Iranian plateau and southwestern Iran and Arabia have been a single geologic zone since the beginning of
Proterozoic time, as shown by the following: (1) the Late Proterozoic through Tertiary stratigraphic correlations and
continuity across Iran and the entire Middle East, from India to Yemen and Jordan; (2) the biozonal correlation from
the Middle East to central Asia; (3) the structural unity across the greater Persian Gulf area; (4) the PrecambrianCambrian salt connection along western India, Pakistan, Iran, Persian Gulf, and Arabia; (5) the existence of strong
seismicity away from the alleged subduction zone; and (6) the random distribution of ophiolites and volcanic rocks in
the Middle East.
The fact that thick evaporites and carbonates are widespread and laterally continuous in the Middle East
(including Iran, Iraq, and Arabia) shows that no suture zone requiring thousands of kilometres of tectonic
transport ever passed through this region in the Palaeozoic or Mesozoic (Meyerhoff et al., 1996b).
Geological and geophysical data simply point to horizontal (tangential) compression of the crust between the
Afro-Arabian block and southwestern Asia.
Southeast Asia
Plate tectonicists and palaeomagneticians clearly have a great fondness for chopping continents up and
moving the individual fragments around. This also applies to China, where the North China and South China
blocks (or microcontinents) were supposedly once far apart; one view is that their eastern ends had collided
by the late Permian, after which they rotated towards one another through an angle of 67º (Zhao & Coe,
1987).
By contrast, Meyerhoff et al. (1991), based on a detailed stratigraphic and palaeogeographic study,
concluded that cohesion has characterized not only China but also large regions of greater Asia through
more than three billion years of geologic time. The idea that Asia is a collage of previously dispersed
microplates or microcontinents that later collided is contradicted by a mass of detailed field data. Field
studies show that many ‘sutures’ are not collision zones but zones where mafic and ultramafic rocks rose to
the surface along fault systems produced in a tensile, rifting, tectonic environment (Meyerhoff, 1995).
31
Based on his surge-tectonic interpretation of Southeast Asia, Meyerhoff (1995) presented maps of surge flow
in Southeast Asia since late Precambrian time, showing that, due to the Earth’s rotation, the overall flow
direction is eastward, but in the western half of the region the surge channels thread their way around the
various massifs lying between the northern (Angaran) and southern (Gondwanan) platform regions, while in
the eastern half the channels fan out to the northeast and southeast (see figures 4.4 and 4.5). Most ancient
cratonal areas and also the Benioff zones in the western Pacific Basin appear to act as barriers to surge
channels, probably because they are rooted in mantle.
The whole pattern resembles one that would be produced by a fluid moving from left to right (west to east)
through a narrow opening into a larger container. The positions of the channels and channel complexes
shifted somewhat over time, but very little. At local level small channels came and went, but the gross
patterns remained unchanged. Meyerhoff (1995, p. 159) writes: ‘The anastomosing patterns of surge
channels and stable blocks (massifs and platforms) ... speak for themselves. No pattern dreamed up for
continental, or microcontinental, “collisions” can explain the simplicity of this pattern which, fundamentally
is a flow pattern.’ The stability of the main flow patterns in Southeast Asia since at least the late Proterozoic
contradicts plate-tectonic models for the region’s geological evolution and also casts doubt on large-scale
polar wander.
Fig. 4.4. Map of southeastern Asia showing platforms, massifs, principal gaps, North-South Zone, Hanoi-Da Hinggan
Ling gradient, and marine magnetic anomalies. The inset map shows the surge-tectonic concept of Asia’s fundamental
‘Y’ structure, splaying toward the east. (Meyerhoff et al., 1996a, fig. 5.2; reprinted with permission from Springer
Science+Business Media BV).
32
Fig. 4.5. Palaeotectonic (surge channel) map of Southeast Asia for late Jurassic-middle Eocene. (Meyerhoff et al.,
1996a, fig. 5.17; reprinted with permission from Springer Science+Business Media BV)
India and Tethys
There is overwhelming geological and palaeontological evidence that India has been an integral part of Asia
since at least mid-Proterozoic time (Chatterjee and Hotton, 1986; Ahmad, 1990; Saxena et al., 1985; Saxena
& Gupta, 1990; Meyerhoff et al., 1991). Yet on the basis of palaeomagnetic data and marine magnetic
anomalies, plate tectonicists claim that India detached itself from Antarctica sometime during the Mesozoic,
and then drifted northeastward up to 7500 km, at speeds of up to 18 cm/yr, until it finally collided with Asia
in the Eocene (55 Ma), pushing up the Himalayas and the Tibetan Plateau. That Asia happened to have an
indentation of approximately the correct shape and size and in exactly the right place for India to ‘dock’ into
would certainly be a remarkable coincidence (Mantura, 1972). Collision models generally assume that the
uplift of the Tibetan Plateau began during or after the early Eocene, whereas palaeontological,
palaeoclimatological, palaeoecological, and sedimentological data conclusively show that major uplift began
in the early Pliocene (5 Ma) and did not reach its present rate (5 mm/yr) until about 2 Ma (Meyerhoff, 1995).
The collision model fails to explain why the beds on either side of the supposed collision zone remain
comparatively undisturbed and low-dipping, whereas the Himalayas have been uplifted, supposedly as a
consequence, some 100 km away, along with the Kunlun Mountains to the north of the Tibetan Plateau.
River terraces in various parts of the Himalayas are almost perfectly horizontal and untilted, suggesting that
the Himalayas were uplifted vertically, rather than as the result of horizontal compression (Ahmad, 1990).
33
The collision zone is supposedly marked by ophiolites, but these are not continuous within the Indus ‘suture’
zone but form irregular outcrops along the Yarlung Zangbo Valley. This ophiolite zone actually parallels
two other zones farther north in Xizang (Tibet) – thereby contradicting conventional plate-tectonic models.
Field data indicate that maximum crustal shortening between central China and the Indian subcontinent does
not exceed 300 to 700 km (Saxena et al., 1985).
The trans-Asiatic lineaments shown in fig. 4.6, which appear to have originated in Precambrian time, and
some of which extend into the Indian Ocean, refute plate-tectonic theories about the long-distance migration
of India (Raiverman, 1992).
Fig. 4.6. Trans-Asiatic lineaments. Legend: 1. orographic axis; 2. ophiolites; 3. lineaments. (Raiverman, 1992, fig. 2)
If the long journey of India had actually occurred, it would have been an isolated island-continent for
millions of years – sufficient time to have evolved a highly distinct endemic fauna during the late Cretaceous
and early Tertiary. However, the Mesozoic and Tertiary faunas show no such endemism, but indicate instead
that India lay very close to Asia throughout this period, and not to Australia and Antarctica (Chatterjee &
Hotton, 1986; Meyerhoff et al., 1996b). 60% of Indian reptilian and amphibian faunas are generically
identical with forms known only from the northern hemisphere, while the remaining taxa are known from
both hemispheres.
Meyerhoff & Meyerhoff (1978) wrote that unless plate tectonicists ‘can provide an acceptable explanation
for the continuity of Precambrian-Cenozoic stratigraphy and biostratigraphy across all of southern Asia, and
for the intertonguing of Gondwana and Tethys biotas and formations in northern India, and even in parts of
the Soviet Union and Tibet, “Greater India’s northward flight from Antarctica and Australia” appears to be
no more than a flight of fancy’.
According to plate tectonics, a triangular Tethys Ocean once separated central and eastern Eurasia on the
north from Arabia, India and Australia on the south. This idea arose because Bullard et al. (1965), in their
efforts to make the continents fit across the Atlantic (using a least-squares criterion), were forced to rotate
Eurasia anticlockwise with respect to North America, and to rotate Africa-Arabia clockwise with respect to
South America. Based on available palaeontological data, Dietz & Holden (1970) placed only northeastern
Africa, the Arabian Peninsula, India, Madagascar, Antarctica, and Australia in Gondwanaland along the
southern shore of the Tethys. In contrast, Drewry et al. (1974) also included Turkey, the Middle East, Iraq,
34
and Iran in Gondwanaland because some Gondwanan biotas had been reported from isolated localities in
those regions. Many different scenarios have been proposed, in which landmasses of various sizes have
migrated northward over great distances, colliding with and accreting to Eurasia, from China to Western
Europe.
Meyerhoff et al. (1996b) showed that 30-50% of the area of
Laurasia contains Gondwanan biotas of various ages, and 5080% of Gondwanaland contains Laurasian biotas of various
ages, and at all times from the Cambrian to the early
Cretaceous a broad biological transition zone is present
between Gondwanaland and Laurasia, in which the biotas of
the two regions are interbedded and, in many cases,
intermingled in the same bed. This renders the concept of a
wide Tethys Ocean untenable. There is abundant evidence
that the Tethys Sea in the region of the present AlpineHimalayan orogenic belt was never a deep, wide ocean but
rather a narrow, predominantly shallow, intracontinental
seaway (Bhat, 1987; Dickins, 1987, 1994c; McKenzie, 1987;
Stöcklin, 1989; Brinkmann, 1972; Trümpy, 1971). All the
biogeographical evidence shows clearly that the strata
everywhere developed in shallow water (Meyerhoff, 1991).
Fig 4.7. The Tethys Sea. (Butler 2004, fig.
10.10)
Fig. 4.8. Cambrian palaeogeography of Tethys, western India to eastern Mediterranean. (Meyerhoff & Meyerhoff,
1974a, fig. 21; Wolfart, 1967; reprinted with permission from Urban-Verlag)
35
Fig. 4.9. Permian palaeogeography of Tethys, western India to eastern Mediterranean. (Meyerhoff & Meyerhoff, 1974a,
fig. 22; Wolfart, 1967; reprinted with permission from Urban-Verlag)
Meyerhoff & Meyerhoff (1974a) wrote:
The rock record shows that Tethys has been a geologic unit from the Atlantic west of Gibraltar to New GuineaAustralia since Proterozoic, and probably Archeozoic, time. North Africa’s rock succession is continuous with that of
Mediterranean Europe and is as closely tied eastward with Arabia. The sedimentary sequence in Arabia and Iran has
been correlated with the succession in the Indian subcontinent, and has been traced in central and eastern Asia north
and south of the Himalayan zone (Lower Himalayas). …
If temperature determines the nature of the fauna ... then Tethys is a classic example of an east-west-oriented zone
whose latitude has not changed significantly since early or middle Proterozoic time. (p. 87, 128)
Exotic terranes
Another dubious, palaeomagnetism-inspired concept is that of ‘allochthonous tectonostratigraphic terranes’,
also known as ‘exotic terranes’ or ‘suspect terranes’ (Frisch & Meschede, 2011). Fault-bounded terranes
range in size from small blocks, less than 100 km2, to fragments of a microcontinent that may be thousands
of square kilometres. They appear to be geologically distinct from neighbouring crustal blocks and, on the
basis of palaeomagnetic data, many are believed to have originated far from their present locations. They
were supposedly carried across ocean basins on moving plates, sometimes for thousands of kilometres, until
they underwent collision and became welded or accreted to another terrane or a larger continental block,
thereby being transferred from one plate to another. After undergoing accretion, they can supposedly slide
along the margin of the continent for several hundred or thousand kilometres. Ocean plateaus – which
usually have a semi-continental crust – are regarded as examples of such terranes that are still travelling over
the Earth’s surface and are too big to eventually disappear down a subduction zone. Island arcs, too, can
36
become accreted terranes.
Most mountain belts – including the Appalachians, Alps, and
Himalayas – allegedly consist largely of allochthonous terranes. The
Appalachian orogen is said to be composed of ‘an embarrassing
number and variety of outboard terranes’ (Williams & Hatcher,
1982). Over 70% of the North American Cordillera is thought to be
a mosaic of ‘suspect terranes’, most of which are believed to have
travelled from far reaches of the Pacific Ocean. The number of
terranes is often put at over 50 (Coney et al., 1980), but there is
disagreement about both the number and the distance they have
travelled (Carlson et al., 2008; Colpron et al., 2007). The circumPacific terrane map shows over 300 terranes in the Pacific
borderlands (Howell et al., 1983).
Fig. 4.10. Map showing alleged terrane
accretion to the western margin of the
North American craton.
(http://pubs.usgs.gov)
Some plate tectonicists have shown that instead of explaining
discordant palaeomagnetic results (i.e. those that do not match the
APW path for the continent in question) by large horizontal
movements, they can be explained equally well by vertical block
rotations, tilting, folding, and/or inclination shallowing (e.g. Butler
et al., 1989; Butler et al., 1991; Butler et al., 2002; Calderwood,
1991; Irving & Archibald, 1990; Hodych & Bijaksana, 1993). In
many cases, field data contradict the long-distance transport of the
terrane (e.g. Laubscher, 1975; Donovan & Meyerhoff, 1982;
Parnell, 1982; McDowell et al., 1984; Saul, 1986; Seiders, 1988;
Hansen, 1988; Newton, 1988).
Careful work by some geologists indicates that the tectonostratigraphic belts from the Sierra Nevada to the
Central and Southern Rocky Mountains of the Western Cordillera are essentially in situ and have not moved
long distances except possibly by thrusting. Some workers are returning to the concept of lithofacies belts or
nappes, a concept demonstrated in the European Alps over a century ago (Meyerhoff et al., 1996a). Sengör
(1990) summarized this old European concept and the reasons why the present allochthonous terrane
concept should be abandoned. As Dickinson (2003) puts it, ‘Preferential reliance on the paleomagnetic
signature of deformed rocks as a faithful record of paleolatitude has the hallmark of mythic thinking, and
seems unnecessarily restrictive. Evaluation of terrane motions can be made with confidence only by
adopting a multidisciplinary strategy.’
Wegener and continental drift
There is a prevalent myth nowadays that Alfred Wegener was a prophetic thinker whose theory of
continental drift was far ahead of its time, that he was thwarted by conservative, dogmatic geologists, and
that plate tectonics has since confirmed many of his insights. Hal Hellman (1998, p. 158), for example,
waxes lyrical: ‘When he [Wegener] died, in 1930, his theory was still in a sort of scientific limbo. His legacy
lives on, however – bigger, grander, more comprehensive, and more majestic than even he could have
imagined.’ Such a view is far removed from reality. Many of the arguments raised against Wegener’s theory
were perfectly valid and some are equally applicable to plate tectonics.
Charles Schuchert (1928, p. 111) stated: ‘Wegener has taken extraordinary liberties with the earth’s rigid
crust, making it pliable so as to stretch the Americas from north to south about 1,500 miles.’ Philip Lake
(1922, p. 344) stated that if, in addition to moving the continental masses, ‘we are also allowed to mould
them as we will, the coincidences that we deduce become evidence of imaginative powers, not of former
realities’.
37
Referring to the geological similarities on opposite sides of the Atlantic, J.W. Gregory (1929) wrote:
The resemblances are most conspicuous between Newfoundland and the southern parts of the British Isles, between the
chains of the Antilles and of the Mediterranean, and between Southern Africa and the opposite parts of South America.
These resemblances are no greater than are to be found along the mountain-chains of Eurasia at similar distances apart.
The differences in detail between Newfoundland and Ireland, between South America and South Africa, … and
similarity between Venezuela and the Atlas are so marked that the countries, although they underwent the same general
geographical vicissitudes, must have been far apart. The resemblances are due to the areas having belonged to the same
tectonic belt; but the differences are sufficient to show that the areas were situated in distant portions of the belt. (p.
116)
Gregory also wrote:
The Appalachian and Armorican Mountains may have belonged to a continuous mountain belt without their having
been actually adjacent, just as the Pyrenees and the Caucasus are regarded as part of one mountain system although
they have always been separated by the full width of Europe. In fact, the differences between the Appalachians and the
corresponding mountains in Western Europe indicate that they were probably formed some distance apart. The problem
is whether the corresponding structures on opposite sides of the Atlantic have been separated by a two-mile subsidence
of the intervening area, or by the horizontal drift of America for 2000 or 3000 miles. (1925, p. 256)
Lake (1922) regarded the fact that the Appalachian folds and the Armorican folds lay on a great circle as a
remarkable fact that suggested they once formed a continuous system, but with the still visible portions
being in their original positions.
Schuchert (1928), too, argued that the geological similarities of the Atlantic continents are far fewer than
would be expected if the drift hypothesis were correct. Brazil and western Africa, for example, have been
independent and far separated since at least the Silurian. While Wegener was correct in connecting the
Caledonian crustal trends of northwestern Europe with those of northern Newfoundland, he was wrong in
connecting them directly. Newfoundland was never a part of Ireland; each land belongs to a widely differing
geological province.
Longwell (1944) objected that Wegener’s emphasis on the widespread distribution of Glossopteris flora
(extinct seed ferns) in southern continents as ‘compelling’ evidence of drift ‘seems dangerously near the
unscientific procedure of selecting evidence to support a favored theory’. Lake (1922), Gregory (1925,
1929), Berry (1928), Simpson (1943) and other workers showed that continental drift created more problems
than it solved in explaining the distribution of flora and fauna. Moreover, palaeontological evidence shows
that land connections across the Atlantic were established and broken several times, so that instead of
drifting steadily westward, the Americas would have to move back and forth like a concertina.
Schuchert (1928) noted that if the drift hypothesis were correct, many fossil marine faunas should have
between 50 and 75% identical species rather than 5%. Closing the Atlantic creates a 600 mile gap between
Siberia and Alaska, which Schuchert described as fatal to Wegener’s hypothesis because fossil evidence
shows that the Bering Sea region has been either a shallow sea or a land bridge since the early Cambrian,
permitting migrations between the Asiatic and American sides of the Pacific and into the Arctic as well. To
explain the dispersion of flora and fauna through geologic time, Schuchert postulated the existence of
various land bridges in different periods (e.g. across the Bering Strait, between Brazil and northern Africa,
between South America and Antarctica, between Antarctica and Australia, and between the latter and
Borneo and Sumatra), together with dispersal along shelf seas, by wind and water currents, and by migratory
birds.
One of Wegener’s allies, Alexander du Toit (1937), responded to the argument that drift would require nearidentity of facies and faunas on opposite coasts by arguing that the shorelines of South America and Africa
were never closer than 400-800 km. While this avoids certain difficulties, it is also an admission that the
case for continental drift based on stratigraphic similarities is less convincing than some drifters would have
us believe (Longwell, 1944).
38
Wegener postulated not only westward drift of the continents but also major polar wander, amounting to 60o
since the Carboniferous and up to 15o even since the late Pliocene. Attempts by Köppen & Wegener (1924)
to explain past climates by means of continental drift and polar wander attracted severe criticism (e.g. Berry,
1927, 1928; Brooks, 1949). Brooks (1949) detailed numerous shortcomings in their theories and noted their
tendency to dismiss evidence that contradicts their claims. He remarked: ‘The drift hypothesis has certainly
not reached a stage of proof in which it can be asserted that evidence which does not fit it is thereby proved
to be false’ (p. 234).
If all the present continents were once joined together in a Pangaea supercontinent, as Wegener proposed,
many of the areas glaciated in the Carboniferous and early Permian would lie next to one another; some
scientists saw this as persuasive evidence of drift. But as pointed out by Coleman (1925, 1932), cold alone
will not produce an ice sheet; moist winds coming from a warm sea are required. In a Pangaea
reconstruction, some of the glaciated areas would be so far inland that they would be out of reach of
moisture-laden winds; that is why Siberia, though beside the Arctic Ocean, was not glaciated to any
significant extent in the Pleistocene.
Lake (1922, p. 338) stated that Wegener ‘is not seeking truth; he is advocating a cause, and is blind to every
fact and argument that tells against it’. Berry (1928) said that his main objection to the drift hypothesis was
that Wegener’s method was ‘not scientific, but takes the familiar course of an initial idea, a selective search
through the literature for corroborative evidence, ignoring most of the facts that are opposed to the idea, and
ending in a state of auto-intoxication in which the subjective idea comes to be considered as an objective
fact’.
True polar wander
Polar motion
The location of the Earth’s rotation poles relative to the crust is not absolutely fixed but subject to slight
variations. This polar motion has two main components: a free oscillation with a period of about 435 days
and a variable amplitude of about 0.1 to 0.2 arc-seconds, known as the Chandler wobble; and a forced annual
oscillation with a nearly constant amplitude of about 0.1 arc-seconds. The oscillations are caused mainly by
oceanic and atmospheric processes (www.iers.org; Gross, 2000). The poles trace spiral paths out of, around,
and then back to their mean positions over a period of about 6.5 years, the maximum separation between the
actual and mean poles during this period averaging about 0.25 arc-seconds (www.britannica.com). Since
1900 the mean pole has shown an irregular drift of about 107 mm/yr in the overall direction of 79ºW, i.e.
towards Hudson Bay (Besse et al., 2011; http://hpiers.obspm.fr). There is no certainty that the poles will
continue to migrate in the same direction for many millions of years.
Polar motion is a sign that the Earth’s rotation axis tends to align itself with the Earth’s axis of maximum
moment of inertia (or axis of figure), i.e. the axis of symmetry of the Earth spheroid, determined by the
distribution of mass within the planet. In general, the rotation/geographic poles will move in response to any
mass redistributions in or on the Earth, including weather systems, the seasonal displacement of air and
water masses, seismic activity, melting ice caps, vertical and horizontal crustal movements, and motions and
density changes within the mantle and core (Dickman, 2000). In plate-tectonic terms, this includes plate
motions, mantle convection, mantle plumes, and plate subduction. Mass redistributions can also cause shortterm fluctuations or long-term changes in the planet’s spin rate.
Some workers believe that three-quarters of the current secular drift of the rotation pole is attributable to
deglaciation at the end of the last ice age, and the rest to mantle convection (Cambiotti, 2012; Peltier & Wu,
1983). Gordon (1995) held that polar wander over the past 10-20 Myr might be connected with the uplift of
the Tibetan Plateau and other mountain ranges.
39
Fig. 5.1. Blue: Path of the geographic pole,
2008-2012. Green: Mean pole positions since
1900. (http://hpiers.obspm.fr)
Variants of true polar wander
Periodic polar motion and any secular drift of the poles are examples of true polar wander (TPW), which is
usually interpreted to mean the displacement of the entire Earth (crust, mantle and core) with respect to the
spin axis, resulting in a change in the position of the geographic poles and equator on the Earth’s surface,
while the tilt of the Earth’s axis remains the same. There are several other possible variants of TPW, in
which it is not the entire Earth that shifts, but only: 1. the crust; 2. the mantle; 3. the lithosphere (crust +
upper mantle); 4. the lithosphere and all or part of the sublithospheric mantle.1
The rapid sliding of the Earth’s entire crust at or above the Mohorovičić discontinuity, triggered by surface
erosion, was proposed by Gussow (1963). The motion of the sublithospheric mantle only, sandwiched
between the lithosphere and core, was proposed by Hargraves & Duncan (1973), who called it ‘mantle roll’.
Damian Kreichgauer (1902) and Charles Hapgood (1958, 1970) proposed that it was the entire lithosphere
that moved; this idea is fashionable among some popular, catastrophist writers (e.g. Hancock, 1995), who
tend to call it ‘crust displacement’. Hapgood believed that there had been three lithosphere displacements
during the past 100,000 years, with the last one taking place between 17,000 and 12,000 years ago when the
north pole allegedly moved 30° (3300 km) to its present location. All these three forms of TPW seem totally
impossible.
The general view among plate tectonicists is that TPW involves a reorientation of the entire planetary body
relative to the spin axis. However, some have proposed the gliding of the crust and mantle over the outer
core (e.g. Andrews, 1985; Kirschvink et al., 1997; Evans, 2003; Raub et al., 2007; Piper, 2006). This has
also been proposed by Pavlenkova (2012), who had earlier proposed that gliding may take place along the
400- or 670-km mantle discontinuity (Pavlenkova, 1995). Based on palaeomagnetic data and plate-tectonic
theories about relatively fixed hotspots, Gordon (1987, 1995) and Kent & May (1987) held that TPW had
not exceeded 20° over the past 200 Myr. Besse & Courtillot (2002, 2003) found 30° of TPW over the last
200 Myr, and Steinberger & Torsvik (2008) found 56º of TPW over the past 320 Myr, but in both cases net
TPW was virtually zero.
Some authors would refer to at least variants (1) and (3) as ‘apparent polar wander’ rather than ‘true polar wander’
(see Northrop & Meyerhoff, 1963). ‘Apparent polar wander’ also refers of course to the plate-tectonic scenario in
which individual lithospheric plates move with respect to the rotation poles.
1
40
Fig. 5.2. Plate-tectonic cartoon of TPW (Evans, 2003; Raub et al., 2007). Changes in Earth’s moment of inertia are
driven by mass redistribution in the mantle and by changes in surface loading. (a) Mantle convection incorporates rising
and sinking density anomalies (light and dark grey, respectively). Due to viscosity, these vertical motions deform the
upper and lower boundaries, and any internal discontinuities, of the mantle. (b) A dynamic planet spins stably and
conserves momentum by shifting positive inertial anomalies (upwellings) toward the equator and negative inertial
anomalies (downwellings) toward the poles via TPW. The outer core-derived geomagnetic field remains aligned with
the spin axis, as does the equatorial bulge (exaggerated) and climatic zonation. Continents ride in unison on top of the
migrating mantle. (Evans, 2003, fig. 1)
Historical background and theoretical debate
In the 18th century, French naturalists Comte Georges-Louis de Buffon and Georges Cuvier considered
polar wandering to be a possible explanation for radically different past climates, such as great warmth in the
polar regions and glaciation near the equator. In the 19th century, this view was held by geologists such as
Henry James (1860) and John Evans (1866), and by astronomers John Lubbock (1848) and Giovanni
Schiaparelli (1889). It was recognized that polar wander would cause compression in regions moving
poleward and tension in regions moving equatorward, and would result in the emergence or submergence of
land.
The feasibility of true polar wander has long been the subject of controversy. George Airy (1860) argued
that, if the Earth were a perfectly rigid sphere, the formation of a mountain with a mass 1/1000th of that of
the equatorial bulge would cause the poles to wander only 2 to 3 miles. James Croll (1886) argued that even
if one-tenth of the Earth’s entire surface were elevated to a height of 10,000 ft, this would cause the poles to
wander only 3º17', and if a continent 10 times the size of Europe were elevated 2 miles, this would do little
more than bring London to the latitude of Edinburgh, or vice versa.
George Darwin (1877, 1878) held that the poles might wander indefinitely if the Earth were more or less
plastic, but not if it were rigid, as he believed it to be. He argued that the pole could move up to 3º in any one
geologic period, and might have moved up to 10 to 15º, ‘in a devious course’, since the Earth’s
consolidation, possibly returning to near its original position. In Darwin’s view, William Thomson (Lord
Kelvin) had demonstrated that the Earth is indeed rigid. Thomson (1876) rejected the idea that the Earth’s
rigid outer shell enclosed a liquid interior and held that sudden large-scale polar wander was impossible; he
believed that in ancient times the geographic poles might have gradually wandered up to 40º or more to their
present positions, ‘without at any time any perceptible sudden disturbance of either land or water’. Although
Thomson granted the possibility of considerable polar wandering during the early plastic stage of the Earth,
he held that virtual rigidity had prevailed throughout most of geologic history (Barrell, 1914).
Chandler’s discovery of the Earth’s polar wobble, with a period of about 428 days, in 1891 showed that the
Earth was not perfectly rigid. The existence of this free nutation had been predicted by Isaac Newton in his
41
Principia Mathematica (1687, bk. 1, prop. 66, cor. 20-22), and by Leonhard Euler in 1765. Euler pointed out
that in a rigid spheroid, if the axis of rotation did not exactly coincide with the axis of figure, the former
would revolve around the latter. Based on the Earth’s ellipticity, he predicted that this revolution would have
a period of 305 days. In 1892 Newcomb showed that the discrepancy between the predicted and actual
period was due to the fact that the Earth is not absolutely rigid. The difference implied that the Earth’s body
had an elasticity comparable to that of steel, but did not possess plasticity (Barrell, 1914).
At the start of the 20th century, a group of Germans proposed a theory of ‘polar pendulation’: the poles
supposedly swung back and forth along the 10ºE meridian, with the axis of oscillation passing through
Ecuador and Sumatra, which therefore never changed their latitude (Reibisch, 1901; Simroth, 1907).
Kreichgauer (1902), on the other hand, postulated that the north pole had wandered from the Antarctic in the
Precambrian, through the Pacific Ocean, across Alaska and Greenland to its present position, while Jacobitti
(1912) held that the north pole lay in the South Atlantic in Cambrian times, then moved across South Africa,
India, Australia, the Pacific Ocean, Canada and Greenland to its present location. In 1912 Alfred Wegener
proposed not only continental drift but also the movement of the north pole from the vicinity of Hawaii to its
present position since Palaeozoic time, largely on the basis of palaeoclimatic evidence (Wegener, 1912,
1929; Köppen & Wegener, 1924).
Barrell (1914) cited various palaeoclimatic and palaeontological data inconsistent with the different theories
of polar wander, but noted that ‘such objections can always be met and conquered by a sufficiently
ingenious advocate’. Some advocates tried to demonstrate polar wander in different periods by citing
evidence of crustal extension in some regions (claimed to have moved equatorward) and compression in
others (claimed to have moved poleward), but Barrell said that there were just as many conflicts as
agreements.
Gold (1955) postulated that large-scale polar wandering could be expected to occur over geologic time in a
plastically deformable Earth: if the Earth were a perfect sphere instead of a flattened spheroid, ‘the smallest
beetle walking over it would be able to change the axis of rotation relative to markings on the sphere by an
arbitrarily large angle; the axis of rotation in space would change by a small angle only’. Gold inferred from
the damping of the Chandler wobble that if a continent the size of South America were raised by 3 metres in
a million years, this would result in a ‘large-angle change’ in the Earth’s orientation relative to its spin axis
during the same period. He believed that the spin axis could have swung through 90o several times during
the Earth’s history on ‘a timescale of the order of 105 or 106 years, but scarcely longer’, leading to ‘drastic
changes of climate’. Jeffreys (1976) argued that there were great difficulties in attributing the damping of the
Chandler wobble to elastoviscosity, adding that if this hypothesis is rejected, ‘the whole explanation of polar
wandering breaks down’ (p. 481).
Gold assumed a Maxwell (elastoviscous) Earth model, in which the Earth has no finite strength, rather than a
Kelvin-Voigt (firmoviscous) model, in which the Earth does possess finite strength. On a Maxwell Earth, the
material forming the Earth yields by flow under stress differences of arbitrarily small magnitude, with the
result that polar wandering occurs in response to any exciting force, however small (including that caused by
Gold’s beetle). However, if the Earth possesses finite strength (a non-zero yield stress), its large moment of
inertia will be overcome and polar wandering will take place only when the excitation stress exceeds the
threshold.
Evidence that the Earth does possesses some strength includes the occurrence of earthquakes to depths of
700 km (Northrop & Meyerhoff, 1963). Further possible evidence is the equatorial bulge, which is believed
to result from the Earth’s rotation but is about 200 metres larger than it should be in a hydrostatic Earth.
MacDonald (1963, 1965) and McKenzie (1966) argued that it arose some 107 years ago, when the Earth used
to rotate faster. This 107-year delay in the Earth’s response implied a lower-mantle viscosity on the order of
1025 Pa s (pascal-seconds), compared with values of 1020 to 1021 Pa s for the upper mantle deduced from
postglacial uplift studies, and would preclude large-scale polar wander.
42
Goldreich & Toomre (1969) dismissed MacDonald’s fossil bulge
argument on the grounds that the excess (nonhydrostatic) part of the
Earth’s equatorial bulge is distinctly triaxial and therefore not the
result of the Earth’s once faster rotation. On the assumption that the
Earth is quasi-rigid and that the mantle does not possesses either
sufficient viscosity or permanent strength to prevent polar
wandering, they held that the modest redistribution of masses within
the Earth by mantle convection would cause large and sometimes
rapid polar wander, which could amount to 90º in 400 Myr. They
conceded that it is by no means certain that the Earth does not
possess finite strength.
Munk & MacDonald (1975) argued that if the mantle were anelastic
enough to permit large-scale polar wander, the poles would shift so
Fig. 5.3. Goldreich & Toomre’s
as to place the continents as well as possible on top of the equatorial
simulated polar wandering curve
bulge. Given the present distribution of the continents, the north
(1969, fig. 3). The meridians and the
circles of latitude are both drawn 30º
geographic pole ought to be located near Hawaii, or at least moving
apart. The markers along the path
towards it. The fact that it is not doing so implies that the Earth (or at
denote ‘time’ t = 0.2, 0.4, 0.6, etc.
least its outer shell) has sufficient finite strength to withstand the
stresses imposed by the continent-ocean system. Munk &
MacDonald pointed out that this conclusion can be avoided by assuming that the stresses in question are
balanced by mantle inhomogeneities. The Earth would require a finite strength of 10 bars (1 MPa) to prevent
polar wandering in response to the continental excitation function. The excitation function arising from
mantle inhomogeneities is probably far larger than that of the continents. Munk & MacDonald say that the
Earth could certainly possess a strength of 100 bars. Moreover, the fact that major gravity anomalies are
associated with Palaeozoic mountain chains indicates that large stress differences can persist for very long
periods and implies a strength of 150-300 bars in the upper 600 km. The Earth might therefore have
sufficient strength to prevent changes in mass distribution on or in the planet from causing significant polar
wander, especially if polar wander involved only a thin outer shell. Munk & MacDonald held that
palaeoclimatic and palaeontological data provided ‘little positive evidence’ for polar wandering on the scale
suggested by palaeomagnetic data. Their overall conclusion was that the problem of polar wander was
‘unsolved’.
An elastoviscous mantle remains a fundamental tenet of plate tectonics. Assuming a simple Maxwell
(elastoviscous) Earth, an average mantle viscosity of 3 x 1022 Pa s, and the existence of mantle convection
cells, and assuming that the nonhydrostatic part of the equatorial bulge is dominated by mantle convection,
Tsai & Stevenson (2007) calculated that the maximum rate of TPW is 61° in 100 Myr and 8° in 10 Myr.
They stated that the maximum TPW speed is 2.4º/Myr, but this is only achieved for a relatively small length
of time, in the middle of a TPW event. They regarded the mantle viscosity structure as the biggest
uncertainty in estimating TPW and acknowledged that the lower mantle could have a far higher viscosity
than currently believed.
Convection and isostasy
In plate tectonics, mantle convection currents were originally considered to be the major driving force of
plate movements, but nowadays the emphasis is placed on ‘slab-pull’, ‘ridge-push’ and ‘trench-suction’,
though their adequacy is very much in doubt (Lowman, 1986; Keith, 1993). As already noted, mantle
convection is regarded as a principal cause of polar wander. However, the existence of large-scale
convection is highly uncertain. Plate tectonicists initially proposed that mantle-deep convection currents
welled up beneath midocean ridges, with downwelling occurring beneath ocean trenches (subduction zones).
The existence of layering in the mantle cast doubt on whole-mantle convection and led to the development
of two-layer convection models. However, seismic tomography has failed to provide clear evidence of large,
plate-propelling convection cells in both the upper and lower mantle (Anderson et al., 1992; Jordan et al.,
43
1993). Large-scale horizontal convection in the upper mantle would be difficult to reconcile with the
existence of mantle roots extending to depths of up to 400 km (O’Reilly et al., 2009). In addition, Sandwell
& Renkin (1988) found no expression in the geoid of convection-cell geometry.
Meyerhoff (1995, p. 165) highlighted another problem:
Almost all theoretical treatments … describe a form of cellular convection generally akin to the Rayleigh-Bénard
cellular convection of the laboratory … The vertical walls of all postulated cells are relatively straight, only mildly
curved. By no stretch of the imagination can any cellular convection model yet proposed accommodate the sinuous and
vortical contortions of the arcs, modern and ancient, of southeastern Asia! To do so would require an organized
convection cell with contorted walls turning 180° upon themselves in several locales from west of India through the
Banda arc to the Philippines arc. The very geometry of the southeastern Asian region simply defies all convective
schemes proposed …
Convection is probably impossible because the mantle appears to obey the modified Lomnitz law of
anelasticity, rather than being elastoviscous, as most geophysicists assume (Jeffreys, 1974, 1976; Wesson,
1974a,b). This means that convection is a self-damping process; if convection ever took place, the velocity
of the currents would decrease toward zero. The modified Lomnitz law is supported by data on ancient
gravity anomalies, Earth’s nonhydrostatic equatorial bulge, the existence and damping of the Chandler
wobble, and the Moon’s orbit, rotation, shape, and free librations.
Wesson (1974a,b) argued that convection cells may be inferred not to exist because there is no convectiongeoid correlation, no convection-heat-flow correlation, and no convection-volcanicity correlation. Chemical
boundaries within the mantle, proposed on geochemical and seismic grounds, are incompatible with
convection, even assuming that currents could pass phase changes in the mantle. Even if only the lower
mantle behaves according to modified Lomnitz law while the upper mantle obeys the elastoviscous law, so
that convection is possible above a depth of 700 km, polar wander would probably not occur due to the very
high viscosity of the lower mantle.
Isostasy refers to the theory that the Earth’s crust responds to an added load (e.g. from a glacier) or a
diminished load (e.g. from erosion) by falling or rising respectively, so as to establish a state of gravitational
equilibrium. It is said to arise from the fact that the crust ‘floats’ on the mantle, or the lithosphere ‘floats’ on
the asthenosphere. Jeffreys (1976) says that isostasy is only a first approximation to the facts. He writes:
If it is true in general, it will follow that every region of positive gravity anomalies is sinking, and every region of
negative ones rising. In any mountain system undergoing denudation, either compensation should keep pace with
denudation and there would be no systematic isostatic gravity anomalies, or it would fail to keep pace and the gravity
anomalies would be negative. Where loads have been added recently, there should similarly be either no gravity
anomalies or systematically positive ones. At every point these consequences are contrary to the facts. (p. 458-9)
In mountain ranges there are often significant residual loads that can be supported only by appreciable
strength in what has been assumed to be a region of great weakness (the asthenosphere).
The postglacial uplift of Scandinavia and part of Canada is cited as a classic case of ‘isostatic rebound’, and
is used to determine the value of the mantle’s elastoviscosity. Jeffreys (1974, 1976) says that the estimated
viscosity derived from a study of Scandinavia and Canada is illusory. The uplift is supposedly due to viscous
recovery following melting of the ice sheet, but the rate of rise within Fennoscandia is far from being closely
correlated with the gravity anomalies. Moreover, Fennoscandia lies within a region of positive anomalies
and therefore ought to be sinking. James (1997) argues that the lithosphere is strong enough to lift
continental ice sheets rather than being depressed by them.
As Jeffreys (1976, p. 459-60) notes, gravity anomalies greater than those of Fennoscandia or Canada exist in
India, but are not associated with systematic vertical movements. Cyprus is a region of strong positive
gravity anomalies but has risen since the Pleistocene. Other parts of the Mediterranean region show similar
anomalies, especially Sicily; some have risen in historical times, others sunk, while some have done first one
44
and then the other. As regards regions with significant and prolonged denudation, there are systematically
positive anomalies over the Welsh mountains and the Highlands of Scotland. The Greater Caucasus is
overloaded, yet it is rising rather than subsiding, while the foredeep to the north is underloaded, but is
subsiding rather than rising (Beloussov, 1980, p. 260).
It is still widely assumed that the Earth is always tending to a state of perfect isostasy, despite evidence that
great changes in the opposite direction have continued over intervals of the order of 107 years. Many old
mountain systems appear to have been originally of similar height to the Alps but have been denuded to
heights of about 1 km. If isostasy had been maintained during this process, the denudation should have
removed all the sediments and cut deeply into the granitic layer. Since this has not happened, the mountains
must have been pulled down from below (Jeffreys, 1976, p. 492). Jeffreys concludes that ‘the hypothesis of
viscous flow, always tending to produce perfect isostasy, is clearly wrong’; rocks at great depths have a nonzero strength and viscous flow is negligible unless the stress differences exceed the strength (p. 460-1).
Significant breaches of isostasy occur mainly in regions of tectonic activity, while stable ancient platforms
and deep oceanic basins are in a state much closer to isostasy.
Current estimates of upper mantle viscosity range from about 3 x 1018 to 3 x 1022 Pa s (Sato, 1991;
Vermeersen et al., 1997). On the basis of crustal deformation in Fennoscandia, Zhao et al. (2012) found an
upper mantle viscosity of between 3.4 and 5.0 x 1020 Pa s, and a lower mantle viscosity of between 7 and 13
x 1021 Pa s. By contrast, MacDonald (1965) argued that a comparison of the figure of the Earth obtained
from satellite observations with that calculated on the assumption that the Earth is in hydrostatic equilibrium
demonstrates that stress differences of the order of 100 bars exist in the mantle. If the mantle is
elastoviscous, the average viscosity would have to be as high as 1025 Pa s, and the existence of ancient
geologic features associated with large gravity anomalies implies that even this figure is too small.
Plate motions plus polar wander
Plate tectonicists believe that true polar wander is the result of changes in the planetary moment of inertia
caused by mass redistributions linked to mantle convection, subducting slabs and upwelling plumes (Besse
et al., 2011). The equatorial bulge tends to stabilize the Earth from tumbling, but the rotation axis adjusts to
the maximum principal inertia axis on a timescale of 2 to 6 Myr (Greff-Lefftz & Besse, 2011). The excess
bulge has been attributed to mantle convection (Cambiotti, 2012).
Many studies of TPW are based on the assumption that hotspots (longstanding active volcanoes) provide a
valid reference frame for the mantle. Hotspots are assumed to be the surface manifestations of plumes,
anchored deep in the mantle, which have left traces in the form of nearly linear chains of extinct volcanoes
on plates passing over them. About 20 hotspots were originally proposed; nowadays there are said to be
about 6 primary deep mantle plumes and as many as 5200 moderate-size plumes. The very existence of deep
mantle plumes is controversial, even among plate tectonicists, as there is growing evidence that hotspots are
mostly shallow features. Seismic tomography provides no clear-cut evidence of narrow upwellings beneath
hotspots (Anderson & Natland, 2005; Anderson, 2007). It is widely accepted that Pacific hotspots move
relative to Indo-Atlantic hotspots and to Iceland. One group maintains that speeds are 3 mm/yr, another 1020 mm/yr or more (Gordon, 1995). However, many workers still believe that certain hotspots are sufficiently
fixed to serve as a valid reference frame. If large-scale plate motion is a myth, hotspot trails have nothing to
do with plates moving over hotspots and must have other causes, e.g. propagating rifts.
Steinberger & Torsvik (2008) studied palaeomagnetically-derived continental motion and rotation over the
past 320 Myr and identified both a steady northward motion and, during certain time intervals, clockwise
and anticlockwise rotations, which they interpreted as evidence for TPW. They found ~18° anticlockwise
rotation at about 250-220 Ma, the same amount of clockwise rotation at 195-145 Ma, ~10° clockwise
rotation at 145-135 Ma, and the same amount of anticlockwise rotation at 110-100 Ma. The overall net
rotation over this period was nearly zero. They noted that some continental motions did not fit this picture,
but attributed this to poor palaeomagnetic data. They considered it unlikely that the steady northward drift
represents TPW. TPW rates are on the order of 0.45-0.8°/Myr but cumulative TPW since the late
45
Carboniferous is close to zero due to both clockwise and anticlockwise episodes of TPW centred on 0°N and
11°E (Torsvik et al., 2012). (NB: 1o ≈ 110 km.)
Fig. 5.4 shows the hotspot-based, episodic TPW path determined by Besse & Courtillot (2002). They found
periods of (quasi) standstill alternating with periods of faster TPW. The path displays a standstill at 160-130
Ma, a quasi-circular track from 130 to 70 Ma (30 km/Myr), a standstill at 50-10 Ma, and then faster motion
up to the present (100 km/Myr). In contrast, Prévot et al. (2000), using a ‘rigorously selected palaeomagnetic
database’, found that over the last 200 Myr there had been two long periods of strict standstill from the
present to 80 Ma and from approximately 150 to 200 Ma, and a single period of TPW between 80 and about
150 Ma (attributed to ‘hyperactivity of mantle plumes’); see fig. 5.5. The latter period culminated around
110 Ma in an abrupt 20° displacement of the pole, during which a speed exceeding 5°/Myr (0.5 m/yr) may
have been reached. Cambiotti et al. (2011), on the other hand, say that since the early Tertiary (50-60 Ma)
the pole has moved about 4-9º toward Greenland; they do not obtain a period of (quasi) standstill at 10-50
Ma or 0-80 Ma.
Fig. 5.4. Hotspot-based true polar wander path during the last
200 Myr, with associated 95% confidence ellipses. (Besse &
Courtillot, 2002; Besse et al., 2011, fig. 3; reprinted with
permission from Springer Science+Business Media BV)
Fig. 5.5. Time-averaged pole positions
and 95% confidence circles with respect
to the Indo-Atlantic hotspot reference
frame for four consecutive periods.
Filled symbols correspond to a pure
dipole field model and empty symbols
to a dipole plus quadrupole field model.
(Prévot et al., 2000, fig. 2a)
In contrast to Prévot et al.’s (2000) claim that TPW as fast as 0.5 m/yr took place during the Cretaceous,
estimates of long-term Mesozoic and Cenozoic TPW rates are typically about 1-5 cm/yr, with TPW
swinging back and forth along 130º and 310ºE (e.g. Besse & Courtillot, 1991). However, rapid bursts of
TPW have been proposed by other workers too. Van der Voo (1994) found rates of 70-110 km/Myr in
certain intervals of the mid-Palaeozoic. Kirschvink et al. (1997), Evans (2003), Raub et al. (2007) and Piper
(2006) discussed the possibility that the Earth had catastrophically exchanged two of its principal axes of
inertia (inertial interchange true polar wander, IITPW), i.e. that the whole lithosphere and mantle may have
rotated by up to 90º in only a few million years. Most plate tectonicists believe that evidence for superfast
events is not compelling.
Kirschvink et al. (1997) speculated that TPW of about 90º occurred from 535 to 520 Ma (525 to 508 Ma,
according to Evans, 2003), at rates of over 600 km/Myr, during the early to middle Cambrian, as a result of a
major reorganization of tectonic plates. Torsvik et al. countered that analysis of a more complete
palaeomagnetic dataset is consistent with ‘conventional plate tectonic systematics’. In response, Kirschvink
and his coworkers stated that their conclusion is based on ‘a more reliable subset of the data’ and noted that
46
the TPW hypothesis neatly explains many features of the early Cambrian geologic record (Torsvik et al.,
1998). Tsai & Stevenson (2007) rejected Kirschvink et al.’s TPW scenario as implausible because it would
require an improbably low mantle viscosity of no more than 7 x 1021 Pa s. Piper (2006) argued for 90°
inertial interchange TPW between 410 and 390 Ma in the Devonian (4.5°/Myr), ‘most plausibly related to
the avalanching of long lithosphere slabs into the lower mantle’. According to Torsvik et al. (2012, p. 362),
‘this captivating claim is not justified by palaeomagnetic data’.
Fig. 5.6. Terminal Proterozoic to late Palaeozoic apparent polar
wander (APW) path for Gondwanaland (Evans, 2003, fig. 3). Evans
(2003) and Raub et al. (2007) interpreted oscillatory APW rotations
as TPW about a common, long-lived, minimum-inertial axis (Imin)
near eastern Australia.
Plate-tectonic estimates of true polar wander are clearly highly inconsistent. They are no more reliable than
the selected palaeomagnetic data and plate-tectonic assumptions on which they are based. If explaining
palaeomagnetic data mainly in terms of large-scale plate motions is wrong, the polar wander hypotheses put
forward to explain remaining features of the data are also likely to be wrong.
Wrench tectonics
Polar wander
Based on selected palaeomagnetic data, wrench tectonics postulates large-scale true polar wander and
‘modest’ continental/plate rotations and translations (Storetvedt, 1997, 2003). Polar wander is said to involve
a reorientation of the entire Earth relative to the spin axis and to be triggered by the large-scale redistribution
of mass in the planet’s interior, mainly resulting from internal degassing and associated crustal delamination.
Polar shifts are said to be relatively rapid events separated by long periods of tectonic tranquillity. The
rotation axis has supposedly wandered by about 70º of latitude relative to the Earth’s surface from the midPalaeozoic to the present in the 0º/180º meridian plane. Around 35 Ma, near the Eocene-Oligocene (E-O)
boundary, the poles allegedly shifted about 35º to their approximate present locations within the space of 2
to 3 Myr.
According to wrench tectonics (Storetvedt, 1997, p. 246-50; Storetvedt & Bouzari, 2012) the overall
wandering of the north pole has been towards the Pacific but there have also been reversals of direction. On
the basis of palaeomagnetic data, it is contended that in the Jurassic and Lower Cretaceous the poles were
not significantly different from the present poles, with the palaeoequator running across the Central Sahara.
At about 100 Ma the equator allegedly shifted to the northern rim of the African continent, but then reverted
to the Central Sahara position at 80-60 Ma, before shifting back to the southern rim of the Mediterranean in
the Lower Tertiary. After the rapid burst of polar wander about 35 Ma, when the equator shifted from the
Mediterranean to around its present position across Central Africa, the equator moved north again in midMiocene time, cutting across Central Arabia, before returning to its present location around 5 Ma.
47
Fig. 6.1. ‘Global polar wander path’ since the midPalaeozoic, based on selected European and African
palaeomagnetic data ‘from which the effect of Alpine age
continental motions are eliminated’. UT = Upper Tertiary,
LT = Lower Tertiary, P = Upper Carboniferous-Permian, LC
= Lower Carboniferous. (Storetvedt, 1997, fig. 9.5)
Wrench tectonics claims that for at least 450 Myr (Storetvedt, 1997, p. 60), polar wander has been confined
to virtually the same plane. The axis about which the Earth is supposedly turning during polar wander passes
through two points on the equator: 0oN, 90oE and 0oN, 90oW (this is virtually the same axis as in the
pendulation theory mentioned earlier). This can only happen if density anomalies continually arise within the
Earth and are centred in the same plane. In terms of the present coordinate system, for the north pole to
move towards the Pacific, positive mass anomalies would have to arise in the northern-hemisphere quadrant
centred on the 180º meridian and/or in the antipodal quadrant; excess mass at 45o north or south would have
most effect. Polar wander (assuming mantle viscosity and strength are sufficient low for it to occur) would
then move these excess masses towards the equator. For polar wander to continue in the same direction,
positive mass anomalies would need to arise in the same quadrants as before (or in the quadrants on the
opposite side of the equator, to reverse the direction of polar wander). However, to move positive masses
already at the equator away from the equator, the new mass anomalies would have to be even greater, unless
the equatorial anomalies decline or disappear. These processes would have to continue for several hundred
million years. Such a scenario, with the poles moving backward and forward in the same plane, is extremely
contrived and seriously strains credulity. If large-scale polar wander were possible, it would surely follow a
more random path, like that simulated by Goldreich & Toomre (1969) (see fig. 5.3 above).
Since the Lower Carboniferous, according to wrench tectonics, the poles took over 300 Myr to wander 35º
(leaving aside reversals of direction), and then only 2 or 3 Myr to wander another 35º – a 100-fold
acceleration. Storetvedt (1990, p. 158) says that this is ‘well within theoretical expectations’ and as a
reference he cites Gold (1955) – who assumed a Maxwell Earth with zero strength and low viscosity. Tsai &
Stevenson (2007) calculated that TPW was limited to 8º in 10 Myr; they assumed a mantle viscosity of 3 x
1022 Pa S (which may be several orders of magnitude too low) together with the existence of plate-tectonic
processes such as mantle convection and slab subduction, as well as rising plumes.
An example of the palaeomagnetic evidence that Storetvedt presents for a 35º displacement of the poles at
the E-O boundary is given in fig. 6.2, which shows selected Mesozoic-Cenozoic palaeomagnetic poles for
Africa and Europe. The symmetrical and oppositely trending APW paths are said to signify that the two
continents rotated in opposite directions until the common cusp at approximately 180ºE, 55ºN, which is
assumed to coincide with the E-O boundary. The subsequent, joint, near-meridional APW track is said to
represent a period of rapid polar wander covering 35º of latitude. Note the complete lack of any age data in
the diagram. Storetvedt (1992, p. 205) simply says that the palaeopoles ‘are inferred to increase in age’ in the
opposite direction to the arrows. The same flaw is found in all his figures purporting to show the 35º
wandering of the poles (e.g. fig. 6.3). It would be more scientific to present the actual age determinations and
then explain why they are considered inaccurate and why assigning ages based on theoretical preconceptions
is more justified. The claim that this rapid TPW explains the major climatic changes at the E-O boundary
will be examined in a later section.
48
Fig. 6.2. Selected Mesozoic-Lower Tertiary
palaeomagnetic poles for Africa (circles), the
Canary and Cape Verde Islands, in comparison
with relevant data for Western Europe (triangles).
(Storetvedt, 1992, fig. 2)
Fig 6.3. Comparison of Mesozoic-Lower Tertiary
palaeomagnetic polar trends of Europe and
cratonic North America. Storetvedt (1992, p. 210)
writes: ‘The age-paleopole relationship should not
be taken too literally as magnetization and
physical rock ages may easily have been
disconnected by remagnetization caused by the
major phase of Alpine plate rotation in the Lower
Tertiary.’ In other words, the actual ages
determined for the virtual poles (which he fails to
give) should not be taken seriously because of
possible remagnetization; instead, we should take
his own ‘inferred’ age sequence (indicated by the
arrows) seriously. (Storetvedt, 1992, fig. 8)
The latitudinal shifts produced by the wandering poles would have their maximum values (a net shift of 35o
from mid-Palaeozoic to the E-O boundary, and 35o within 2-3 Myr thereafter) only at locations on the
0o/180o meridian; for other locations they would be smaller. The palaeomagnetic data presented by Besse &
Courtillot (2002, 2003) for the various continents and by Beaman et al. (2007) for the Pacific plate do not
appear to provide any clear support for the above wrench-tectonic claims. For instance, Greenland (relatively
near the supposed polar-wander meridian) fails to show latitude changes of the magnitude required. At the
very least this shows that very varied conclusions can be drawn on the basis of palaeomagnetism, depending
on which data are selected and how they are manipulated and interpreted.
49
If palaeomagnetic data are considered reliable, a ‘global’ polar wander path needs to be based on global data,
not just on data for Europe and Africa; it would be interesting to see how much data selection and processing
are required to obtain a consistent scenario. Fig. 6.4 shows Besse & Courtillot’s (2002) master apparent
polar wander path for Africa for the past 200 Myr. Note that it doubles back on itself in the 200-90 Ma
interval. This can also be seen in the South African APWP for the Phanerozoic presented in fig. 6.5. Fig. 6.6
shows the South American APWP, which is remarkable in that it displays very little apparent polar wander;
most poles have latitudes higher than 80º. It would be interesting to hear how the APWPs for these two
continents are to be reconciled with the wrench-tectonic claim of 70º of northward TPW in the past 350 Myr,
and 35º of TPW around 35-32 Ma – especially if continental rotations are supposed to have begun only in
the late Cretaceous. The APWPs for Siberia and Baltica shown in fig. 3.8 would also provide an interesting
challenge for wrench tectonics. The more palaeomagnetic data that are taken into account, the more rotations
and translations, and possible disruption of geological continuity, there will need to be.
Fig. 6.4. Master APWP for Africa from the past 200 Myr, with 95%
confidence ellipses (averages every 10 Myr, with a 20 Myr sliding
window); mean ages for each time window are indicated. (Besse &
Courtillot, 2002, fig. 1)
Fig. 6.5. South African APWP for the Phanerozoic. (a) South
African poles only. (b) Smoothed APWP spline path using
master path approach for Gondwana in South African
coordinates. (Tauxe, 2013, fig. 16.10)
Fig. 6.6. South American APWP for the past 200
Myr. (Besse & Courtillot, 2002, fig. 9)
50
Fig. 6.7. Wrench-tectonic palaeoequators
for the Carboniferous, Permian, Lower
Tertiary, and Upper Precambrian-Lower
Palaeozoic. (Storetvedt & Bouzari, 2012,
fig. 3)
Storetvedt (1997) claims that the Caledonian-Appalachian,
Hercynian, and Alpine-Himalayan foldbelts (or at least parts of
them) lay roughly along contemporary palaeoequators (see fig. 6.7).
He attributes heightened tectonic activity at the equator to the effect
of centrifugal force (maximum at the equator) on mantle pluming
(occurring in pulses). It is therefore strange that we do not see more
tectonic and volcanic activity along the present equator. According
to wrench tectonics, all of the Earth’s surface between the present
equator and the Lower Carboniferous equator must have had an
equatorial bulge at least once; since the poles are supposed to have
moved back and forth, some areas must have been part of the
equatorial bulge up to four times. If we consider the geographic
distribution of orogenic activity in different periods, we can see that
most of it does not coincide with wrench-tectonic palaeoequators
(figs. 6.8-6.10). Storetvedt also argues that heightened tectonic
activity would be expected at near-perpendicular angles to
palaeoequators, and gives the Ural belt, Oslo Rift, and Rhine
Graben as examples (p. 359). But again, the pickings are
unimpressive.
Fig. 6.8. Continental orogenic belts, by time
of major orogenic distribution. (After
Burchfiel, 1990; www.accessscience.com)
Fig. 6.9. Distribution of
orogenies with similar
ages to the Caledonian/
Variscan orogeny (380 to
280 Ma).
(en.wikipedia.org)
51
Fig. 6.10. Areas with
mid-Permian folding
and/or distinctive
transgressive
unconformity associated
with the beginning of the
Hunter-Bowen
(Indosinian) orogenic
folding phase. The terms
used in different parts of
the world for this
tectonic phase are
shown. (Dickins, 1994a,
fig. 2)
McKenzie & Priestley (2008) note that the tectonic deformation of the continents has been dominated by
continental ‘cores’ – i.e. areas underlain by thick, cold lithosphere; these area do not include all cratons,
often extend beyond the boundaries of present cratons, and also include non-cratonic areas. The North
American core (fig. 6.11) is fringed by the Appalachians in the east and the Rocky Mountains in the west.
The western part of the North Eurasian core (which comprises the Baltic Shield and the Russian and Siberian
Platforms; fig. 6.12) has controlled the tectonic evolution of Europe, where the Caledonian and Hercynian
foldbelts wrap round its margins. The Urals cut across the core, and were presumably underlain by thinner
lithosphere when they formed.
Fig. 6.11. (a) Tectonic map of North America (Holmes, 1965, fig. 811). (b) Contours of lithospheric thickness,
calculated from shear-wave velocity. Magenta circles = locations of diamond-bearing kimberlites; yellow circles =
locations of alkali basalts containing mantle nodules whose mineral compositions were used to estimate the lithospheric
thickness. Numbers in white boxes show the estimated thickness of the lithosphere (thicknesses are about 100 km less
than those found by Shapiro et al., 2004; see fig. 2.8). The yellow line shows the approximate boundary of the North
American craton, labelled ‘a’. (McKenzie & Priestley, 2008, fig. 2)
52
Fig. 6.12. (a) Tectonic map of Europe (Holmes, 1965, fig. 807). (b) The yellow line shows the approximate boundary
of the East European Craton, labelled ‘a’. See key to fig. 6.11. (McKenzie & Priestley, 2008, fig. 3)
Also worth noting is that, according to theory (Jeffreys, 1976) and laboratory experiment (Bucher, 1956),
heated spheres cool by rupture along great circles. Pre-Tertiary orogenic belts fall along approximate greatcircle belts or on parts of such belts. But only two partial great-circle fracture zones survive on Earth today:
the highly active circum-Pacific seismotectonic belt (Benioff zones) and the almost defunct AlpineHimalayan (Tethyan) belt (Meyerhoff et al., 1996a; Meyerhoff & Meyerhoff, 1974a). Both belts probably
originated in the Precambrian and have been episodically reactivated.
Spinning continents
According to wrench tectonics, since the late Cretaceous continents have undergone various degrees of
rotation, together with part of the surrounding oceanic lithosphere, perhaps as far as the official ‘plate
boundaries’. Although the proposed motions are more modest than in plate tectonics, they are still
problematic, given the dramatic variations in lithospheric thickness and the lack of a universal
asthenosphere. A summary of the claimed movements is given below. When palaeomagnetic data require us
to believe in geologically implausible events, a reasonable conclusion is that the data should not be taken too
seriously.
North America is said to have rotated clockwise by ~25° relative to Europe/Eurasia about an Euler pole in
the northern part of the continent (Storetvedt, 1997, p. 72; Storetvedt &Longhinos, 2011, p. 21). Storetvedt
& Longhinos (2012) say that North America has rotated ~30º clockwise relative to the more sluggish
Eurasian landmass, but rather than an in situ rotation of the entire landmass, a substantial portion is the result
of internal deformation, with increasing clockwise torsion in the southern half of the continent, totalling
some 55°.
Prior to the Eocene-Oligocene boundary, Europe rotated ~25º clockwise, and Africa rotated anticlockwise by
the same amount (Storetvedt, 1997, p. 246). Africa and the whole of Eurasia are said to be subject to
ongoing rotation, though Storetvedt (1997, p. 248) also states that the relative rotation of Africa and Europe
ceased at the E-O boundary.
South America rotated ~20º clockwise, probably in the Eocene, around an Euler pole in Northeast Brazil, but
at the same time the northern half of the continent (north of 20oS) rotated anticlockwise, which contributed
to tectonic bending of the Central Andean chain (Storetvedt, 1992, p. 211; 1997, p. 307-12). The rotation of
South America has since been revised to ~10º. South America is also said to have shifted southward by ~20º
(Storetvedt & Longhinos, 2012). The inertial forces that supposedly cause continents to rotate should make
South America rotate anticlockwise, but ‘tectonic interaction’ across the relatively narrow equatorial
Atlantic allegedly gave rise to considerable ‘tectonic pressure’ that forced South America to swing
clockwise, though at the same time the stress field caused the northern part of the continent to swing
anticlockwise!
53
Greater India rotated ~135º clockwise near the Cretaceous-Tertiary boundary before redocking with Asia in
the Lower Palaeocene in approximately its present orientation (Storetvedt, 1990, p. 141; 1997, p. 243).
Antarctica rotated 140o clockwise and Australia 70o anticlockwise (Storetvedt, 1997, p. 335, 350). The
rotation of Australia continued past the E-O boundary (1992, p. 209), and Antarctica is still rotating. In
addition, the Antarctica-Australia-New Zealand-Melanesia block has moved 1700 km (15o) northeastward
relative to Africa (1997, p. 350).
Fig. 6.13. The nodding of Iberia, according to wrench
tectonics. (Storetvedt, 1997, fig. 9.24)
Central Iran has rotated more than 90° anticlockwise
since the Triassic (Storetvedt & Bouzari, 2012). The
Italian microplate rotated 10-15º clockwise in the
Lower Tertiary (in contrast to the 30-40º
anticlockwise rotation postulated by plate tectonics)
(Storetvedt, 1997, p. 271-2). During the Tertiary,
Madeira rotated ~50º anticlockwise followed by
~25º clockwise (1990, p. 166). In the Upper
Cretaceous, Iberia rotated at least 40º anticlockwise
at 100-90 Ma, followed by ~70º clockwise at 75-65
Ma (1997, p. 281).
Fig. 6.14. Major tectonic trends on and around the South American continent. CPM = Central Pacific Megatrend. NPM
= North Pacific Megatrend. (Choi, 2002, fig. 4)
54
The Earth’s surface is crisscrossed by structural lineaments, apparently originating in Precambrian time,
which often run for thousands of kilometres across ocean basins and continents – thereby refuting the
lithospheric motions postulated by plate tectonics. Storetvedt & Longhinos (2012) say that the continental
and seafloor movements proposed by wrench tectonics have not significantly disrupted the continuity of
these lineaments. They should, however, have left some trace. But the lineaments in fig. 6.14 show no
obvious evidence of South America’s alleged 20º southward motion and 10º clockwise rotation, and Africa’s
25º anticlockwise rotation. The lineaments running across Australia and into the seafloor (fig. 6.15)
contradict the 70º anticlockwise rotation of Australia. Similarly, the ~135º clockwise pirouetting of Greater
India is contradicted by the lineaments and stratigraphic evidence shown in figures 4.6, 4.8 and 4.9.
Fig. 6.15. Ocean-floor lineaments around Australia with
major continental lineaments superimposed (O’Driscoll,
1986, Elliott, 1994). Deep-sea drillholes and dredging
sites are indicated. CLPT F.Z. = Clipperton Fracture
Zone. (Choi, 1997, fig. 2)
The GPS velocity vectors in fig. 6.16 show no ongoing clockwise rotation of North and South America, or
anticlockwise rotation of Africa and Australia. Storetvedt & Longhinos (2012) argue that GPS data do
support continuing clockwise rotation of the whole of Eurasia. They cite Zemtsov (2007), but the latter
provides some important additional information. First, the epicentre of the present rotation would be located
in the Eastern Himalayas (95ºE, 30ºN), in China, far from the geometrical centre of the continent. Second,
the angular rotation vectors increase from the periphery of the continent to the central domain of rotation –
the opposite of what we would expect if Eurasia were rotating as a rigid unit. It would be strange if the
largest continent were still rotating (and deforming) – faster towards the centre than on the periphery – while
several smaller continents have stopped rotating.
Storetvedt & Bouzari (2012) conclude from the GPS velocity field shown in fig. 6.17 that the Middle East is
currently rotating counterclockwise. However, the velocity vectors clearly show that the region is not
rotating as a coherent, relatively rigid unit. The rotation pole could be placed in the eastern Mediterranean
(perpendicular to most of the velocity vectors), but the velocities should then increase with increasing
distance from the rotation pole, which they clearly do not. As Reilinger et al. (2006) say, ‘The velocity field
is characterized well by a system of undeforming regions separated by concentrated zones of deformation
(widths <<100 km). Deformation zones correlate closely with mapped, active faults and historic seismicity,
55
and coherent regions with seismically quiet zones ...’ Storetvedt & Bouzari (2012) go as far as to infer from
fig. 6.17 that not only Arabia, Iran, Anatolia and the Aegean but also Africa as a whole are still rotating
anticlockwise – but they fail to show the velocity field for Africa. As seen in fig. 6.16, there is no compelling
evidence of anticlockwise rotation.
Fig. 6.16. Crustal motions according to GPS. (http://en.wikipedia.org)
Fig. 6.17. Map showing decimated GPS-derived
velocities for the Middle East relative to Eurasia,
with 1σ error ellipses. (Reilinger et al., 2006, fig. 2)
Fig. 6.18. Horizontal station velocities from the
SCAR GPS network in Antarctica, showing one of
the solutions for the rotation pole. (Dietrich et al.,
2004; http://rses.anu.edu.au)
56
GPS data are also said to demonstrate the ongoing clockwise rotation of Antarctica (Storetvedt, 2010). What
is certain is that Antarctica is not rotating about a point within itself; the southern-hemisphere Euler pole lies
in the surrounding ocean (Bouin & Vigny, 2000; Donnellan & Luyendyk, 2004; Jiang et al., 2009; Dietrich et
al., 2004) (fig. 6.18). Jiang et al. (2009) say that, in general, the Antarctic plate is moving towards the South
American plate, while departing gradually from the Australian plate.
Space-geodetic data provide valuable information on current crustal motions and local stress fields, but
drawing generalized conclusions about the motion of entire lithospheric ‘plates’ and extrapolating the data
millions of years into the past or future are highly dubious practices.
Inertial forces
According to wrench tectonics (Storetvedt, 2007), hemispherical torsion, in-situ continental/plate rotations,
and the formation of tectonomagmatic belts are mainly caused by the centrifugal force, the Coriolis force,
the Eötvös (Polflucht or pole-fleeing) force, and tidal forces (the gravitational pull of the Sun and Moon).
Wegener’s theory invoked the same forces (and later, mantle convection) to explain the presumed westward
drift of the continents through the denser seafloor. However, the objection was raised that none of these
forces could exert shear stresses of more than about 4000 dynes/cm2 at the base of a continent, whereas the
outer mantle appears to have a long-term strength of at least 108 dynes/cm² (100 bars) (Chadwick, 1962;
Jeffreys, 1976).
The Coriolis effect is caused by the Earth’s rotation and the inertia of the body experiencing the effect. The
Coriolis effect refers to the fact that a horizontally moving body appears to be deflected to the right in the
northern hemisphere and to the left in the southern hemisphere (Pedlosky, 1979; Price, 2006). In addition,
objects travelling upwards or downwards are deflected to the west or east respectively, though this effect is
far less significant. ‘For a given horizontal motion the strongest horizontal deflection is at the poles and there
is no horizontal deflection at the equator; for vertical motion the opposite is true’ (Persson, 1998). In
addition to its horizontal component, the Coriolis effect has a vertical component, known as the Eötvös
effect. The Eötvös effect means that eastward-travelling objects are deflected upwards (feel lighter), while
westward-travelling objects are deflected downwards (feel heavier). This aspect of the Coriolis effect is
greatest near the equator. The term ‘Coriolis effect’ is often used to refer only to the horizontal component.
The Coriolis effect is proportional to the Earth’s rotation rate and to the speed of the moving object in
question. Because the Earth rotates only once a day, the effect is small, but it becomes noticeable for largescale atmospheric and oceanic circulation. It explains why high-pressure wind systems rotate clockwise in
the northern hemisphere and counter-clockwise in the southern hemisphere; low-pressure systems rotate in
the opposite direction. A similar situation applies to ocean gyres: circulation is clockwise in the northern
hemisphere and anticlockwise in the southern hemisphere in the case of high-pressure gyres; the opposite
applies in low-pressure gyres.
Ricard (2007) states that the Coriolis force is 20 trillion times weaker than the force of gravity, and the
centrifugal force is 291 times weaker, and that, even on the most generous assumptions, ‘inertia and Coriolis
accelerations still play a negligible role in mantle dynamics’. Goldreich & Toomre (1969), too, held that
within the mantle ‘all Coriolis forces must be utterly negligible’. Hughes (1973), on the other hand, held that
Coriolis force perturbations of mantle convection could explain fracture zones at the Earth’s surface,
including the lineations on the Pacific and Indian Ocean floors, and the East African and mid-Atlantic rift
systems. He argued that the Eurasian block and the Pacific Ocean floor are rotating anticlockwise. However,
his model makes various unreasonable assumptions: two-phase mantle-wide convection involving four
convection cells in a tetrahedral arrangement with polar symmetry; the existence of very high flow rates and
extremely low viscosity in the asthenosphere; and the existence of a global and continuous asthenosphere
100 km below the Earth’s surface.
The pole-fleeing force (another aspect of the Eötvös effect) is the combined effect of Earth rotation and the
principle of isostasy. It was first noted by Kreichgauer in 1900 (Wegener, 1929, p. 179), and was explained
57
in more detail by Eötvös (1913) (see Scheidegger, 1963). The Eötvös force tends to displace a floating body
towards regions where gravity is least. In particular, an isostatically compensated landmass will tend to
move towards the equator. However, according to Jeffreys (1976), the stress produced is only about 4400
dynes/cm2. Wegener assumed that the Eötvös force, in tending to make a floating body move towards the
equator, would, on a rotating planet, produce a steady drift to the west. Jeffreys (1976, p. 479-83) objected
that since the Eötvös force is a shear applied to the surface, mantle viscosity would have to be only ~10 15
Pa s – a value so low that the Chandler wobble would be damped out within a few days. Assuming a
viscosity of at least 5 x 1019 Pa s, implied by latitude variation data, the crust would be displaced through a
radian (~57o) in about 3000 Myr. Assuming the viscosities inferred from the Fennoscandian uplift, this
period would have to be multiplied by about 100. Jeffreys concluded that the Eötvös force cannot produce
displacements of geological importance.
According to Gasperid & Chierici (1996), the Eötvös force is no more than few mGals if the decoupling
surface lies at the crust-mantle boundary (as in Wegener’s theory), but is an order of magnitude larger (2030 mGals at mid-latitudes) if the same surface lies at the lithosphere-asthenosphere boundary. Caputo (1986)
argued that if the density of a slice of lithosphere is lower than that of the surrounding rocks, the block will
move toward the equator (Polfluchtkraft), whereas if the density is higher, the block will move toward the
poles (Äquatorfluchtkraft). Gasperid & Chierici say that the Polfluchtkraft could have contributed to
Gondwanaland’s alleged 4000-km-long drift towards the equator and to its breakup. They ignore the major
variations in lithospheric thickness and lack of a universal asthenosphere.
According to wrench tectonics, wrench deformation is at its maximum in the palaeoequatorial zone and is
governed by inertial forces such as the Coriolis force and centrifugal force (Storetvedt & Bouzari, 2012).
These forces are also known as ‘fictitious forces’ or ‘pseudo-forces’ because they result not from any
physical interaction but from the acceleration of the non-inertial reference frame, i.e. the rotating Earth (Iro,
2010; Price, 2006). The centrifugal force is strongest at the equator whereas the Coriolis force is weakest at
the equator as far as horizontal flows are concerned. Storetvedt (1992, p. 217) wrote that the mobility of the
lithosphere relative to the underlying mantle was probably caused by ‘thermal convection in the mantle, …
controlled by the Earth’s axial spin’. Later, however, he argued that ‘mantle convection is no longer needed
as a driving force in tectonics’ (Storetvedt, 2007). It is therefore unclear how the Coriolis force is supposed
to spin entire continents, even a continent as big as Eurasia. The Earth’s rotation will cause overall eastward
flow in the asthenosphere (and also in lithospheric magma channels) (Meyerhoff et al., 1992b, 1996a), but it
is hardly conceivable that inertial forces could alter asthenospheric flow in a way that would cause
individual continents/‘plates’ to rotate. Wrench tectonics has yet to back up its rather inflated claims about
inertial forces with a quantitative analysis. The forces that have supposedly shifted South America (and the
surrounding seafloor) 20o southward, and the ‘Antarctica-Australia-New Zealand-Melanesia block’ 15o
northeastward are also unknown.
The rotation poles can only be in one place at a time. So once polar wanderers who reject large-scale plate
motions but take (some) palaeomagnetic data seriously, have decided on the poles’ location for a particular
period, remaining components of palaeomagnetic data must be either explained by rotations and translations
of individual continents or continental blocks, or dismissed as errors.2 If wrench tectonics were to
incorporate more regional palaeomagnetic data, more and more movements of intracontinental blocks would
be required: for instance, the North and South China blocks, Siberia and ‘Baltica’, the northern and southern
halves of Britain and Ireland, and hundreds of ‘exotic terranes’ will all have to undergo independent
rotations/translations, and there is no guarantee that these rotations will be in the same direction as that in
which the continent as a whole is supposedly rotating. If the postulated movements are geologically
unrealistic, and if it is necessary to be arbitrarily selective in deciding which palaeomagnetic data to use, it is
An extreme example of this is Pavlenkova’s (2012) fluid-rotation model, in which she adopts Storetvedt’s polar
wander scenario but rejects the proposed rotational and translational movements of continents/plates. She only accepts
the overall northward continental shifts implied by palaeomagnetic data, but arbitrarily rejects any longitudinal shifts.
She ignores the fact that even the latitudinal changes implied by palaeomagnetic data for different continents or blocks
are not always compatible with the proposed 70 o of polar wander since the Carboniferous.
2
58
probably a sign that the data themselves are unreliable guides to past positions of the rotation poles.
Palaeoclimate
The boundaries of the world’s present climate zones do not run parallel to the equator and are not perfectly
axisymmetric (fig. 7.1). This is due to the distribution of land and sea, continental topography and oceanic
bathymetry, and related atmosphere and ocean circulation.
Fig. 7.1. World climate zones. (http://en.wikipedia.org)
Over geologic time, the Earth has passed through numerous cycles of warming and cooling, on different
time scales, during which the width of climate zones has varied significantly. Global cycles of climate
change and the varying width of climate zones help to explain the presence of large Cretaceous dinosaurs
and trees in high-latitude localities such as Svalbard and the North Slope of Alaska, late Palaeocene-midEocene forests on Ellesmereland with crocodilian bones, palm trees in west-central Greenland and southern
Alaska, nummulitic (Tethyan) limestone on the Hatton-Rockall Plateau, mangrove swamps in the LondonParis basin, and large fossil trees and coal seams within 3º of the south pole. Even since the early Pliocene
the width of the temperate zone has changed by more than 15° (1650 km) in both the northern and southern
hemispheres (Meyerhoff et al., 1996b). High-resolution proxy data confirm that the Earth as a whole passes
through warmer and cooler periods, as well as highlighting regional palaeoclimate variations. The ‘big
picture’ is shown in figures 7.2 and 7.3.
The global climate is a complex, chaotic, nonlinear system and our understanding of how it works is still in
its infancy. There is huge uncertainty regarding the relative importance of, and dynamic interaction between,
solar factors, orbital factors (obliquity, eccentricity, precession), oceanic factors, tectonic factors, positive
and negative feedbacks, and the hydrological and carbon cycles. Major uncertainty surrounds the dynamic
role of clouds in modulating the Earth’s temperature.
59
Fig. 7.2. Global temperature since the late Proterozoic (Scotese et al., 1999, fig. 2; revised by C.R. Scotese). Cenozoic
temperature is based on oxygen-isotope ratios, and pre-Cenozoic temperature on lithological climate indicators (coals,
evaporites, bauxites, tillites, etc.).
Fig. 7.3. Phanerozoic temperatures based on oxygen-isotope data. (For sources, see: http://en.wikipedia.org.)
The obliquity of the ecliptic determines the latitudinal distribution of insolation, the range of seasonal
changes, and the widths of the tropical and polar zones. The higher the obliquity, the greater the seasonal
contrast, especially at high latitudes. The obliquity is currently 23.44º and is slowly decreasing; over the past
5 Myr it is calculated to have varied between 22.08º and 24.54º, with a mean period of about 41,000 years
(Berger & Loutre, 1991). A similar oscillation is generally assumed to have applied for most of the Earth’s
history, but some workers postulate far greater changes in axial tilt. Williams (1993) argued that the
obliquity was above 54° for most of the Precambrian, and decreased rapidly from about 60° to 26° between
650 and 430 Ma. Some modelling studies suggest that high obliquity (up to 70o) throughout the Precambrian
could help to explain warm temperatures during the Archaean and/or at least some of the Proterozoic
glaciations (Jenkins, 2000, 2004; Donnadieu et al., 2002), though explanations assuming a lower obliquity
have also been proposed (Hoffman et al., 1998).
To explain the generally warmer and more equable climate of the Mesozoic and early Cenozoic, several
workers have suggested that the obliquity may have been between 0º and 15º (Douglas & Williams, 1982;
Xu Qinqi, 1979, 1980; Allard, 1948). On the basis of palaeobotanical evidence, Wolfe (1978, 1980) argued
60
that the obliquity decreased gradually from around 10° to 5° from the Palaeocene to the mid-Eocene, then
began to increase slightly until the end of the Eocene, when it increased rapidly to 25-30°, before decreasing
again. Barron (1984) argued that although a smaller obliquity would increase winter insolation at high
latitudes, the mean annual insolation would decrease, leading to cooler polar temperatures, whereas the
evidence points to warmer polar temperatures in Mesozoic and early Cenozoic time. Wolfe (1978), however,
suggested that at some critical value of axial inclination, the atmospheric circulation changes from one that is
predominantly cellular (as it is today) to one that is predominantly meridional, which would more than
compensate for decreased annual insolation at high latitudes.
About 60% of the northern hemisphere is land, compared with only 20% in the southern hemisphere. Since
water has a greater heat capacity than land, temperature gradients and climate extremes are greatest in the
northern hemisphere, while the southern hemisphere has a more uniform annual climate. The greater land
area in the northern hemisphere helps explain why the meteorological equator is shifted north of the equator
much of the time over much of the globe. Marked by the Intertropical Convergence Zone (ITCZ), it is north
of the equator during the northern-hemisphere summer months and moves south of the equator only in some
areas during the southern-hemisphere summer.
Ocean currents, which are closely linked to wind systems, are a vital factor in understanding regional climate
(figs. 7.4 and 7.5). They explain why the average temperature of the North Atlantic at 50-60ºN is nearly 7ºC
higher than that of the South Atlantic at 50-60ºS, and the whole of the North Atlantic between 30ºN and the
Arctic circle is on average about 5º warmer than the South Atlantic between 30ºS and Antarctic circle. Twothirds of the Atlantic equatorial current are deflected into the northern hemisphere and only one-third into
the southern hemisphere. Non-drift palaeogeographic reconstructions show that, during most of geologic
time, virtually all the equatorial currents were deflected into the northern hemisphere (Simpson et al., 1930).
Fig. 7.4. Surface ocean currents. (www.physicalgeography.net)
Knowledge of the past distribution of land and sea, and past atmosphere and ocean circulation patterns is
very limited. While we have reasonable knowledge of marine inundations of the present continental areas,
the task of gathering sufficient data to determine the size, distribution and evolution of former landmasses in
the present oceans has barely begun (see figs. 2.2 and 2.4); palaeoclimatic and palaeontological data from
such palaeolands are nonexistent.
61
Fig. 7.5. The thermohaline circulation, or ‘great ocean conveyor’. (en.wikipedia.org)
Dickins (1994a) proposed that a land barrier to the west of North and South America would explain the
anomalous warm water temperatures along the ‘Andean’ and ‘Rocky Mountain’ seaways in the Lower
Permian (fig. 7.6); data supporting the existence of such a barrier is presented by Dickins et al. (1992).
Brooks (1949) argued that the distribution of full-grown and dwarf rudist molluscs in the Upper Cretaceous
could be explained by a different ocean current system, arising from a different arrangement of land and sea
(fig. 7.7).
Fig. 7.6. Postulated ocean current systems to explain water-temperature distribution in the Lower Permian. (Dickins,
1994a, fig. 9)
62
Fig. 7.7. Geography of the Upper Cretaceous, showing land and sea distribution, probable ocean currents, and
distribution of reef-building rudists. (Brooks, 1949, fig. 28)
Continental and polar stability
The geographic distribution of palaeoclimatic indicators such as evaporites, carbonate rocks, coals, and
tillites since Proterozoic time is best explained by stable rather than shifting continents, in conjunction with
periodic changes in climate, from globally warm or hot to globally cool (Meyerhoff & Meyerhoff, 1974a;
Meyerhoff et al., 1996b). Shifting the continents or poles may explain local or regional palaeoclimatic
features for a particular period, but it invariably fails to explain the global climate for the same period.
Evaporites are commonly thought to form in warm or hot climates where water evaporation exceeds water
influx, though there is very good geochemical evidence that factors other than climate are also extremely
important in evaporite formation (Hardie, 1990, 1991). Meyerhoff (1970a,b) found that 95% of all
evaporites from the late Proterozoic to the present, by volume and by area, lie in regions which today receive
less than 100 cm of annual rainfall, i.e. in today’s dry-wind belts. At least 35% of these evaporites are prePermian. Late Palaeozoic evaporites are associated with areas where reefs and fusulinids thrived, and were
generally deposited next to areas where warm oceanic currents were present, even at high latitude, in late
Proterozoic and early Palaeozoic times (Meyerhoff et al., 1996b).
On the basis of evaporite distribution, Meyerhoff (1970b) compiled a graph showing periods of worldwide
warming and cooling (fig. 7.8). The vertical coordinate is geologic time, and the horizontal coordinate is
absolute degrees of latitude across which evaporites were deposited. Right-hand deflections of the curve
represent periods when the Earth was uniformly warm, and are called evaporite-maximum periods (e.g. late
Proterozoic-Cambrian, Devonian, and Permo-Triassic). At such times only two climatic zones are present: a
broad torrid zone (90-120o wide), and temperate zones in the polar regions. Left-hand deflections of the
curve represent periods when the Earth was cool, and are called evaporite-minimum or glacial-maximum
periods. At such times there are three climatic zones (as today): a torrid zone (60-70o wide), two temperate
zones at middle latitudes, and frigid zones in the polar regions. Note the overall correspondence of the curve
in fig. 7.8 with the curves in figures 7.2 and 7.3.
These different climate states are reflected in the meridional thermal gradient (ΔT), which is partly related to
63
equator-to-pole heat transport efficiency. Today, during a relatively warm period in an otherwise cold
Cenozoic era, ΔT is about 33ºC. During the globally warm Mesozoic climate, ΔT was about 19-23ºC, and
during the nearly as warm Palaeocene-Eocene Thermal Maximum (55 Ma) it was 15ºC, whereas during the
last glacial maximum (21-22 ka) it was about 50ºC (Cronin, 2010).
Fig. 7.8. Evaporite-maximum and evaporite-minimum
periods. (Meyerhoff & Meyerhoff, 1974a, fig. 2)
Fig. 7.9. Map showing Permian
evaporite, coal, and tillite
distributions. X = coal; solid
black = evaporites; solid
triangles = tillites;
predominantly coal areas are
separated from predominantly
evaporite areas by heavy black
lines (note northward deflection
in area of modern Gulf StreamNorth Atlantic Drift). Present
warm ocean currents are shown
by single-line arrows, and cold
ocean currents by double-line
arrows. Horizontally ruled areas
are those areas which today
receive more than 100 cm of
annual rainfall. (Meyerhoff,
1973, fig. 2; reprinted with
permission from University of
Chicago Press)
64
Fig 7.10. Map showing
Triassic evaporite, coal,
and aeolian sandstone
distributions. The
Triassic was an
evaporite-maximum
period; note the great
width of the evaporite
zone. Symbols are the
same as for fig. 7.9, with
two exceptions: (a) there
are no known Triassic
tillites, and (2) aeolian
and partly aeolian
sandstone deposits are
shown by stippled
pattern. (Meyerhoff,
1973, fig. 3; reprinted
with permission from
University of Chicago
Press)
During most times after the Devonian, two axisymmetrical globe-encircling belts of coal deposits were
present, one lying north of the evaporite belt, and the other south of it. A third – tropical – coal belt existed at
times, especially during the Cenozoic. 88% of the world’s economic coal deposits are on the eastern sides of
the continents or in northwestern Europe and on the Arctic coast of northwestern Asia – i.e. in the regions
which today receive the heaviest rainfall in the temperate zones. Northwestern Europe and the adjacent part
of Arctic Asia are exceptions because only in the North Atlantic do a major moisture-bearing wind system
and warm ocean current, the Gulf Stream, cross from the eastern side of a continent (North America) to the
northwestern side of another (Eurasia) (Meyerhoff et al., 1996b; Meyerhoff & Teichert, 1971). It is
noteworthy that both the evaporite and coal zones show a pronounced northward offset similar to today’s
northward offset of the meteorological equator.
Meyerhoff (1970a,b) presents many maps showing the general axisymmetry of Devonian to Miocene
evaporite and coal deposits. He notes that the axisymmetry of high-latitude coal belts and the low-latitude
evaporite belts is not so apparent for pre-Triassic and Miocene times, and identifies several reasons for this,
connected with the location of orogenic belts, the uneven worldwide distribution of Palaeozoic rocks, and
the history of the North Atlantic and Arctic Oceans.
He notes that high-latitude evaporites – late Proterozoic through early Permian – are widespread in Canada
and Eurasia, whereas evaporite deposition of the same ages in the southern hemisphere remained nearly
constant (20-40ºS) at all times during post-Proterozoic history. Almost all ancient high-latitude evaporite
deposits are associated with marine sequences deposited in seas which invaded the continents from the
Eurasian-Arctic basin and from the North Atlantic, but not from the Canadian-Arctic basin, which is
separated from the former by the Proterozoic Lomonosov sill and from the North Pacific by the Archaean
Bering-Chukotsk Shelf. After Devonian time, evaporite depocentres shifted systematically Atlanticward
with the formation of the Franz Josef and Faeroes-Greenland sills. The required temperature for late
Proterozoic-Palaeozoic evaporite deposition in high latitudes during evaporite-maximum periods can be
attributed to the persistence of the Gulf Stream-North Atlantic Drift system since mid-Proterozoic time.
Meyerhoff traces the history of sill development across the Arctic and North Atlantic Oceans in great detail
and states that this provides the only known explanation for high northern latitude evaporite deposition in the
past, and the gradual shift of evaporite deposition southward.
65
Fig 7.11. Latitudinal distribution of ammonoids, carbonate rocks, terrigenous clastic rocks, and tetrapod faunas for
Mesozoic time. 1 = northern and southern boundaries of carbonate rocks; 2 = southern limit of Boreal and some Tethys
and Pacific ammonoids; 3 = central part of carbonate-rock belt (presumed meteorological equator). (Khudoley, 1974,
fig. 1)
66
Khudoley (1974) found that the latitudinal distribution of carbonate rocks, terrigenous clastic rocks,
ammonoids (extinct group of molluscs), and tetrapod faunas for Mesozoic time was consistent with the
present positions of the continents and geographic poles (fig. 7.11). Carbonate rocks extended on average to
40-45ºN and 30ºS, showing the pronounced northward offset of the thermal equator. During the Mesozoic,
the carbonate belt decreased in width from about 90º in the Triassic to 65º in the Cretaceous. The average
northward offset of the median line of carbonate rocks was about 24º in the Triassic, 19º in the Jurassic, and
12º in the Cretaceous, reflecting global climatic changes.
Permo-Carboniferous glaciations
In the Carboniferous and Permian, glaciers covered parts, if not all, of Antarctica, the Malvinas Islands,
South Africa, South America, India, and Australia. Proponents of continental drift claim that this glaciation
can be explained in terms of Gondwanaland, which was then situated near the south pole. However,
glaciation requires the interaction of warm ocean currents, moisture-laden warm air, and cold winds
generated by glacial ice (Coleman, 1925, 1932; Brooks 1949). Moisture to sustain continental glaciers
cannot be carried more than 2500 km, whereas the central parts of the proposed supercontinents are 3000 to
4000 km from the nearest ocean moisture source. In all Laurasia-Gondwanaland reconstructions, not even
one-tenth of this amount of moisture could have reached the supercontinents’ interiors (Meyerhoff &
Harding, 1971). Glaciers would have formed only at the margins of Pangaea, while the interior would have
been a vast, frigid desert, like parts of interior Siberia today (Meyerhoff, 1970a; Meyerhoff & Teichert,
1971). Shallow epicontinental seas within Pangaea could not have provided the required moisture because,
like Hudson Bay today, they would freeze over during winter months, thereby preventing evaporation.
During the late Ordovician, too, large areas of Gondwanaland and Laurasia that were glaciated would have
been too far inland for moist ocean-air currents to reach them.
Fig. 7.12. Locations of original Carboniferous and Permian glacial centres – most of them in mountain highlands.
Arrows show interpreted movement directions. Except possibly in Antarctica, no glacial centre was more than a quarter
of the size of the Pleistocene Keewatin ice sheet of North America. (Meyerhoff & Teichert, 1971, fig. 3; reprinted with
permission from University of Chicago Press)
The Permo-Carboniferous glaciations are easier to explain in terms of the continents’ present positions:
nearly all the continental ice centres were adjacent to or near present coastlines, or in high plateaus and/or
mountainlands not far from the present coasts. Some of southern and most of central Africa, much of the
Andes, and parts of Australia – as well as possibly large areas in India and Brazil – were the sites of welldeveloped mountain glaciation. There were more large glacial centres in the southern hemisphere than in the
67
northern hemisphere because the average elevation of the continents during Permo-Carboniferous time was
greater in the southern hemisphere. With the exception of parts of northern and eastern Siberia and the
present Urals, low-lying lands warmed by epeiric seas characterized the northern hemisphere. The change
from high elevations in the southern hemisphere (vs. low elevations in the northern hemisphere) during
Permo-Carboniferous time to high northern-hemisphere elevations (vs. low southern-hemisphere elevations)
during Pliocene-Holocene time, together with the progressive southward development of sills in the Arctic
and North Atlantic, provide a sound explanation for changes in locale of continental glaciation from PermoCarboniferous to Pliocene-Holocene time. There is no need to invoke continental drift or polar wander.
Fig. 7.13. Gondwanaland reconstruction for Permo-Carboniferous time, showing icecap centres. Icecaps are separate
and distinct, and interior icecaps (in eastern Brazil, Africa, Malagasy, and parts of Australia) are 3000-4000 km from
the coast, well out of reach of moisture-bearing winds. (Meyerhoff & Teichert, 1971, fig. 13; reprinted with permission
from University of Chicago Press)
Moisture was also needed to nourish luxuriant growth of the type known in Carboniferous and younger
sediments of Gondwanaland. The extensive Carboniferous, Permian, and younger coal deposits of interior
Gondwanaland and Laurasia would have required 150-200 cm of annual rainfall, so these, too, could not
have formed if the continents had been joined together; deserts, not swamps, would have been present
instead (Meyerhoff & Harding, 1971; Meyerhoff et al., 1996b).
Brooks (1949) argued that the distribution and elevation of landmasses in the late Carboniferous (without
assuming drift or polar wander), and the prevailing wind and ocean currents, could explain why ice sheets
developed in low latitudes while a comparatively mild climate existed further north. At that time the
diversion of heat from the southern to the northern hemisphere is thought to have been greater than at any
subsequent time. Brooks postulated a wide funnel opening to the Pacific, continuing in subtropical latitudes
along the Tethys Sea into the Mediterranean and through it into the Atlantic. In this configuration, the Tethys
Sea would have received all the warm equatorial water of the Pacific (Brooks, in Simpson et al., 1930).
Brooks argues that in the mid-Carboniferous Gondwanaland was not continuous from South America to
Australia, but was probably broken up into three or four separate landmasses by straits leading from north to
south. This would allow free circulation between the Tethys Sea and the Southern Ocean, considerably
raising the temperature of the southern hemisphere and helping to account for the great climatic difference
between the Middle and Upper Carboniferous.
68
Fig. 7.14. Geography of the Upper Carboniferous according to Brooks (1949, fig. 29). In the Lower Carboniferous
the areas of Gondwanaland drawn with broken lines were probably under water.
Eocene-Oligocene boundary
According to wrench tectonics, around the Eocene-Oligocene (E-O) boundary (33.9 Ma), the poles were
displaced 35º to approximately their present locations within 2 to 3 Myr (Storetvedt, 1997, 2003). This is
said to explain the drastic cooling at that time, which put an end to the warm conditions in northern Europe,
Iceland and Greenland. The poles are said to have shifted back to their previous position in the midMiocene, explaining why temperatures in Europe reached a new maximum and why the eastern
Mediterranean was again characterized by relatively flat-lying palaeomagnetic inclinations. Finally, the
poles are said to have shifted back to their present position around 5 Ma (Storetvedt & Bouzari, 2012).
Storetvedt cites Buchardt (1978) as evidence for the abrupt polar wandering events at the E-O boundary and
in the mid-Miocene. Buchardt’s isotopic palaeotemperature curve (fig. 7.15) is based on benthic isotopes
from the southern North Sea. This was the first such curve for the Tertiary period in northwestern Europe.
Higher-resolution proxy data, both global and regional, have since become available.
Fig. 7.15. Isotopic temperature curve for the Tertiary North Sea. Uncertainty is indicated by shaded area. (Buchardt,
1978)
69
Zachos et al.’s (2001) benthic foraminiferal oxygen-isotope (δ18O) curve for the Cenozoic is shown in fig.
7.16. Its variations reflect episodes of global warming and cooling, and ice-sheet growth and decay. The past
65 Myr have seen progressive planetary cooling, but with many fluctuations. A series of warmer intervals
(e.g. the Middle Eocene Climatic Optimum, the Late Oligocene Warming Event, and Mid-Miocene Climatic
Optimum), and colder intervals (e.g. the Oi and Mi glaciations) have been recognized and correlated in both
hemispheres.
Fig. 7.16. Global deep-sea oxygen isotope record based on data from over 40 DSDP and ODP sites. (Zachos et al.,
2001, fig. 2).
Fig. 7.16 shows a temperature decrease of over 4ºC at the E-O boundary and a similar temperature increase
in the late Oligocene. Zachos et al. (2001) say that the entire increase in δ18O prior to the late Eocene can be
attributed to a drop in deep-sea temperature from ~12° to ~4.5°C. All subsequent δ18O change reflects the
combined effect of ice-volume and temperature, particularly the rapid step at 34 Ma. Following the cooling
70
and rapid expansion of Antarctic continental ice-sheets in the earliest Oligocene, deep-sea δ18O values
remained relatively high, indicating permanent ice sheets, with a mass of up to 50% of that of the presentday ice sheet, and bottom temperatures of ~4°C. These ice sheets persisted until the latter part of the
Oligocene (26 to 27 Ma), when a warming trend reduced the extent of Antarctic ice. From this point until the
mid-Miocene, global ice volume remained low and bottom water temperatures trended slightly higher, with
the exception of several brief periods of glaciation (e.g. the Mi events). This warm phase peaked in the latemiddle Miocene climatic optimum (17 to 15 Ma), and was followed by a gradual cooling and
reestablishment of a major ice-sheet on Antarctica by 10 Ma.
Small ephemeral ice-sheets appeared on Antarctica throughout the late Eocene. In the earliest Oligocene a
climatic threshold was apparently reached, allowing for the rapid expansion of large ice-sheets, resulting in
the deep but short-lived Oi-1 glaciation. Regional sea surface temperatures decreased by over 4ºC and deep
waters cooled by 3-4ºC (Salamy & Zachos, 1999). The Oi-1 event lasted 400 kyr. It involved the
reorganization of the climate/ocean system, as shown by global shifts in the distribution of marine biogenic
sediments and an overall increase in ocean fertility, and by a major drop in the calcium carbonate
compensation depth (Zachos et al., 2001). High-resolution isotopic records show that the ubiquitous δ18O
increase marking the Oi-1 event occurred in less than 350 kyr, with over half the transition occurring in the
final 40-50 kyr. This period of lower temperatures and widespread continental glaciation persisted for
roughly 400 kyr. This interval was characterized by at least two ~100-kyr waxing and waning cycles (Oi-1a
and Oi-1b) and possibly several higher-frequency events (Zachos et al., 1996).
De Man & Van Simaeys’ (2004) benthic foraminiferal palaeotemperature curve for the Oligocene southern
North Sea Basin does not cover the E-O boundary but it does show an abrupt temperature increase of ~25ºC
at the boundary of the Rupelian and Chattian stages (28.1 Ma) of the Oligocene. In general, the Rupelian
assemblages are dominated by cold to cold-temperate taxa and the calculated bottom-water
palaeotemperatures never exceed 10°C. The base of the Chattian, on the other hand, is characterized by
abundant warm temperate, tropical to subtropical taxa and the calculated bottom-water palaeotemperatures
exceed 20°C. Higher up the Chattian, the warm temperate to subtropical species are less abundant, and coldtemperate taxa become more frequent. Previous studies – using independent tools of palaeoclimate
reconstruction – confirm that warm to tropical conditions emerged during the late Oligocene in Northwest
Europe.
The wrench-tectonic attempt to explain certain climatic changes during the Cenozoic in terms of polar
wander is ill-conceived. Moving the poles will bring about global shifts in the locations of climate zones,
with some areas becoming warmer and some cooler. It will not automatically bring about global warming or
cooling. The rapid polar wander event that began around the E-O boundary allegedly triggered the Oi-1
glaciation, but the glaciation ended within a million years, while the poles were supposedly still wandering
to their present positions. A major abrupt warming event took place in the late Oligocene, but the next
climate transition that wrench tectonics singles out as being linked to polar wander is the mid-Miocene
warming. More polar wander events could be added, but it would still be necessary to invoke other factors to
explain global climate changes, and the more the poles are moved back and forth like a yoyo, the more
implausible the proposed scenario becomes. Moreover, since wrench tectonics claims to have solid
palaeomagnetic foundations, global palaeomagnetic data supporting these additional events would have to be
found – though this could be facilitated by the tactic of dismissing palaeopole positions or age
determinations that are inconvenient.
The following table shows published sea surface temperature (SST) estimates from marine records for the
Lomonosov Ridge (87.87ºN) (O’Regan et al., 2011). The first two entries are for Palaeocene/Eocene time,
the next nine are for Eocene time, and the last two are for Miocene time.
Age
(Ma)
≤55
55
≤53.5
53.5
49
49
48
46
45
44.4
44.4
18
18
SST
(ºC)
17.5
23
22
26.5
25
9
13.5
15
8.2
4.7
10
19.7
13
71
From a wrench-tectonic perspective, the Lomonosov Ridge site would have been located at about 55ºN or
less for all these ages except the last two, when it would presumably have been at or near its present
location. A sudden 35º latitude shift around 35 Ma does not help to explain these data.
There is evidence of cooling at all latitudes during the E-O transition. Coxall & Pearson (2007) reported
cooling of low-, middle-, and high-latitude continents. On the basis of proxy records of sea surface
temperatures from multiple ocean localities, Liu et al. (2009) concluded that high-latitude sea surface
temperature cooled by an average of ~4.8ºC during the E-O climate transition (37-33 Ma). Greenland
experienced gradual long-term cooling of ~3-5°C starting near the E-O boundary (Schouten et al. 2008). In
the Norwegian-Greenland Sea, there was a cooling of ~5ºC across E-O boundary (O’Regan et al., 2011). In
the interior of North America, the Eocene mean annual air temperature was stable but a cooling of 8±3°C
over about 400 kyr occurred during the early Oligocene (Zanazzi et al., 2007). The oxygen isotope values in
low-northern-latitude nearshore marine fish otoliths (ear stones) and mollusc shells show little reduction in
summer or mean annual temperatures, but significant lowering of winter temperatures after the end of the
Eocene (Ivany et al., 2000; Kobashi et al., 2001). Although global cooling was particularly notable at high
latitudes, there was also a ~2.5°C decrease in tropical surface-water temperature (Lear et al., 2008).
There is evidence of ice-sheet growth on Antarctica as early as 42-38 Ma (Tripati et al., 2005). The
conventional view used to be that ice-sheet growth in the northern hemisphere began no earlier than about 15
Ma. But ice-rafted debris points to ice on Greenland as early as 44-38 Ma (Tripati et al., 2008; Eldrett et al.
2007), and there is evidence of episodic sea-ice formation in the Arctic around 46 Ma (St. John, 2008;
O’Regan et al., 2011). At that time, central Greenland (72oN, 40oW) would have been at a latitude no higher
than 40oN according to wrench tectonics.
The Cenozoic global ocean was initially relatively weakly stratified, with relatively warm surface and
bottom temperatures, but became strongly stratified, with warmer surface temps and cooler deep-sea
temperatures separated by strong vertical thermal gradients (Cronin, 2010). A progressive 12ºC cooling
occurred in four steps: in the early to mid-Eocene, at the E-O boundary, during the late-middle Miocene, and
in the Plio-Pleistocene. These patterns generally match climate trends inferred from deep-sea foraminiferal
isotopes and other marine and continental records.
The opening and closing of ocean gateways are recognized as having a major impact on ocean circulation
and heat transport. Antarctica is currently isolated from the rest of the world by the strong Antarctic
Circumpolar Current (ACC). The isolation of Antarctica and the development of strong zonal (west-to-east)
flow by the ACC are the result of the opening of two major ocean gateways – the Tasmanian Gateway
between Australia and Antarctica, and the Drake Passage between South America and Antarctica. The
opening of the Tasmanian Gateway in particular has been linked to the E-O climate transition. The opening
and widening of these gateways are usually attributed to plate motions and seafloor spreading, but vertical
tectonics will do the job just as well.
The Tasmanian land bridge subsided sufficiently to create a shallow-water opening in the Tasmanian
passage region in the earliest Oligocene, when a shallow ACC was established in this region. The deepwater (2000 m) passage may have been fully established during or just after 34 Ma. The early Tertiary land
connection between Antarctica and South America also underwent gradual subsidence, and the Drake
Passage gradually opened. During the late Eocene (~50 Ma), reconstructions show shallow-water pathways
across subsiding continental shelf areas, and by 34-30 Ma a continuous (1000-3000 m) deeper-water
connection had formed, establishing full circumpolar oceanic flow around the entire Antarctic continent
(Cronin, 2010). These events prevented meridional oceanic heat flow from low to high southern-hemisphere
latitudes, led to the thermal isolation of Antarctica, fostered the growth of the Antarctic ice sheet, and
initiated thermohaline circulation similar to that of today (Kennett & Shackleton, 1976; Kennett & Exon
2004).
72
Palaeobiogeography
Meyerhoff et al. (1996b) showed in a detailed study that most major biogeographical boundaries, based on
floral and faunal distributions, do not coincide with the partly computer-generated plate boundaries
postulated by plate tectonics. Nor do the proposed movements of continents correspond with the known, or
necessary, migration routes and directions of biogeographical boundaries. In most cases, the discrepancies
are very large, and not even an approximate match can be claimed. The authors comment: ‘What is puzzling
is that such major inconsistencies between plate tectonic postulates and field data, involving as they do
boundaries that extend for thousands of kilometers, are permitted to stand unnoticed, unacknowledged, and
unstudied’ (p. 3). A group of earth science graduates who were asked to comment on the authors’ manuscript
commented: ‘If this global study of biodiversity through time is correct, and it is very convincingly
presented, then a lot of what we are being taught about plate tectonics should more aptly be called
“Globaloney” ’ (p. ix).
Meyerhoff et al. (1996b) also demonstrated the existence during most if not all of Phanerozoic time of a
broad geographical zone, ranging from several hundred to 5000 km in width, where strata bearing ‘northern’
biotas are intercalated with strata bearing ‘southern’ biotas and, in many areas, admixtures of northern and
southern taxa are present in the same beds. The warm realms of northern origin include the post-Palaeozoic
Tethyan Realm, while the cooler, southern realms include the mid-Palaeozoic Malvinokaffric Realm and the
late Palaeozoic and younger Gondwana Realm.
Figures 8.1 and 8.2 demonstrate that the faunal and floral distributions of mid-Cambrian to early Cretaceous
times are nearly identical with those of Cenozoic time. The northern and southern limits of the ‘intercalary
zone’, and the northern and southern limits of specific groups of organisms shown in fig. 8.2, demonstrate a
distinct bipolarity in the distribution of mid-Cambrian to early Cretaceous organisms and are explained most
simply in terms of north-south migrations on a modern globe, without any large-scale continental drift or
polar wander. During no interval of the past is there good evidence for a perfectly latitudinal biotic
distribution, owing to a variety of geographic and climatic barriers.
Fig. 8.1. Map showing the intercalary zone of Cambrian to early Cretaceous time (diagonal shading). This is the zone
where northern and southern biotas are mixed. The heavy line marks the northern known limit of the earliest Permian
marine fossil localities. The northern limit of the intercalary zone is the northernmost reported limit of southern taxa of
all ages and the southern limit is the southernmost reported limit of northern taxa of all ages. (Meyerhoff et al., 1996b,
fig. 17; reprinted with permission from the Geological Society of America)
In plate-tectonic terms, most of South America north of the Amazon River, an area of 3,300,000 km2, has
belonged biogeographically to the North American ‘plate’ since Ordovician time. Furthermore, an important
73
north-south biogeographic boundary occupied the present Andean region, and separated warm-water biotas
on the west from cool-water biotas on the east. The boundary first appeared in the Cambrian and persisted
into the Permian, shifting back and forth (east-west) with time. If plate-tectonic criteria were consistently
applied, a 7500 km suture zone would be postulated in the western part of South America, 100-300 km
inland from the Pacific Ocean. Plate models require a suture zone running the length of the Mediterranean
Sea, despite the stratigraphic continuity between Europe and Africa. But biogeographical data, beginning in
Devonian time, show a break only within the African continent, extending from near Dakar to Arabia
(Meyerhoff et al., 1996b).
Fig. 8.2. Map showing: (1) the northern limit of Malvinokaffric faunas during Ordovician time; (2) the northernmost
known occurrences of Gondwanan Triassic tetrapods; (3) the outline of early Permian ‘Gondwanaland’; and (4) the
southernmost reported occurrences of Triassic temperate and warm-water marine invertebrates. (Meyerhoff et al.,
1996b, fig. 18; reprinted with permission from the Geological Society of America)
In Asia and the Southwest Pacific, the boundary commonly used to separate the Malvinokaffric and
Gondwana Realms from their northern equivalents (e.g. Tethyan Realm in post-Palaeozoic time) is the
Indus-Yarlung ‘suture zone’. It extends over 5000 km across south-central and southeastern Asia alone, and
is projected westward to the Taurus Mountains of Turkey and Troodos massif of Cyprus, and southeastward
to Papua New Guinea. In many plate-tectonic models it forms the boundary between the northern and
southern plates. When Gondwanan taxa were found north of the suture, their identifications were first
ignored. Once some of them were accepted, the collision zone was shifted northward to embrace larger and
larger segments of Asia. Meyerhoff et al. (1996b) write:
Tethyan taxa are known from many places on the Indian craton and other places south of the suture, yet the suture’s
position is always shifted to accommodate Gondwanan taxa, never the reverse! When Tethyan forms are found south of
the suture, they are said to mark the northern shore of Gondwanaland. This leads to speculation on the fate of the
southern shore of Angaraland, yet this point is rarely discussed in the literature. (p. 8)
Similarly, when a representative Tethyan taxon is found in, say, a New Zealand setting, it becomes evidence
of an exotic terrane, but when the opposite situation is encountered – e.g. a Gondwanan taxon is found in
northern Tibet – the whole suture zone between the southern and northern continents is shifted northward for
whatever period of time is necessary.
The Taurus-Zagros-Indus-Yarlung boundary is not useful for the Malvinokaffric Realm because the biota
74
does not change across the suture. The boundary has been applied mainly to the younger Gondwana Realm.
But here, too, it is not useful because Gondwana elements (early Permian to early Cretaceous) extend
northward to the Tunguska basin of Siberia, Mongolia, northeastern China, the Parimor’ye region (north of
Vladivostok), and the Kolyma River basin of northeastern Siberia. Conversely, northern – and especially
Tethyan Realm – biotas extend southward to New Zealand, Western Australia, the Northern Territory of
Australia, southern India, and Saudi Arabia (Meyerhoff et al., 1996b). The widespread distribution of the
Glossopteris flora in the southern continents is frequently claimed to support the former existence of
Gondwanaland, but it is rarely pointed out that this flora has also been found in northeast Asia (Smiley,
1976).
As mentioned in a previous section, plate-tectonic models requiring a broad, deep ocean between India and
Asia are untenable because of the gradual, not abrupt, faunal changes that take place across the QinghaiXizang (Tibet) Plateau and the Himalayas. Particularly strong evidence of India’s proximity to Asia is
provided by the pre-Early Permian stratigraphy and biota. The most common plate-tectonic ‘explanation’
(which ignores pre-Permian problems) is that India drifted rapidly northward and attached itself to Asia
along a suture zone north of the present Indus-Yarlung suture in the earliest Permian. Later, Gondwanaland
collided with Asia in the late Permian, and then the Indus-Yarlung suture opened during the late Triassic and
India moved south – only to shuttle northward again and close the suture zone in the Eocene (Meyerhoff et
al., 1996b).
In his global analysis of Mesozoic ammonoids and associated sediments, Khudoley (1974, 1988) found
evidence of latitudinal zoning: a cooler (Boreal) ammonoid-clastic sediment realm around the present
rotation pole; a warmer (Tethyan) realm of circum-equatorial distribution; and some evidence for a cool
realm in high southern latitudes (Antiboreal) (see fig. 7.11). The boundaries of these temperature-controlled,
faunal-sediment realms shifted latitudinally as global climates changed. Khudoley concluded that the
ammonoid zonal belts and dispersal patterns contradicted large-scale continental drift and polar wandering.
Smiley (1967, 1974, 1976, 1992) argued that the later Palaeozoic, Mesozoic, and Cenozoic distributions of
terrestrial plants and vertebrate tetrapods, and of contemporaneous marine faunal realms, do not support
significant continental movements or polar wander. Global isotherm lines and latitudinal zonations of plants
and animals generally conform to the present rotational poles rather than past palaeomagnetic poles, and the
boundaries of past vegetation zones show a high degree of conformity with present isotherm lines. Smiley
acknowledges that specific data when taken out of their global context may lead to quite different
interpretations of global tectonics. Axelrod (1963, 1964), too, argued that floral evidence from late
Palaeozoic to late Mesozoic time suggested stable rather than shifting continents and poles.
Stehli & Grant (1971) examined the variation in the number of species within various groups of
Carboniferous and Permian faunas and demonstrated that the number increases systematically towards (or
away from) the present equator (fig. 8.3). Stehli (1957) concluded that the Permian north pole was at or close
to its present position. Smiley (1967) stated that if polar wander had occurred, vegetational sequences (at
least in the vicinity of the polar path) would be expected to show a cooling trend in one place as the pole
approached, and a warming trend in another as the pole receded, but this is not evident in the rich Mesozoic
floral records. Some polar wanderers have claimed that in the Lower Tertiary the equator ran along the
Mediterranean, placing Central Europe in the tropical rain belt (Storetvedt, 2003, p. 48-9). But early Tertiary
marine fauna were of tropical character in the London, Paris and Volga basins, in northwestern India,
southern Japan, South Africa, southwest Australia, southern New Zealand, and Patagonia, as well as in the
present tropics – consistent with the present position of the poles (Day & Runcorn, 1955).
More migration problems are raised by joining the continents in the past than by keeping them separated
(Simpson, 1943; Teichert & Meyerhoff, 1972; Teichert, 1974; Meyerhoff & Meyerhoff, 1974a). It is
unscientific to select a few faunal identities and ignore the vastly greater number of faunal dissimilarities
from different continents which were supposedly once joined. For the later Palaeozoic, on the basis of
ecotonal floras, Smiley (1992) argued that continental India was near the Angara and Cathaysia land floral
provinces of Eurasia, and the Australia-New Guinea continental block was near the Cathaysia province of
75
southeastern Eurasia. The positions of India and Australia at high southern latitudes in a classical Pangaea
reconstruction would result in oceanic separation so great that interchange of viable flora structures would be
impossible. The evidence of floral interchange between Eurasia and western North America implies the
presence of a land dispersal route across the Beringian region at least from later Palaeozoic time, whereas
plate tectonics requires an oceanic separation perhaps thousands of kilometres wide through this region prior
to the Cenozoic (Smiley, 1976).
Fig. 8.3. Map showing sampling stations in
which Permian genera endemic to the Tethys
have been found. Note the Tethyan belt’s
clear relationship to the present equator, the
extension of Tethyan forms to high latitudes
in the Ural Seaway, and the strong suggestion
of thermal asymmetry in the two hemispheres
with the southern-hemisphere Tethyan region
much less extensive than the northern. (Stehli
& Grant, 1971, fig. 4)
Some of the palaeontological evidence appears to require the alternate emergence and submergence of land
dispersal routes only after the supposed breakup of Pangaea. For example, mammal distribution indicates
that there were no direct physical connections between Europe and North America during late Cretaceous
and Palaeocene times (unless the land connections contained barriers to migration), but suggests a temporary
connection with Europe during the Eocene (Simpson, 1943; Meyerhoff & Meyerhoff, 1974a). Continental
drift, on the other hand, would have resulted in an initial disconnection with no subsequent reconnection.
A few drifters have recognized the need for intermittent land bridges after the supposed separation of the
continents (e.g. Du Toit, 1937; Tarling, 1982b; Briggs, 1987). Various oceanic ridges, rises, and plateaus
could have served as land bridges, as many are known to have been partly above water at various times in
the past (Schuchert, 1932; Willis, 1932; Meyerhoff et al., 1996b; Meyerhoff & Meyerhoff, 1974a). They
include the Faeroe-Greenland Ridge, the Mid-Pacific Mountains, the Cocos Ridge, the South Madagascar
Ridge, the Kerguelen-Gaussberg Ridge, the Rio Grande Rise-Walvis Ridge, and parts of the Mid-Atlantic
Ridge. On geological and palaeontological grounds, Gregory (1925, 1929, 1930) argued that there had been
numerous landmasses in the world’s oceans. Simpson (1943), however, challenged the necessity of so many
transoceanic continents for migration purposes. This does not alter the fact that there is mounting evidence
for the existence of extensive former landmasses in the present oceans, most of which had disappeared by
Miocene time.
Conclusion
Fossil magnetism in rocks can be affected by many factors, with the result that derived virtual magnetic
poles for specific periods show a wide scatter. The fundamental assumption that averaged palaeomagnetic
poles approximately coincided with past geographic poles is unproved. There is great scope for subjectivity
in the selection, processing and interpretation of palaeomagnetic data, and this is reflected in inconsistent
and often contradictory reconstructions of plate motions and true polar wander through time. The major
tenets of plate tectonics and the associated theories of continental fragmentation and assembly are
contradicted by a wealth of evidence. The wrench-tectonic theory of large-scale polar wander and ‘minor’
plate rotations and translations, based on selected palaeomagnetic data, is also open to serious objections.
Geological, geophysical, palaeontological and palaeoclimatic data do not require large-scale plate motion or
76
polar wander; they point to nondrifting continents and stable poles, with vertical tectonic movements causing
periodic changes in the distribution of land and sea.
Acknowledgements: I would like to thank Dong Choi, Takao Yano and Giovanni Gregori for their valuable comments.
I would also like to thank all the organizations, publishers and individuals who have allowed use of illustrations.
References
Agocs, W.B., Meyerhoff, A.A., & Kis, K., 1992. Reykjanes Ridge: quantitative determinations from magnetic
anomalies. In: Chatterjee, S., & Hotton, N., eds., New Concepts in Global Tectonics. Lubbock, TX: Texas Tech
University Press, p. 221-238.
Ahmad, F., 1990. The bearing of palaeontological evidence on the origin of the Himalayas. In: Barto-Kyriakidis, A.,
ed., Critical Aspects of the Plate Tectonics Theory. Athens: Theophrastus Publications, v. 1, p. 129-142.
Airy, G., 1860. Change of climate. Athenaeum, no. 7171, p. 384.
Allard, H.A., 1948. Length of day in the climates of past geological eras and its possible effects upon changes in plant
life. In: Murneek, A.E., & Whyte, R.O., eds., Vernalization and photoperiodism: a symposium. Waltham, MA:
Chronica Botanica, p. 101-119.
Anderson, D.L., 2007. New Theory of the Earth. Cambridge: Cambridge University Press, www.gps.caltech.edu.
Anderson, D.L., Tanimoto, T., & Zhang, Y., 1992. Plate tectonics and hotspots: the third dimension. Science, v. 256, p.
1645-1651.
Anderson, D.L., & Natland, J.H., 2005. A brief history of the plume hypothesis and its competitors: concept and
controversy. In: Foulger, G.R., Natland, J.H., Presnall, D.C., & Anderson, D.L., eds., Plates, Plumes, and
Paradigms. Geological Society of America Special Paper 388, p. 119-145.
Andrews, J.A., 1985. True polar wander: an analysis of Cenozoic and Mesozoic paleomagnetic poles. Journal of
Geophysical Research, v. 9, p. 7737-7750.
Artemieva, I.M., & Mooney, W.D., 2002. On the relation between cratonic lithosphere thickness, plate motions, and
basalt drag. Tectonophysics, v. 358, p. 211-231.
Aumento, F., & Loncarevic, B.D., 1969. The Mid-Atlantic Ridge near 45°N. III: Bald Mountain. Canadian Journal of
Earth Sciences, v. 6, p. 11-23.
Axelrod, D.I., 1963. Fossil floras suggest stable, not drifting, continents. Journal of Geophysical Research, v. 68, no.
10, p. 3257-3263.
Axelrod, D.I., 1964. Reply. Journal of Geophysical Research, v. 69, no. 8, p. 1669-1971.
Bachtadse, V., & Briden, J.C., 1990. Paleomagnetic constraints on the position of Gondwana during Ordovician to
Devonian times. In: McKerrow, W.S., & Scotese, C.R., eds., Palaeozoic Palaeogeography and Biogeography.
Geological Society of London, Memoir 12, p. 43-48.
Bacon, M., & Gray, F., 1971. Evidence for crust in the deep ocean derived from continental crust. Nature, v. 229, p.
331-332.
Barrell, J., 1914. The status of hypotheses of polar wanderings. Science, v. 40, p. 333-340.
Barron, E.J., 1984. Climatic implications of the variable obliquity explanation of Cretaceous-Paleogene high-latitude
floras. Geology, v. 12, p. 595-598.
Barron, E.J., Harrison, C.G.A., & Hay, W.W., 1978. A revised reconstruction of the southern continents. American
Geophysical Union Transactions, v. 59, p. 436-439.
Beaman, M., Sager, W.W., Acton, G.D., Lanci, L., & Pares, J., 2007. Improved late Cretaceous and early Cenozoic
paleomagnetic apparent polar wander path for the Pacific plate. Earth and Planetary Science Letters, v. 262. p. 120.
Begg, G.C., Griffin, W.L., Natapov, L.M., O’Reilly, S.Y., Grand, S.P., O’Neill, C.J., Hronsky, J.M.A., Poudjom
Djomani, Y., Swain, C.J., Deen, T., & Bowden, P., 2009. The lithospheric architecture of Africa: seismic
tomography, mantle petrology, and tectonic evolution. Geosphere, v. 5, no. 1, p. 23-50, doi:10.1130/GES00179.1,
http://geosphere.gsapubs.org.
Beloussov, V.V., 1990. Certain trends in present-day geosciences. In: Barto-Kyriakidis, A., ed., Critical Aspects of the
Plate Tectonics Theory. Athens: Theophrastus Publications, v. 1, p. 3-15.
Berger, A., & Loutre, M.F., 1991. Insolation values for the climate of the last 10 million years. Quaternary Science
Reviews, v. 10, p. 297-317.
Berry, E.W., 1927. The term Oligocene and certain climatic considerations. American Journal of Science, v. 13, p. 252256.
Berry, E.W., 1928. Comments on the Wegener hypothesis. In: Waterschoot van der Gracht, W.A.J.M. van, ed., The
Theory of Continental Drift: A symposium. Tulsa, OK: American Association of Petroleum Geologists, p. 194-196.
77
Besse J., & Courtillot V., 1991. Revised and synthetic apparent polar wander paths of the African, Eurasian, North
American and Indian Plates, and true polar wander since 200 Ma. Journal of Geophysical Research, v. 96, p. 40294050.
Besse, J., & Courtillot, V., 2002. Apparent and true polar wander and the geometry of the geomagnetic field over the
last 200 Myr. Journal of Geophysical Research, v. 107, no. B11, 2300, doi:10.1029/2000JB000050, 6 - 1-31.
Besse, J., & Courtillot, V., 2003. Correction to ‘Apparent and true polar wander and the geometry of the geomagnetic
field over the last 200 Myr’. Journal of Geophysical Research, v. 108, no. B10, 2469, 3 - 1-2,
doi:10.1029/2003JB002684.
Besse, J., Courtillot V., & Greff, M., 2011. Paleomagnetism, polar wander. In: Gupta, H.K., ed., Encyclopedia of Solid
Earth Geophysics. Springer, p. 945-954.
Bhat, M.I., 1987. Spasmodic rift reactivation and its role in pre-orogenic evolution of the Himalayan region.
Tectonophysics, v. 134, p. 103-127.
Biggin, A.J., Hinsbergen, D.J.J. van, Langereis, C.G., Straathof, G.B., & Deenen, M.H.L., 2008. Geomagnetic secular
variation in the Cretaceous Normal Superchron and in the Jurassic. Physics of the Earth and Planetary Interiors, v.
169, p. 3-19.
Bonatti, E., 1990. Subcontinental mantle exposed in the Atlantic Ocean on St Peter-Paul islets. Nature, v. 345, p. 800802.
Bonatti, E., & Crane, K., 1982. Oscillatory spreading explanation of anomalously old uplifted crust near oceanic
transforms. Nature, v. 300, p. 343-345.
Bonatti, E., & Honnorez, J., 1971. Nonspreading crustal blocks at the Mid-Atlantic Ridge. Science, v. 174, p. 13291331.
Bouin, M.-N., & Vigny, C., 2000. New constraints on Antarctica plate motion and deformation from GPS data. Journal
of Geophysical Research, v.105, p. 28,279-28,293.
Briggs, J.C., 1987. Biogeography and Plate Tectonics. Amsterdam: Elsevier.
Brinkmann, R., 1972. Mesozoic troughs and crustal structure in Anatolia. Geological Society of America Bulletin, v. 83,
no. 3, p. 819-826.
Brooks, C.E.P., 1949. Climate through the Ages: A study of the climatic factors and their variations. 2nd ed., London:
Ernest Benn, Dover reprint 1970.
Buchardt, B., 1978. Oxygen isotope palaeotemperatures from the Tertiary period in the North Sea area. Nature, v. 275,
p. 121-123.
Bucher, W.H., 1933. The Deformation of the Earth’s Crust. Princeton, NJ: Princeton University Press.
Bucher, W.H., 1956. Role of gravity in orogenesis. Geological Society of America Bulletin, v. 67, p. 1295-1318.
Budinger, T.F., & Enbysk, B.J., 1967. Late Tertiary date from the East Pacific Rise. Journal of Geophysical Research,
v. 72, no. 8, p. 2271-2274.
Bullard, E.C., Everett, J.E., & Smith, A.G., 1965. The fit of the continents around the Atlantic. In: A Symposium on
Continental Drift, Royal Society of London Philosophical Transactions, Series A, v. 258, p. 41-51.
Burchfiel, B.C., 1990. The continental crust. In: Moores, E.M., ed., Shaping the Earth: Tectonics of continents and
oceans. New York: W.H. Freeman & Co.
Butler, R.F., 2004. Paleomagnetism: Magnetic domains to geologic terranes. Electronic edition, www.pmc.ucsc.edu.
Butler, R.F., Gehrels, G.E., McClelland, W.C., May, S.R., & Klepacki, D., 1989. Discordant paleomagnetic data from
the Canadian Coast Plutonic Complex: regional tilt rather than large-scale displacement? Geology, v. 17, no. 8, p.
691-694.
Butler, R.F., Dickinson, W.R., & Gehrels, G.E., 1991. Paleomagnetism of coastal California and Baja California:
alternatives to large-scale northward transport. Tectonics, v. 10, p. 561-576.
Butler, R.F., Gehrels, G.E., & Baldwin, S.L., 2002. Paleomagnetism and geochronology of the Ecstall pluton in the
Coast Mountains of British Columbia: evidence for local deformation rather than large-scale transport. Journal of
Geophysical Research, v. 107, p. 3–1-13.
Calderwood, A.R., 1991, No suspect terranes; the case against paleomagnetic data. Geological Society of America,
Abstracts with Programs, v. 23, no. 5, p. 136.
Cambiotti, G., 2012. Modelling the earth: compressible viscoelastodynamics, gravitational seismology and true polar
wander (doctoral thesis). University of Milan.
Cambiotti, G., Ricard, Y., & Sabadini, R., 2011. New insights into mantle convection true polar wander and rotational
bulge readjustment. Earth and Planetary Science Letters, v. 310, p. 538-543.
Caputo, M., 1986. The Polfluchtkraft revisited. Bollettino Geofisica Teorica Applicata, v. 28. p. 199-214.
Carlson, D.H., Plummer, C.C., & McGeary, D., 2008. Physical Geology: Earth revealed. 7th ed., New York: McGrawHill.
Chadwick, P., 1962. Mountain-building hypotheses. In: Runcorn, S.K., ed., Continental Drift. New York & London:
Academic Press, p. 195-234.
Chatterjee, S., & Hotton, N., 1986. The paleoposition of India. Journal of Southeast Asian Earth Sciences, v. 1, p. 145-
78
189.
Choi, D.R., 1997. Geology of the oceans around Australia, parts I-III. New Concepts in Global Tectonics Newsletter,
no. 3, p. 8-13.
Choi, D.R., 1999. Geology of East Pacific: Middle America Trench. New Concepts in Global Tectonics Newsletter, no.
12, p. 10-16.
Choi, D.R., 2000. Subduction does not exist – from seismic data interpretation. New Concepts in Global Tectonics
Newsletter, no. 15, p. 9-14.
Choi, D.R., 2001. Submarine geology of the oceans around Australia. Himalayan Geology, v. 22, p. 81-90.
Choi, D.R., 2002. Deep earthquakes and deep-seated tectonic zones. Part 2: South America. New Concepts in Global
Tectonics Newsletter, no. 24, p. 9-21.
Choi, D.R., Vasil’yev, B.I., & Tuezov, I.K., 1990. The Great Oyashio Paleoland: a Paleozoic-Mesozoic landmass in the
northwestern Pacific. In: Barto-Kyriakidis, A., ed., Critical Aspects of the Plate Tectonics Theory. Athens:
Theophrastus Publications, v. 1, p. 197-213.
Choi, D.R., Vasil’yev, B.I., & Bhat, M.I., 1992. Paleoland, crustal structure, and composition under the northwestern
Pacific Ocean. In: Chatterjee, S., & Hotton, N., eds., New Concepts in Global Tectonics. Lubbock, TX: Texas Tech
University Press, p. 179-191.
Coleman, A.P., 1925. Permo-Carboniferous glaciation and the Wegener hypothesis. Nature, v. 115, no. 2895, p. 602.
Coleman, A.P., 1932. Glaciation and continental drift. The Geographical Journal, v. 79, no. 3, p. 252-255.
Colpron, M., Nelson, J.L., & Murphy, D.C., 2007. Northern Cordilleran terranes and their interactions through time.
GSA Today, v. 17, no. 4/5, doi:10.1130/GSAT01704-5A.1.
Coney, P.J., Jones, D.L., & Monger, J.W.H., 1980. Cordilleran suspect terranes. Nature, v. 288, p. 329-333.
Conrad, C.P., & Lithgow-Bertelloni, C., 2006. Influence of continental roots and asthenosphere on plate-mantle
coupling. Geophysical Research Letters, v. 33, L05312, doi:10.1029/2005GL025621.
Cottrell, R.D., & Tarduno, J.A., 2000. Late Cretaceous true polar wander: not so fast. Science, v. 288, 2283a.
Coxall, H.K., & Pearson, P.N., 2007. The Eocene-Oligocene transition. In: Williams, M., Haywood, A.M., Gregory,
F.J., & Schmidt, D.N., eds., Deep-time Perspectives on Climate Change: Marrying the signal from computer
models and biological proxies. Geological Society, London, Micropalaeontological Society, p. 351-387.
Croll, J., 1886. Discussions on Climate and Cosmology. London: Edward Stanford.
Cronin, T.M., 2010. Paleoclimates: Understanding climate change past and present. New York: Columbia University
Press.
Dalziel, I.W.D., Mosher, S., & Gahagan, L.M., 2000. Laurentia-Kalahari collision and the assembly of Rodinia. The
Journal of Geology, v. 108, no. 5, p. 499-513.
Darwin, G.H., 1877. On the influence of geological changes on the earth’s axis of rotation. Philosophical Transactions
of the Royal Society of London, A., v. 167, p. 271-312.
Darwin, G.H., 1878. On Professor Haughton’s estimate of geological time. Proceedings of the Royal Society of London,
v. 27, p. 179-183.
Day, A., & Runcorn, S., 1955. Polar wandering: some geological, dynamical and paleomagnetic aspects. Nature, v.
176, p. 422-426.
De Man, E., & Van Simaeys, S., 2004. Late Oligocene warming event in the southern North Sea Basin: benthic
foraminifera as paleotemperature proxies. Netherlands Journal of Geosciences / Geologie en Mijnbouw, v. 83, no.
3, p. 227-239, www.njgonline.nl.
Dickins, J.M., 1987. Tethys – a geosyncline formed on continental crust? In: McKenzie, K.G., ed., Shallow Tethys 2,
International Symposium, Wagga Wagga, 1986. Rotterdam: A.A. Balkema, p. 149-158.
Dickins, J.M., 1994a. What is Pangaea? In: Embry, A.F., Beauchamp, B., & Glass, D.G., Pangea: Global environments
and resources. Canadian Society of Petroleum Geologists, Memoir 17, p. 67-80.
Dickins, J.M., 1994b. The nature of the oceans or Gondwanaland, fact and fiction. In: Gondwana Nine, Ninth
International Gondwana Symposium, Hyderabad, India. Rotterdam: A.A. Balkema, p. 387-396.
Dickins, J.M., 1994c. The southern margin of Tethys. In Gondwana Nine, Ninth International Gondwana Symposium,
Hyderabad, India. Rotterdam: A.A. Balkema, p. 1125-1134.
Dickins, J.M., Choi, D.R., & Yeates, A.N., 1992. Past distribution of oceans and continents. In: Chatterjee, S., &
Hotton, N., eds., New Concepts in Global Tectonics. Lubbock, TX: Texas Tech University Press, p. 193-199.
Dickinson, W.R., 2003. The place and power of myth in geoscience: an associate editor’s perspective. American
Journal of Science, v. 303, p. 856-864.
Dickman, S.R., 2000. Tectonic and cryospheric excitation of the Chandler wobble and a brief review of the secular
motion of the earth’s rotation pole. IAU Colloquium 178: Polar Motion: Historical and Scientific Problems, ASP
Conference Series, v. 208. p. 421-435.
Dietrich, R., Rülke, A., Ihde, J., Lindner, K., Miller, H., Niemeier, W., Schenke, H.-W., & Seeber, G., 2004. Plate
kinematics and deformation status of the Antarctic Peninsula based on GPS. Global and Planetary Change, v. 42,
nos. 1-4, p. 313-321.
79
Dietz, R.S., & Holden, J.C., 1970. Reconstruction of Pangaea: breakup and dispersion of continents, Permian to
present. Journal of Geophysical Research, v. 75, no. 26, p. 4939-4956.
Domeier, M., Van der Voo, R., & Torsvik, T.H., 2012. Paleomagnetism and Pangea: the road to reconciliation.
Tectonophysics, v. 514, p. 14-43.
Donnadieu, Y., Ramstein, G., Fluteau, F., Besse, J., & Meert, J., 2002. Is high obliquity a plausible cause for
Neoproterozoic glaciations? Geophysical Research Letters, v. 29, no. 23, doi:10.1029/2002GL015902.
Donnellan, A., & Luyendyk, B.P., 2004. GPS evidence for a coherent Antarctic plate and for postglacial rebound in
Marie Byrd Land. Global and Planetary Change, v. 42, nos. 1-4, p. 305-311.
Donovan, R.N., & Meyerhoff, A.A., 1982. Paleomagnetic evidence for a large (~2,000 km) sinistral offset along the
Great Glen fault during Carboniferous time: comment. Geology, v. 10, no. 11, p. 604-605.
Douglas, J.G., & Williams, G.E., 1982. Southern polar forests: the early Cretaceous floras of Victoria and their
palaeoclimatic significance. Palaeogeography, Palaeoclimatology, Palaeoecology, v. 39, p. 171-185.
Drewry, G.E., Ramsay, A.T.S., & Smith, A.G., 1974. Climatically controlled sediments, the geomagnetic field, and
trade wind belts in Phanerozoic time. Journal of Geology, v. 82, p. 531-553.
Du Toit, A.L., 1937. Our Wandering Continents. Edinburgh and London: Oliver and Boyd.
Eldrett. J.S., Harding, I.C., Wilson, P.A., Butler, E., & Roberts, A.P., 2007. Continental ice in Greenland during the
Eocene and Oligocene. Nature, v. 446, p. 176-179.
Elliott, C.I., 1994. Lineament tectonics: an approach to basin analysis and exploration. In: Purcell, P.G., & Purcell,
R.R., eds., Sedimentary Basins of Western Australia. Petroleum Exploration Society of Australia Conference
Proceedings, p. 77-90.
Eötvös, R. von, 1913. Verhandlungen der 17 Allgemeinen Konferenz der Internationalen Erdmessung, Berlin, Part 1, p.
111.
Evans, D.A.D., 2003. True polar wander and supercontinents. Tectonophysics, v. 362, p. 303-320.
Evans, D.A.D., 2009. The palaeomagnetically viable, long-lived and all-inclusive Rodinia supercontinent
reconstruction. In: Murphy, J.B., Keppie, J.D., & Hynes, A., eds., Ancient Orogens and Modern Analogues.
Geological Society of London Special Publication, v. 327, p. 371-404.
Evans, J., 1866. On a possible geological cause of changes in the position of the axis of the earth’s crust. Proceedings
of the Royal Society of London, v. 15, p. 46-54.
Frisch, F., & Meschede, M., 2011. Terranes. In: Frisch, F., Meschede, M., & Blakey, R.C., Plate Tectonics:
Continental drift and mountain building. Heidelberg: Springer, p. 131-138.
Gasperid, M., & Chierici, F., 1996. The Eötvös force revisited. Terra Nova, v. 8, no. 4, p. 356-360.
Gold, T., 1955. Instability of the earth’s axis of rotation. Nature, v. 175, p. 526-529.
Goldreich, P., & Toomre, A., 1969. Some remarks on polar wandering. Journal of Geophysical Research, v. 74, no. 10,
p. 2555-2567.
Gordon, R.G., 1987. Polar wandering and paleomagnetism. Annual Reviews of Earth and Planetary Sciences, v. 15, p.
567-593.
Gordon, R.G., 1995. Plate motions, crustal and lithospheric mobility, and paleomagnetism: prospective viewpoint.
Journal of Geophysical Research, v. 100, p. 24,367-24,392.
Gordon, R.G., & Stein, S., 1992. Global tectonics and space geodesy. Science, v. 256, p. 333-342.
Graham, J.W., Buddington, A.F., & Balsley, J.R., 1957. Stress-induced magnetizations of some rocks with analyzed
magnetic minerals. Journal of Geophysical Research, v. 62, no. 3, p. 465-474.
Grant, A.C., 1980. Problems with plate tectonics: the Labrador Sea. Bulletin of Canadian Petroleum Geology, v. 28, p.
252-278.
Grant, A.C., 1992. Intracratonic tectonism: key to the mechanism of diastrophism. In: Chatterjee, S., & Norton, N.,
eds., New Concepts in Global Tectonics. Lubbock: Texas Tech University Press, p. 65-73.
Greff-Lefftz, M., & Besse, J., 2012. Paleo movement of continents since 300 Ma, mantle dynamics and large wander of
the rotational pole. Earth and Planetary Science Letters, v. 345-348, p. 151-158.
Gregory, J.W., 1899. The plan of the earth and its causes. The Geographical Journal, v. 13, p. 225-250.
Gregory, J.W., 1925. Continental drift (review of The Origin of Continents and Oceans by A. Wegener). Nature, v.
115, no. 2886, p. 255-257.
Gregory, J.W., 1929. The geological history of the Atlantic Ocean. Quarterly Journal of Geological Society, v. 85, p.
68-122.
Gregory, J.W., 1930. The geological history of the Pacific Ocean. Quarterly Journal of Geological Society, v. 86, p. 72136.
Gross, R.S., 2000. The excitation of the Chandler wobble. Geophysical Research Letters, v. 27, no. 15, p. 2329-2332.
Gung, Y., Panning, M., & Romanowicz, B., 2003. Global anisotropy and the thickness of continents. Nature, v. 422, p.
707-711.
Gussow, W.C., 1963. Metastasy. In: Munyan, A.C., ed., Polar Wandering and Continental Drift. Society of Economic
Paleontologists and Mineralogists, Spec. Publ. no. 10, p. 146-169.
80
Hall, J.M., & Robinson, P.T., 1979. Deep crustal drilling in the North Atlantic Ocean. Science, v. 204, p. 573-586.
Hall, J.M., & Ryall, P.J.C., 1977. Paleomagnetism of basement rocks, Leg 37. In: Aumento, F., Melson, W.B., et al.,
Initial Reports of the Deep Sea Drilling Project, Leg 37. Washington, DC: US Government Printing Office, p. 425448.
Hallam, A., 1976. How closely did the continents fit together? Nature, v. 262, p. 94-95.
Hancock, G., 1995. Fingerprints of the Gods. London: Heinemann.
Hansen, V.L., 1988. A model for terrane accretion: Yukon-Tanana and Slide Mountain terranes, northwest North
America. Tectonics, v. 7, no. 6, p. 1167-1177.
Hapgood, C.H., 1958. Earth’s Shifting Crust. New York: Pantheon.
Hapgood, C.H., 1970. The Path of the Pole. Philadelphia: Chilton Book Company.
Hardie, L.A., 1990. The role of rifting and hydrothermal CaCl 2 brines in the origin of potash evaporites: an hypothesis.
American Journal of Science, v. 290, p. 1-42.
Hardie, L.A., 1991. On the significance of evaporites. Annual Review of Earth and Planetary Sciences, v. 219, p. 131168.
Hargraves, R.B., & Duncan, R.A., 1973. Does the mantle roll? Nature, v. 245, p. 361-363.
Harrison, C.G.A., 1968. Antipodal locations of continents and ocean basins. Science, v. 153, no. 3741, p. 1246-1248.
Heezen, B.C., Tharp, M., & Ewing, M., 1959. The floors of the oceans. I: The North Atlantic. Geological Society of
America Special Paper 65.
Hellman, H., 1998. Great Feuds in Science: Ten of the liveliest disputes ever. New York: John Wiley & Sons.
Hirth, G., Evans, R.L., & Chave, A.D., 2000. Comparison of continental and oceanic mantle electrical conductivity: is
the Archean lithosphere dry? Geochemistry, Geophysics, Geosystems, v. 1, no. 12, 1030,
doi:10.1029/2000GC000048.
Hodych, J.P., & Bijaksana, S., 1993. Can remanence anisotropy detect paleomagnetic inclination shallowing due to
compaction? A case study using Cretaceous deep-sea limestones. Journal of Geophysical Research, v. 98, no. B12,
p. 22429-22441.
Hoffman, P.F., Kaufman, A.J., Halverson, G.P., & Schrag, D.P., 1998. A Neoproterozoic snowball earth. Science, v.
281, p. 1342-1346.
Holmes, A., 1965. The Principles of Physical Geology. Edinburgh: Nelson.
Hou, G., Santosh, M., Qian, X., Lister, G.S., & Li, J., 2008. Configuration of the late Paleoproterozoic supercontinent
Columbia: insights from radiating mafic dyke swarms. Gondwana Research, v. 14, no. 3, p. 395-409.
Houghton, R.L., Thomas, J.E., Diecchio, R.J., & Tagliacozzo, A., 1979. Radiometric ages of basalts from DSDP Leg
43, Sites 382 and 385 (New England Seamounts), 384 (J-anomaly), 386 and 387 (central and western Bermuda
Rise). In: Tucholke, B.E., Vogt, P.R., et al., Initial Reports of the Deep Sea Drilling Project, Washington, DC: US
Government Printing Office, v. 43, p. 739-753.
Howell, D.G., Schermer, E.R., Jones, D.L., Ben-Avraham, Z., & Scheibner, E., 1983. Tectonostratigraphic terrane map
of the Circum-Pacific region. USGS Open-File Report: 83-716.
Hughes, T., 1973. Coriolis perturbation of mantle convection related to a two-phase convection model. Tectonophysics,
v. 18, nos. 3-4, p. 215-230.
Iro, H., 2002. A Modern Approach to Classical Mechanics. Singapore: World Scientific.
Irving, E., 1977. Drift of the major continental blocks since the Devonian. Nature, v. 270, p. 304-309.
Irving, E., & Archibald, D.A., 1990. Bathozonal tilt corrections to paleomagnetic data from mid-Cretaceous plutonic
rocks: examples from the Omineca belt, British Columbia. Journal of Geophysical Research, v. 95, no. B4, p. 45794585.
Ivany, L.C., Patterson, W.P., & Lohmann, K.C., 2000. Cooler winters as a possible cause of mass extinctions at the
Eocene/Oligocene boundary. Nature, v. 407, p. 887-890.
Jacobitti, E., 1912. Mobilità dell’asse terrestre: studio geologico. Torino: Società Tipografico-Editrice Nazionale.
James, H., 1860. Athenaeum, 25 Aug.
James, K., 2012. Whence the Caribbean? New Concepts in Global Tectonics Newsletter, no. 64, p. 19-23.
James, P., 1997. Is isostasy a real phenomenon? New Concepts in Global Tectonics Newsletter, no. 3, p. 3-4.
Jeffreys, H., 1974. Theoretical aspects of continental drift. In Kahle, C.F., ed., Plate Tectonics – Assessments and
Reassessments. Tulsa, OK: American Association of Petroleum Geologists, Memoir 23, p. 395-405.
Jeffreys, H., 1976. The Earth: Its origin, history and physical constitution. 6th ed., Cambridge: Cambridge University
Press.
Jenkins, G.S., 2000. Global climate model high-obliquity solutions to the ancient climate puzzles of the faint-young-sun
paradox and low-latitude Proterozoic glaciation. Journal of Geophysical Research, v. 105, no. D6, p. 7357-7370.
Jenkins, G.S., 2004. High obliquity as an alternative hypothesis to early and late Proterozoic extreme climate
conditions. In: Jenkins, G.S., McMenamin, M.A.S., McKey, C.P., & Sohl, L., eds., The Extreme Proterozoic:
Geology, geochemistry, and climate. Washington, DC: American Geophysical Union, Geophysical Monograph 146,
p. 183-192.
81
Jiang, W-P., Dong-Chen, E., Zhan, B-W, & Liu, Y-W., 2009. New model of Antarctic plate motion and its analysis.
Chinese Journal of Geophysics, v. 52, no. 1, p. 23-32.
Jones, E.J.W., & Ewing, J.I., 1969. Age of the Bay of Biscay: evidence from seismic profiles and bottom samples.
Science, v. 166, no. 3901, p. 102-105.
Jordan, T.H., 1975. The continental tectosphere. Reviews of Geophysics and Space Physics, v. 13, p. 1-12.
Jordan, T.H., 1978. Composition and development of the continental tectosphere. Nature, v. 274, p. 544-548.
Jordan, T.H., 1988. Structure and formation of the continental tectosphere. Journal of Petrology, Special Lithosphere
Issue, p. 11-37.
Jordan, T.H., Puster, P., Glatzmaier, G.A., & Tackley, P.J., 1993. Comparisons between seismic earth structures and
mantle flow models based on radial correlation functions. Science, v. 261, p. 1427-1431.
Kashfi, M.S., 1992. Geological evidence for a simple horizontal compression of the crust in the Zagros Crush Zone. In:
Chatterjee, S., & Norton, N., eds., New Concepts in Global Tectonics. Lubbock: Texas Tech University Press, p.
119-130.
Keith, M.L., 1993. Geodynamics and mantle flow: an alternative earth model. Earth-Science Reviews, v. 33, p. 153337.
Kennett, J.P., & Shackleton, N.J., 1976. Oxygen isotopic evidence for the development of the psychrosphere 38 Myr
ago. Nature, v. 260, p. 513-515.
Kennett, J.P., & Exon, N.F., 2004. Paleoceanographic evolution of the Tasmanian Seaway and its climatic implications.
In: Exon, N.F., Kennett, J.P., & Malone, M.J., eds., The Cenozoic Southern Ocean: Tectonics, sedimentation and
climate change between Australia and Antarctica. Washington, DC: American Geophysical Union, Geophysical
Monograph Series 151, p. 345-367.
Kent, D.V., & May, S.R., 1987. Polar wander and paleomagnetic reference pole controversies. Reviews of Geophysics,
v. 25, no. 5, p. 961-970.
Kent, D.V., & Van der Voo, R., 1990. Palaeozoic palaeogeography from palaeomagnetism of the Atlantic-bordering
continents. In: McKerrow, W.S., & Scotese, C.R., eds., Palaeozoic Palaeogeography and Biogeography.
Geological Society of London, Memoir 12, p. 49-56.
Kent, D.V., & Smethurst, M.A., 1998. Shallow bias of paleomagnetic inclinations in the Paleozoic and Precambrian.
Earth and Planetary Science Letters, v. 160, p. 391-402.
Kent, P.E., 1969. The geological framework of petroleum exploration in Europe and North Africa and the implications
of continental drift hypotheses. In: The Exploration for Petroleum in Europe and North Africa. London: Institute of
Petroleum, p. 3-17.
Khudoley, K.M., 1974. Circum-Pacific Mesozoic ammonoid distribution: relation to hypotheses of continental drift,
polar wandering, and earth expansion. In: Kahle, C.F., ed., Plate Tectonics – Assessments and Reassessments.
American Association of Petroleum Geologists, Memoir 23, p. 295-330.
Khudoley, K.M., 1988. The paleobiogeography of the Atlantic Ocean in the Jurassic period. International Geology
Review, v. 30, p. 623-634.
King, R.E., 1971. Geology and history of Sicily. American Association of Petroleum Geologists Bulletin, v. 55, no. 6, p.
901-902.
Kirschvink, J.L., Ripperdan, R.L., & Evans, D.A., 1997. Evidence for a large-scale reorganization of early Cambrian
continental masses by inertial interchange true polar wander. Science, v. 277, p. 541-545.
Kobashi, T., Grossman, E.L., Yancey, T.E., & Dockery, D.T., 2001. Reevaluation of conflicting Eocene tropical
temperature estimates: molluskan oxygen isotope evidence for warm low latitudes. Geology, v. 29, p. 983-986.
Köppen, W., & Wegener, A., 1924. Die Klimate der geologischen Vorzeit. Berlin: Gebrüder Bornträger.
Kreichgauer, P.D., 1902. Die Äquatorfrage in der Geologie. Steyl: Druck und Verlag der Missionsdruckerei; 2nd ed.
1926.
Kustowski, B., Ekström, G., & Dziewonski, A.M., 2008. Anisotropic shear-wave velocity structure of the Earth’s
mantle: a global model. Journal of Geophysical Research, v. 113, B06306, doi:10.1029/2007JB005169.
Laing, A.C.M., 1998. The myth of plate tectonics. New Concepts in Global Tectonics Newsletter, no. 6, p. 14-24.
Lake, P., 1922. Wegener’s displacement theory. Geology Magazine, v. 59, p. 338-346.
Larson, R.I., & Chase, C.G., 1972. Late Mesozoic evolution of the western Pacific Ocean. Geological Society of
America Bulletin, v. 83, no. 2, p. 3627-3643.
Laubscher, H.P., 1975. Plate boundaries and microplates in Alpine history. American Journal of Science, v. 275, no. 8,
p. 865-876.
Lear, C.H., Bailey, T.R., Pearson, P.N., Coxall, H.K., & Rosenthal, Y., 2008. Cooling and ice growth across the
Eocene-Oligocene transition. Geology, v. 36, no. 3, p. 251-254.
Legendre, C.P., Meier, T., Lebedev, S., Friederich, W., & Viereck-Götte, L., 2012. A shear wave velocity model of the
European upper mantle from automated inversion of seismic shear and surface waveforms. Geophysical Journal
International, v. 191, no. 1, p. 282-304.
Le Grand, H.E., 1998. Drifting Continents and Shifting Theories. Cambridge: Cambridge University Press.
82
Li, Z.X., Powell, C.McA., Thrupp, G.A., & Schmidt, P.W., 1990. Australian Palaeozoic palaeomagnetism and tectonics
- II. A revised apparent polar wander path and palaeogeography. Journal of Structural Geology, v. 12, p. 567-575.
Li, Z.X., Bogdanova, S.V., Collins, A.S., Davidson, A., et al., 2008. Assembly, configuration, and break-up history of
Rodinia: a synthesis. Precambrian Research, v. 160, p. 179-210.
Liu, Z., Pagani, M., Zinniker, D., DeConto, R., Huber, M., Brinkhuis, H., Shah, S.R., Leckie, R.M., & Pearson, A.,
2009. Global cooling during the Eocene-Oligocene climate transition. Science, v. 323, no. 5918, p. 1187-1190.
Longwell, C.R., 1944. Some thoughts on the evidence of a continental drift. American Journal of Science, v. 242, p.
218-231.
Lowman, P.D., 1986. Plate tectonics with fixed continents: a testable hypothesis – II. Journal of Petroleum Geology, v.
9, p. 71-87.
Lubbock, J., 1848. On change of climate resulting from a change in the earth’s axis of rotation. Quarterly Journal of
the Geological Society, v. 4, p. 4-7.
Luyendyk, B.P., & Melson, W.G., 1967. Magnetic properties and petrology of rocks near the crest of the Mid-Atlantic
Ridge. Nature, v. 215, no. 5097, p. 147-149.
MacDonald, G.J.F., 1963. The deep structure of continents. Reviews of Geophysics, v. 1, p. 587-665.
MacDonald, G.J.F., 1965. Continental structure and drift. In: A Symposium on Continental Drift. Royal Society of
London Philosophical Transactions, Series A, v. 258, p. 215-227.
Macdougall, D., 1971. Deep sea drilling: age and composition of an Atlantic basaltic intrusion. Science, v. 171, p.
1244-1255.
Mantura, A.J., 1972. New global tectonics and ‘the new geometry’. American Association of Petroleum Geologists
Bulletin, v. 56, p. 2451-2455.
Masters, G., Johnson, S., Laske, G., & Bolton, H., 1996. A shear-velocity model of the mantle. Philosophical
Transactions of the Royal Society of London A, v. 354, p. 1385-1411.
Maxwell, J.C., 1970. The Mediterranean, ophiolites, and continental drift. In: Johnson, H., & Smith, B.L., eds., The
Megatectonics of Continents and Oceans. Rutgers, NJ: Rutgers University Press, p. 167-193.
McDowell, F.W., Lehman, D.H., Gucwa, P.R., Fritz, D., & Maxwell, J.C., 1984. Glaucophane schists and ophiolites of
the northern California Coast Ranges: isotopic ages and their tectonic implications. Geological Society of America
Bulletin, v. 95, no. 11, p. 1373-1382.
McElhinny, M., & McFadden, P., 2000. Paleomagnetism: Continents and oceans. San Diego, CA: Academic Press.
McKenzie, D.P., 1966. The viscosity of the lower mantle. Geophysical Journal of the Royal Astronomical Society, v.
14, p. 297-305.
McKenzie, D., & Priestley, K. 2008. The influence of lithospheric thickness variations on continental evolution. Lithos,
v. 102, nos. 1-2, p. 1-11, http://bullard.esc.cam.ac.uk.
McKenzie, K.G., 1987. Tethys and her progeny. In: McKenzie, K.G., ed., Shallow Tethys 2, International Symposium,
Wagga Wagga, 1986. Rotterdam: A.A. Balkema, p. 501-523.
Meert, J.G., 2012. What’s in a name? The Columbia (Paleopangaea/Nuna) supercontinent. Gondwana Research, v. 21,
no. 4, p. 987-993.
Melson, W.G., Hart, S.R., & Thompson, G., 1972. St. Paul’s Rocks, equatorial Atlantic: petrogenesis, radiometric ages,
and implications on sea-floor spreading. In: Shagam, R., Hargraves, R.B., Morgan, W.J., Van Houten, F.B., Burk,
C.A., Holland, H.D., & Hollister, L.C., eds., Studies in Earth and Space Sciences. Boulder, CO: Geological Society
of America, Memoir 132, p. 241-272.
Merrill, R.T., McElhinny, M.W., & McFadden, P.L., 1996. The Magnetic Field of the Earth. San Diego, CA: Academic
Press.
Meyerhoff, A.A., 1970a. Continental drift: implications of paleomagnetic studies, meteorology, physical oceanography,
and climatology. Journal of Geology, v. 78, p. 1-51.
Meyerhoff, A.A., 1970b. Continental drift. II: High latitude evaporite deposits and geologic history of Arctic and North
Atlantic oceans. Journal of Geology, v. 78, p. 406-444.
Meyerhoff, A.A., 1973. Origin of Arctic and North Atlantic oceans. In: Pitcher, M.G., ed., Arctic Geology. Tulsa, OK:
American Association of Petroleum Geologists, Memoir 19, p. 562-582.
Meyerhoff, A.A., 1974. Crustal structure of northern North Atlantic Ocean – a review. In: Kahle, C.F., ed., Plate
Tectonics – Assessments and Reassessments. Tulsa, OK: American Association of Petroleum Geologists, Memoir
23, p. 423-433.
Meyerhoff, A.A., 1995. Surge-tectonic evolution of southeastern Asia: a geohydrodynamics approach. Journal of
Southeast Asian Earth Sciences, v. 12, p. 143-247.
Meyerhoff, A.A., & Harding, J.L., 1971. Some problems in current concepts of continental drift. Tectonophysics, v. 12,
p. 235-260.
Meyerhoff, A.A., & Hatten, C.W., 1974. Bahamas salient of North America. In: Burk, C.A., & Drake, C.L., eds., The
Geology of Continental Margins. Berlin: Springer-Verlag, p. 429-446.
Meyerhoff, A.A., & Meyerhoff, H.A., 1974a. Tests of plate tectonics. In: Kahle, C.F., ed., Plate Tectonics –
83
Assessments and Reassessments. Tulsa, OK: American Association of Petroleum Geologists, Memoir 23, p. 43-145.
Meyerhoff, A.A., & Meyerhoff, H.A., 1974b. Ocean magnetic anomalies and their relations to continents. In: Kahle,
C.F., ed., Plate Tectonics – Assessments and Reassessments. Tulsa, OK: American Association of Petroleum
Geologists, Memoir 23, p. 411-422.
Meyerhoff, A.A., & Teichert, C., 1971. Continental drift. III: Late Paleozoic glacial centers and Devonian-Eocene coal
distribution. Journal of Geology, v. 79, p. 285-321.
Meyerhoff, A.A., Kamen-Kaye, M., Chen, C., & Taner, I., 1991. China – Stratigraphy, Paleogeography and Tectonics.
Dordrecht: Kluwer.
Meyerhoff, A.A., Agocs, W.B., Taner, I., Morris, A.E.L., & Martin, B.D., 1992a. Origin of midocean ridges. In:
Chatterjee, S., & Norton, N., eds., New Concepts in Global Tectonics. Lubbock: Texas Tech University Press, p.
151-178.
Meyerhoff, A.A., Taner, I., Morris, A.E.L., Martin, B.D., Agocs, W.B., & Meyerhoff, H.A., 1992b. Surge tectonics: a
new hypothesis of earth dynamics. In: Chatterjee, S., & Norton, N., eds., New Concepts in Global Tectonics.
Lubbock: Texas Tech University Press, p. 309-409.
Meyerhoff, A.A., Taner, I., Morris, A.E.L., Agocs, W.B., Kaymen-Kaye, M., Bhat, M.I., Smoot, N.C., & Choi, D.R.,
1996a. Surge Tectonics: A new hypothesis of global geodynamics (D. Meyerhoff Hull, ed.). Dordrecht: Kluwer.
Meyerhoff, A.A., Boucot, A.J., Meyerhoff Hull, D., & Dickins J.M., 1996b. Phanerozoic Faunal & Floral Realms of
the Earth: The intercalary relations of the Malvinokaffric and Gondwana faunal realms with the Tethyan faunal
realm. Boulder, CO: Geological Society of America, Memoir 189.
Meyerhoff, H.A., & Meyerhoff, A.A., 1978. Spreading history of the eastern Indian Ocean and India’s northward flight
from Antarctica and Australia: discussion and reply. Geological Society of America Bulletin, v. 89, p. 637-640.
Morel, P., & Irving, E., 1978. Tentative paleocontinental maps for the early Phanerozoic and Proterozoic. Journal of
Geology, v. 86, p. 535-561.
Morel, P., & Irving, E., 1981. Paleomagnetism and the evolution of Pangea. Journal of Geophysical Research, v. 86,
no. B3, p. 1858-1872.
Müller, R.D., Sdrolias, M., Gaina, C., & Roest, W.R., 2008. Age, spreading rates and spreading symmetry of the
world’s ocean crust. Geochemistry, Geophysics, Geosystems, v. 9, Q04006, doi:10.1029/2007GC001743.
Munk, W.H., & MacDonald, G.J.F., 1975. The Rotation of the Earth: A geophysical discussion. 2nd ed., Cambridge:
Cambridge University Press.
Newton, C.R., 1988. Significance of ‘Tethyan’ fossils in the American Cordillera. Science, v. 242, no. 4877, p. 385391.
Northrop, J.W., & Meyerhoff, A.A., 1963. Validity of polar and continental movement hypotheses based on
paleomagnetic studies. American Association of Petroleum Geologists Bulletin, v. 47, no. 4, p. 575-585.
Oard, M.J., 2000a. Literature criticisms of plate tectonics. In: Reed, J.K., ed., Plate Tectonics: A different view. St.
Joseph, MO: Creation Research Society Books, p. 25-64.
Oard, M.J., 2000b. Subduction unlikely – plate tectonics improbable. In Reed, J.K., ed., Plate Tectonics: A different
view. St. Joseph, MO: Creation Research Society Books, p. 93-145.
O’Driscoll, E.S.T., 1986. Observations of the lineament-ore relation. Philosophical Transactions of the Royal Society of
London A, v. 317, p. 195-218.
Opdyke, N.D., & Hekinian, R., 1967. Magnetic properties of some igneous rocks from the Mid-Atlantic Ridge. Journal
of Geophysical Research, v. 72, no. 8, p. 2257-2260.
O’Regan, M., Williams, C.J., Frey, K.E., & Jakobsson, M., 2011. A synthesis of the long-term paleoclimatic evolution
of the Arctic. Oceanography, v. 24, no. 3, p. 66-80, http://www.tos.org.
O’Reilly, S.Y., Zhang, M., Griffin, W.L., Begg, G., & Hronsky, J., 2009. Ultradeep continental roots and their oceanic
remnants: a solution to the geochemical ‘mantle reservoir’ problem? Lithos, 211S, p. 1043-1054.
Orlenok, V.V., 1986. The evolution of ocean basins during Cenozoic time. Journal of Petroleum Geology, v. 9, p. 207216.
Ozima, M., Ozima, M., & Kaneoka, I., 1968. Potassium-argon ages and some magnetic properties of some dredge
submarine basalts and their geophysical implications. Journal of Geophysical Research, v. 73, no. 2, p. 711-723.
Parnell, J., 1982. Paleomagnetic evidence for a large (~2,000 km) sinistral offset along the Great Glen fault during
Carboniferous time: comment. Geology, v. 10, no. 11, p. 605.
Pavlenkova, G.A., & Pavlenkova, N.I., 2006. Upper mantle structure of the Northern Eurasia from peaceful nuclear
explosion data. Tectonophysics, v. 416. p. 3-52.
Pavlenkova, N.I., 1990. Crustal and upper mantle structure and plate tectonics. In: Barto-Kyriakidis, A., ed., Critical
Aspects of the Plate Tectonics Theory. Athens: Theophrastus Publications, v. 1, p. 73-86.
Pavlenkova, N.I., 1995. Structural regularities in the lithosphere of continents and plate tectonics. Tectonophysics, v.
243, p. 223-229.
Pavlenkova, N.I., 1996. General features of the uppermost mantle stratification from long-range seismic profiles.
Tectonophysics, v. 264, p. 261-278.
84
Pavlenkova, N.I., 2012. The earth’s degassing, rotation and expansion as sources of global tectonics. New Concepts in
Global Tectonics Newsletter, no. 63, p. 49-71.
Pedlosky, J., 1979. Geophysical Fluid Dynamics. New York: Springer-Verlag.
Peltier, W.R., & Wu, P., 1983. Continental lithospheric thickness and deglaciation-induced true polar wander.
Geophysical Research Letters, v. 10, p. 181-184.
Persson, A., 1998. How do we understand the Coriolis force? Bulletin of the American Meteorological Society, v. 79,
no. 7, p. 1373-1385, www.aos.princeton.edu.
Pilot, J., Werner, C.-D., Haubrich, F., & Baumann, D., 1998. Paleozoic and Proterozoic zircons from the Mid-Atlantic
Ridge. Nature, v. 393, p. 676-679.
Piper, J.D.A., 2006. A ~90° late Silurian-Early Devonian apparent polar wander loop: the latest inertial interchange of
planet earth? Earth and Planetary Science Letters, v. 250, nos. 1-2, p. 345-357.
Pollack, H.N., & Chapman, D.S., 1977. On the regional variation of heat flow, geotherms, and lithospheric
thickness. Tectonophysics, v. 38, p. 279-296.
Poupinet, G., & Shapiro, N.M., 2009. Worldwide distribution of ages of the continental lithosphere derived from a
global seismic tomographic model. Lithos, v. 109, nos. 1-2, p. 125-130, 10.1016/j.lithos.2008.10.023.
Pratsch, J.-C., 1990. Relative motions in geology: some philosophical differences. Journal of Petroleum Geology, v. 13,
p. 229-234.
Pratt, D., 2000. Plate tectonics: a paradigm under threat. Journal of Scientific Exploration, v. 14, no. 3, p. 307-352,
www.scientificexploration.org.
Pratt, D., 2001. Problems with plate tectonics. New Concepts in Global Tectonics Newsletter, no. 21, p. 10-24.
Prévot, M., Mattern, E., Camps, P., & Daignières, M., 2000. Evidence for a 20° tilting of the earth’s rotation axis 110
million years ago. Earth and Planetary Science Letters, v. 179, nos. 3-4, p. 517-528.
Price, J.F., 2006. A Coriolis tutorial. Woods Hole Oceanographic Institute, version 3.3, www.whoi.edu.
Priestley, K., & McKenzie, D., 2006. The thermal structure of the lithosphere from shear wave velocities. Earth and
Planetary Science Letters, v. 244, p. 285-301, www.seg2.ethz.ch.
Raiverman, V., 1992. Trans-Asiatic lineaments and Himalayan orogeny. In: Sinha, A.K., ed., Himalayan Orogen and
Global Tectonics. International Lithosphere Programme, publication no. 197, p. 121-155.
Raub, T.D., Kirschvink, J.L., & Evans, D.A.D., 2007. True polar wander: linking deep and shallow geodynamics to
hydro- and bio-spheric hypotheses. In: Schubert, G., ed., Treatise on Geophysics. Amsterdam: Elsevier, v. 5, ch. 14,
p. 565-589.
Reibisch, P., 1901. Ein Gestaltungsprinzip der Erde. 27. Jahresbericht des Vereins für Erdkunde zu Dresden, p. 105124.
Reilinger, R., McClusky, S., Vernant, P., Lawrence, S., et al., 2006. GPS constraints on continental deformation in the
Africa-Arabia-Eurasia continental collision zone and implications for the dynamics of plate interactions. Journal of
Geophysical Research, v. 111, no. B5, doi:10.1029/2005/JB004051.
Reynolds, P.H., & Clay, W., 1977. Leg 37 basalts and gabbro: K-Ar and 40Ar-39Ar dating. In: Aumento, F. Melson,
W.G., et al., Initial Reports of the Deep Sea Drilling Project. Washington, DC: US Government Printing Office, v.
37, p. 629-30.
Rezanov, I.A., 1968. Paleomagnetism and continental drift. International Geology Review, v. 10, p. 765-776.
Ricard, Y., 2007. Physics of mantle convection. In: Schubert, G., ed., Treatise on Geophysics. Amsterdam: Elsevier, v.
7: Mantle dynamics.
Rogers, J.J.W., & Santosh, M., 2002. Configuration of Columbia, a Mesoproterozoic supercontinent. Gondwana
Research, v. 5, p. 5-22.
Ruditch, E.M., 1990. The world ocean without spreading. In: Barto-Kyriakidis, A., ed., Critical Aspects of the Plate
Tectonics Theory. Athens: Theophrastus Publications, v. 2, p. 343-395.
Rudnick, R.L., McDonough, W.F., & O’Connell, R.J., 1998. Thermal structure, thickness and composition of
continental lithosphere. Chemical Geology, v. 145, p. 395-411.
Russell, C.T., & Dougherty, M.K., 2010. Magnetic fields of the outer planets. Space Science Reviews, v. 152, p. 251269.
Sager, W.W., Weiss, C.J., Tivey, M.A., & Johnson, H.P., 1998. Geomagnetic polarity reversal model of deep-tow
profiles from the Pacific Jurassic Quiet Zone. Journal of Geophysical Research, v. 103, no. B3, p. 5269-5286.
Salamy, K.A., & Zachos, J.C., 1999. Latest Eocene-early Oligocene climate change and Southern Ocean fertility:
inferences from sediment accumulation and stable isotope data. Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 145, p. 61-77.
Sandwell, D.T., & Renkin, M.L., 1988. Compensation of swells and plateaus in the north Pacific: no direct evidence for
mantle convection. Journal of Geophysical Research, v. 93, no. B-4, p. 2775-2783.
Sato, H., 1991. Viscosity of the upper mantle from laboratory creep and anelasticity measurements in peridotite at high
pressure and temperature. Geophysical Journal International, v. 105, no. 3, p. 587-599.
Saul, L.R., 1986. Pacific west coast Cretaceous molluscan faunas: time and aspect of changes. In: Abbott, P.L., ed.,
85
Cretaceous stratigraphy, western North America. Los Angeles, CA: Society of Economic Paleontologists and
Mineralogists, Pacific Section, Book 46, p. 131-135.
Saxena, M.N., & Gupta, V.J., 1990. Role of foredeep tectonics, geological and palaeontological data, gravity tectonics
in the orogeny and uplift of the Himalayas, vis-a-vis continental drift and plate tectonics concepts. In: BartoKyriakidis, A., ed., Critical Aspects of the Plate Tectonics Theory. Athens: Theophrastus Publications, v. 1, p. 105128.
Saxena, M.N., Gupta, V.J., Meyerhoff, A.A., & Archbold, N.W., 1985. Tectonic and spatial relations between India
and Asia since Proterozoic time. In: Gupta, V.J., ed., Contributions to Himalayan Geology, v. 3: Geology of
Western Himalayas. Delhi: Hindustan Publishing Corporation, p. 187-207.
Scheidegger, A.E., 1963. Principles of Geodynamics. 2nd ed., New York: Academic Press.
Schiaparelli, G.V., 1889. De la rotation de la terre sous l’influence des actions géologiques. St. Petersburg: Imprimerie
de l’Académie Impériale des Sciences.
Schmidt, P.W., Powell, C.McA., Li, Z.X., & Thrupp, G.A., 1990. Reliability of Palaeozoic palaeomagnetic poles and
APWP of Gondwanaland. Tectonophysics, v. 184, p. 87-100.
Scotese, C.R., Boucot, A.J., & McKerrow, W.S., 1999. Gondwanan palaeogeography and palaeoclimatology. Journal
of African Earth Sciences, v. 28, no. 1, p. 99-114.
Schouten, S., Eldrett, J., Greenwood, D.R., Harding, I., Baas, M., & Sinninghe Damsté, J.S., 2008. Onset of long-term
cooling of Greenland near the Eocene-Oligocene boundary as revealed by branched tetraether lipids. Geology, v. 36,
no. 2, p. 147-150.
Schuchert, C., 1928. The hypothesis of continental displacement. In: Waterschoot van der Gracht, W.A.J.M. van, ed.,
The Theory of Continental Drift: A symposium. Tulsa, OK: American Association of Petroleum Geologists, p. 104144.
Schuchert, C., 1932. Gondwana land bridges. Geological Society of America Bulletin, v. 43, p. 875-915.
Scotese, C.R., & Barrett, S.F., 1990. Gondwana’s movement over the south pole during the Palaeozoic: evidence from
lithological indicators of climate. In: McKerrow, W.S., & Scotese, C.R., eds., Palaeozoic Palaeogeography and
Biogeography. Geological Society of London, Memoir 12, p. 75-85.
Seiders, V.M., 1988. Origin of conglomerate stratigraphy in the Franciscan assemblage and Great Valley sequence,
northern California. Geology, v. 16, no. 9, p. 783-787.
Sengör, A.M.C., 1990. Lithotectonic terranes and the plate tectonic theory of orogeny: a critique of the principles of
terrane analysis. In: Wiley, T.J., Howell, D.G., & Wong, F.L., eds., Terrane Analysis of China and the Pacific Rim.
Houston, TX: Circum-Pacific Council for Energy and Mineral Resources Earth Science Series, v. 13, p. 9-44.
Shapiro, M.N., 1990. Is the opening of the Atlantic manifested in the structure of the northern framing of the Pacific
Ocean? Barto-Kyriakidis, A., ed., Critical Aspects of the Plate Tectonics Theory. Athens: Theophrastus
Publications, v. 1, p. 179-196.
Shapiro, N.M., Ritzwoller, M.H., Mareschal, J.C., & Jaupart, C., 2004. Lithospheric structure of the Canadian Shield
inferred from inversion of surface-wave dispersion with thermodynamic a priori constraints. In: Curtis, A., &
Wood, R., eds., Geological Prior Information: Informing science and engineering. Geological Society of London,
Special Publication, v. 239, p. 175-194.
Simpson, G.C., Seward, A.C., Gregory, J.W., Mitchell, P., Brooks, C.E.P., Tate Regan, C., & Thomas, H., 1930.
Discussion on geological climates. Proceedings of the Royal Society of London, Series B, v. 106, no. 744, p. 299317.
Simpson, G.G., 1943. Mammals and the nature of continents. American Journal of Science, v. 241, p. 1-31.
Simroth, H., 1907. Die Pendulationstheorie. Leipzig: Grethlein.
Smiley, C.J., 1967. Paleoclimatic interpretations of some Mesozoic floral sequences. American Association of
Petroleum Geologists Bulletin, v. 51, p. 849-863.
Smiley, C.J., 1974. Analysis of crustal relative stability from some late Paleozoic and Mesozoic floral records. In:
Kahle, C.F., ed., Plate Tectonics – Assessments and Reassessments. American Association of Petroleum Geologists,
Memoir 23, p. 331-360.
Smiley, C.J., 1976. Pre-Tertiary phytogeography and continental drift – some apparent discrepancies. In: Gray, J., &
Boucot, A.J., eds., Historical Biogeography, Plate Tectonics, and the Changing Environment. Corvallis: Oregon
State University Press, p. 311-319.
Smiley, C.J., 1992. Paleofloras, faunas, and continental drift: some problem areas. In: Chatterjee, S., & Horton, N.,
eds., New Concepts in Global Tectonics. Lubbock: Texas Tech University Press, p. 241-257.
Smith, A.G., & Hallam, A., 1970. The fit of the southern continents. Nature, v. 225, p. 139-144.
Smith, A.G., Hurley, A.M., & Briden, J.C., 1981. Phanerozoic Paleocontinental World Maps. Cambridge: Cambridge
University Press.
Smoot, N.C., 1997. Aligned buoyant highs, across-trench deformation, clustered volcanoes, and deep earthquakes are
not aligned with plate-tectonic theory. Geomorphology, v. 18, p. 199-222.
Steers, J.A., 1950. The Unstable Earth. 5th ed., London: Methuen.
86
Stehli, F.G., 1957. Possible Permian climatic zonation and its implications. American Journal of Science, v. 255, p.
607-618.
Stehli, F.G., & Grant, R.E., 1971. Permian brachiopods from Axel Heiberg Island, Canada, and an index of sampling
efficiency. Journal of Paleontology, v. 45, no. 3, p. 502-521.
Steinberger, B., & Torsvik, T.H., 2008. Absolute plate motions and true polar wander in the absence of hotspot tracks.
Nature, v. 452, p. 620-623.
St. John, K.E.K., 2008. Cenozoic ice-rafting history of the central Arctic Ocean: terrigenous sands on the Lomonosov
Ridge. Paleoceanography, v. 23, PA1505, doi:10.1029/2007PA001483.
Stöcklin, J., 1989. Tethys evolution in the Afghanistan-Pamir-Pakistan region. In: Sengör, A.M.C., ed., Tectonic
Evolution of the Tethyan Region. Dordrecht: Kluwer, p. 241-264.
Storetvedt, K.M., 1990. The Tethys Sea and the Alpine-Himalayan orogenic belt; mega-elements in a new global
tectonic system. Physics of the Earth and Planetary Interiors, v. 62, p. 141-184.
Storetvedt, K.M., 1992. Rotating plates: new concept of global tectonics. In: Chatterjee, S., & Horton III, N., eds., New
Concepts in Global Tectonics. Lubbock: Texas Tech University Press, p. 203-220.
Storetvedt, K.M., 1997. Our Evolving Planet: Earth history in new perspective. Bergen, Norway: Alma Mater.
Storetvedt, K.M., 2003. Global Wrench Tectonics: Theory of earth evolution. Bergen, Norway: Fagbokforlaget.
Storetvedt, K.M., 2007. Review of Tectonic Consequences of the Earth’s Rotation by R.C. Bostrom. Journal of
Scientific Exploration, v. 21, no. 2, p. 405-412.
Storetvedt, K.M., 2010. World magnetic anomaly map and global tectonics. New Concepts in Global Tectonics
Newsletter, no. 57, p. 27-53.
Storetvedt, K.M. & Longhinos, B., 2011. Evolution of the North Atlantic: paradigm shift in the offing. New Concepts in
Global Tectonics Newsletter, no. 59, p. 9-48.
Storetvedt, K.M., & Longhinos, B., 2012. The Atlantic and its bordering continents – a wrench tectonic analysis:
lithospheric deformation, basin histories and major hydrocarbon provinces. New Concepts in Global Tectonics
Newsletter, no. 64, p. 30-68.
Storetvedt, K.M., & Bouzari, S., 2012. The Tethys configuration and principal tectonic features of the Middle East: a
wrench tectonic survey. New Concepts in Global Tectonics Newsletter, no. 65, p. 103-142.
Tarling, D.H., 1982a. A Proterozoic Pangaea? Nature, v. 299, p. 207.
Tarling, D.H., 1982b. Land bridges and plate tectonics. Geobios, Mémoire Spécial 6, p. 361-374.
Tauxe, L., 2013. Essentials of Paleomagnetism. University of California Press, 2010; 2nd web edition, 2013,
http://magician.ucsd.edu/Essentials_2.
Teichert, C., 1974. Marine sedimentary environments and their faunas in Gondwana area. In: Kahle, C.F., ed., Plate
Tectonics – Assessments and Reassessments. American Association of Petroleum Geologists, Memoir 23, p. 361394.
Teichert, C., & Meyerhoff, A.A., 1972. Continental drift and the marine environment. 24th International Geology
Conference, Montreal, p. 339-349.
Thomson, W., 1876. Presidential Address, British Association for the Advancement of Science. Reprinted in
Mathematical and Physical Papers, 3, Cambridge University Press, p. 320-335.
Torsvik, T.H., Meert, J.G., Smethurst, M.A., Evans, D.A., Ripperdan, R.L., & Kirschvink, J.L., 1998. Polar wander and
the Cambrian (comment by Torsvik, Meert & Smethurst, and reply by Evans, Ripperdan & Kirschvink). Science, v.
279, no. 5347, p. 9.
Torsvik, T.H., Van der Voo, R., Preeden, U., Niocaill, C.M., et al., 2012. Phanerozoic polar wander, palaeogeography
and dynamics. Earth-Science Reviews, v. 114, nos. 3-4, p. 325-368.
Tripati, A., Backman, J., Elderfield, H., & Ferretti, P., 2005. Eocene bipolar glaciation associated with global carbon
cycle changes. Nature, v. 436, p. 341-346.
Tripati, A.K., Eagle, R.A., Morton, A., Dowdeswell, J.A., et al., 2008. Evidence for glaciation in the northern
hemisphere back to 44 Ma from ice-rafted debris in the Greenland Sea. Earth and Planetary Science Letters, v. 265,
nos. 1-2, p. 112-122.
Trümpy, R., 1971. Stratigraphy in mountain belts. Quarterly Journal of the Geological Society of London, v. 126, p.
293-318.
Tsai, V.C., & Stevenson, D.J., 2007. Theoretical constraints on true polar wander. Journal of Geophysical Research, v.
112, B05415, doi:10.1029/2005JB003923, www.gps.caltech.edu.
Umbgrove, J.H.F., 1947. The Pulse of the Earth. 2nd ed., The Hague: Martinus Nijhoff.
Van der Voo, R., 1990. The reliability of paleomagnetic data. Tectonophysics, v. 184, p. 1-9.
Van der Voo, R., 1993. Paleomagnetism of the Atlantic, Tethys and Iapetus Oceans. Cambridge: Cambridge University
Press.
Van der Voo, R., 1994. True polar wander during the middle Paleozoic? Earth and Planetary Science Letters, v. 122,
nos. 1-2, p. 239-243.
Van der Voo, R. 1998. A complex field. Science, v. 281, p. 791-792.
87
Van der Voo, R., & French, R.B., 1974. Apparent polar wandering for the Atlantic-bordering continents: late
Carboniferous to Eocene. Earth-Science Reviews, v. 10, no. 2, p. 99-119.
Van Summeren, J., Conrad, C.P., & Lithgow-Bertelloni, C., 2012. The importance of slab pull and a global
asthenosphere to plate motions. Geochemistry, Geophysics, Geosystems, v. 13, Q0AK03,
doi:10.1029/2011GC003873.
Vasiliev, B.I., & Yano, T., 2007. Ancient and continental rocks discovered in the ocean floors. New Concepts in Global
Tectonics Newsletter, no. 43, p. 3-17.
Vasiliev, B.I., & Choi, D.R., 2008. Geology and tectonic development of the Pacific Ocean. Part 3: Structure and
composition of the basement. New Concepts in Global Tectonics Newsletter, no. 48, p. 23-51.
Vermeersen, L.L.A., Fournier, A., & Sabadini, R., 1997. Changes in rotation induced by Pleistocene ice masses with
stratified analytical earth models. Journal of Geophysical Research, v. 102, no. B12, p. 27,689-27,702,
doi:10.1029/97JB01738.
Wanless, R.K., Stevens, R.D., Lachance, G.R., & Edmonds, C.M., 1968. Age determinations and geological studies. KAr isotopic ages, report 8. Geological Survey of Canada, Paper 67-2, pt. A, p. 140-141.
Wegener, A., 1912. Die Entstehung der Kontinente. Geologische Rundschau, v. 3, p. 276-292.
Wegener, A., 1929. The Origin of Continents and Oceans. English translation of 4th German edition, New York:
Dover, 1966.
Wells, J.M., & Verhoogen, J., 1967. Late Paleozoic paleomagnetic poles and the opening of the Atlantic Ocean.
Journal of Geophysical Research, v. 72, no. 6, p. 1777-1782.
Wesson, P.S., 1974a. Problems of plate tectonics and continental drift. In: Kahle, C.F., ed., Plate Tectonics –
Assessments and Reassessments. American Association of Petroleum Geologists, Memoir 23, p. 146-154.
Wesson, P.S., 1974b. Lomnitz law versus elasticoviscous controversy. In: Kahle, C.F., ed., Plate Tectonics –
Assessments and Reassessments. American Association of Petroleum Geologists, Memoir 23, p. 448-462.
Williams, G.E., 1993. History of the earth’s obliquity. Earth-Science Reviews, v. 34, p. 1-45.
Williams, H., & Hatcher, R.D., 1982. Suspect terranes and accretionary history of the Appalachian orogen. Geology, v.
10, p. 530-536.
Willis, B., 1932. Isthmian links. Geological Society of America Bulletin, v. 43, p. 917-952.
Wolfart, R., 1967. Zur Entwicklung der paläozoischen Tethys in Vorderasien. Erdöl und Kohle-Erdgas-Petrochemie, v.
20, no. 3, p. 168-180.
Wolfe, J.A., 1978. A paleobotanical interpretation of Tertiary climates in the northern hemisphere. American Scientist,
v. 66, p. 694-703.
Wolfe, J.A., 1980. Tertiary climates and floristic relationships at high latitudes in the northern hemisphere.
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 30, p. 313-323.
Xu Qinqi, 1979. On the causes of ice ages. Scientia Geologica Sinica, v. 7, p. 252-263.
Xu Qinqi, 1980. Climatic variation and the obliquity. Vertebrata PalAsiatica, v. 18, p. 334-343.
Yano, T., Choi, D.R., Gavrilov, A.A., Miyagi, S., & Vasiliev, B.I., 2009. Ancient and continental rocks in the Atlantic
Ocean, New Concepts in Global Tectonics Newsletter, no. 53, p. 4-37.
Yano, T., Vasiliev, B.I., Choi, D.R., Miyagi, S., Gavrilov, A.A., & Adachi, H., 2011. Continental rocks in the Indian
Ocean. New Concepts in Global Tectonics Newsletter, no. 58, p. 9-28.
Yuan, H., & Romanowicz, B., 2010. Lithospheric layering in the North American craton. Nature, v. 466, p. 1063-1068,
doi:10.1038/nature09332.
Zachos, J.C., Quinn, T.M., & Salamy, K.A., 1996. High-resolution (104) deep-sea foraminiferal stable isotope records
of the Eocene-Oligocene climate transition. Paleoceanography, v. 11, no. 3, p. 251-266, www.mbari.org.
Zachos, J., Pagani, M., Sloan, L., Thomas, E., & Billups, K., 2001. Trends, rhythms, and aberrations in global climate
65 Ma to present. Science, v. 292, no. 5517, p. 686-693, www.essc.psu.edu.
Zanazzi, A., Kohn, M.J., MacFadden, B.J., & Terry, D.O., 2007. Large temperature drop across the Eocene-Oligocene
transition in central North America. Nature, v. 445, p. 639-642.
Zemtsov, V.A., 2007. Influence of earth rotation on continental motions. Journal of Gondwana Research, v. 12, no. 5,
p. 242-251.
Zhao, G., Sun, M., Wilde, S.A., & Li, S.Z., 2004. A Paleo-Mesoproterozoic supercontinent: assembly, growth and
breakup. Earth-Science Reviews, v. 67, nos. 1-2, p. 91-123.
Zhao, S., Lambeck, K., & Lidberg, M., 2012. Lithosphere thickness and mantle viscosity inverted from GPS-derived
deformation rates in Fennoscandia. Geophysical Journal International, v. 190, p. 278-292, doi:10.1111/j.1365246X.2012.05454.x.
Zhao, X., & Coe, R.S., 1987. Paleomagnetic constraints on the collision and rotation of North and South China. Nature,
v. 327, p. 141-144.
Zhou, Y., Nolet, G., Dahlen, F.A., & Laske, G., 2006. Global upper-mantle structure from finite-frequency surfacewave tomography. Journal of Geophysical Research, v. 111, B04304, doi:10.1029/2005JB003677.