ARTICLE INFORMATION

Author(s): K.M. Storetvedt
Publication: New Concepts in Global Tectonics Journal
Reference: V. 3, No. 4, December 2015. Page 516-528
KEYWORDS: academic laziness | fictitious knowledge | omissions | social pressure | wishful thinking

Geoscientific Urban Legends

DISCUSSION

“[A well-trained rationalist] will obey the mental image of his master, he will conform to the standards of argumentation he has learned, he will adhere to these standards no matter how great the confusion in which he finds himself, and he will be quite incapable of realizing that what he regards as the ‘voice of reason’ is but a causal after-effect of the training he had received.”

– Paul Feyerabend, in: Against Method (1988, p. 17)

“Whenever there is a breakthrough, a really important new discovery, this means that the experts have been proved wrong, and that the facts, the objective facts, were different from what the experts expected them to be.”

– Karl Popper, in: A World of Propensities (1990, p. 33)

Synopsis

According to Rekdal (2014, p. 638), “Many of the messages presented in respectable scientific publications are, in fact, based on various forms of rumours. Some of these rumours appear so frequently, and in such complex, colourful, and entertaining ways that we may speak of them as academic urban legends”. Rekdal further submits that the explanation for this phenomenon is usually that authors have “lazily, sloppily, or fraudulently” used selected second order sources – easily introducing false or twisted information which the review process has not discovered. The same disturbing pattern is readily exposed in the saga of continental drift/plate tectonics. Though critical facts contradicted Wegener’s drift hypothesis, which during the 1960s was ‘upgraded’ by plate tectonic mechanisms, a multitude of human factors were unable to stop it from an ultimate triumphal procession. With hindsight it is safe to say that this scientific transition was largely a fashion-like trendy contraption – being without significant problem-solving capacity. In this article, I discuss how a complex mix of early disregarded critical observations, professional alienation, competition and socio-political pressure gave rise to a kind of ‘unconscious’ professional state during the run up to the plate tectonic revolution. Henceforth, the Earth sciences had fallen victim of an overarching urban legend – a colourful story that during the last half century has brought global geology astray.

Keywords: academic laziness, wishful thinking, omissions, social pressure, fictitious knowledge

Faulty Towers of Science

There is a popular misconception that science is an impersonal, dispassionate, and thoroughly
objective enterprise. Whereas most other human activities are dominated by fashions, fads, and
personalities, science is supposed to be constrained by agreed rules of procedures and rigorous tests.
It is the results that count, not the people who produce them. This is, of course, manifest nonsense
[my italics].


Science is a people-driven activity like all human endeavor, and just as subject to fashion and whim.
In this case fashion is set not so much by choice of subject matter, but by the way scientists think
about the world. Each age adopts its particular approach to scientific problems, usually following
the trail blazed by certain dominant figures who both set the agenda and define the best methods to
tackle it”
.

(Davies, 1995)


This candid description of the human aspects of science, given above by physics professor Paul Davies, is indeed overwhelmingly demonstrated by the 20th century development of mobilistic global geology which in the late 1960s culminated in terms of the plate tectonic (PT) revolution. This ‘new global tectonics’ spread like an infectious epidemic and soon became the ruling dogma of the Earth sciences. In retrospect is not difficult to see that the new global hypothesis – being critically dependent on the validity of Wegener’s Gondwana assembly, was crooked right from the beginning. When palaeomagnetism in the mid-1950s re-opened the debate on continental drift, Wegener’s disregard of essential palaeoclimate observations from Antarctica remained overlooked and his Gondwana configuration taken for granted. Hence, in less than a decade lateral continental drift, along with its plate tectonic sub-mechanisms, developed into a penetrating urban legend – an all-embracing global model dominated by wishful thinking, auxiliary mechanisms and ad hoc fixes. In a recent essay in this journal – When global tectonics became a ‘pathological science’ (Storetvedt, 2014) – I described a number of important cases of counter evidence, negligence and misunderstandings which accumulated during the first decade following the plate tectonics revolution at the end of the 1960s. Here I discuss some critical cases of pervasive human weaknesses and professional alienation that became decisive turning points during the run up to this ‘coup d’état’.

For the growing number of young aspiring geophysicists in the early 1960s (including myself), the traditional arguments against Wegener’s continental drift were casually mentioned, but commonly regarded as out of date ‘tactics’ by the old boys. In the new progressive climate of opinion, it became important to give the ‘new global tectonics’ a polished façade. The answer was given, so nature had to adapt to the new promising view, not the vice versa. An advancing popular opinion will always be a very strong power although thinking in grooves may evolve in unclear ways depending not least on the wishes and persuasive powers of a few dominant individuals. Anyway, lateral continental drift was a glorious story to convey to my students; con-arguments were not saleable. Wegener’s book had traditionally caught little interest in Bergen, so in my time as a student and young researcher knowledge about continental drift had to be found in ‘second order’ texts – primarily from Arthur Holmes’ textbook Principles of Physical Geology (Holmes 1944, and later editions). When I many years later found du Toit’s book Our Wandering Continents (du Toit 1937) in the University library, having originally been part of a professorial book collection in the geological institute, it was characterized by not having been opened at all. Arthur Holmes and Alexander du Toit had been the two principal supporters of Wegener before World War II.

Wishful thinking prevails over conflicting evidence

Holmes presented Wegener’s model in a rather polished and sympathetic way, ignoring critical counter evidence. The Gondwana assembly was an essential element in the pre-Mesozoic continental configuration, with Antarctica in a pivotal position at the southern geographic pole. However, since the turn of the century, accumulated fossil and rock evidence for palaeoclimate had indicated strongly that Antarctica had had a prolonged Lower Palaeozoic to Lower Tertiary history of tropical to warm intermediate climate – not the ice-house conditions which Wegener, and later Holmes, took for granted. For the drift model, the wide range of palaeoclimate evidence from Antarctica was quite destructive, but both Wegener and Holmes basically ignored the increasing number of critical contradictory observations that had been collected by British, Swedish and Norwegian expeditions since the late 19 th century. Nevertheless, the chapter on continental drift in Holmes’ textbook seems to have been the principal reference source for the leading British palaeomagnetists when they opted for Wegener’s continental drift in the mid-late 1950’s (see also below).

Fig. 1. Permo-Carboniferous glacial centres of the southern hemisphere (white) as had been reported by the late 1920s; a) is after Wegener (1929) and b) is from Holmes (1954). The crucial aspect of the drift hypothesis was that these ice centres had once been united – forming a Gondwana continental assemblage, and subsequently split apart during post-Palaeozoic times. The glaciated Gondwana was thought to have been grouped around Antarctica which was given a nodal polar position despite the fact that all palaeoclimatic data from that land mass suggested tropical conditions during the Palaeozoic – and subtropical conditions as late as the Lower Tertiary. The fact that glacial traces had been reported also from the Northern Hemisphere was again disregarded. In other words, there was something fundamentally wrong with the drift-related Gondwana proposition. Note also the difference between the two presentations; though they both show Antarctica without glacial activity, the one by Holmes (b) gives the impression of having a persuasive intent.

In his efforts to match the contours of the adjoining Atlantic margins, Wegener (1929) had repeatedly been accused of having taken undue liberties with the crust; he had stretched and bent the crust as needed – or simply – eliminated parts of it, in order to achieve his preconceived continental match. In Wegener’s model, the South Atlantic was just an enormously widened rift the edges of which were once directly connected; if so, comparison of the geological structure of the adjacent continental margins ought to provide a clear-cut test. Wegener (1929) referred frequently to the work of du Toit (1927) who had done extensive work on the geological comparison of South America and Africa. Du Toit, however, concluded that in order to allow for differences in observed rock facies of the opposing continental margins a pre-drift separation of at least 400-800 km was required. This meant that the geological resemblance across the South Atlantic was not that obvious after all. If du Toit had to allow for a continental separation of 400-800 km or more, to obtain a ‘sensible geological match’ between the adjoining South Atlantic continents, how could he really set an upper limit to the suggested initial gap between the two land masses? Though Wegener in principle accepted du Toit’s reservations regarding the initial close continental contact, he nevertheless argued that the bilateral agreement of geological features across the South Atlantic “is just as if we were to refit the torn pieces of a newspaper by matching their edges” (Wegener, 1929 and 1966, p. 77). This was hardly anything but an emotional statement – not more than a socio-political attempt to bolster his pre-conceived Gondwana assembly.

On the other hand, Wegener (1929) seems not to have been completely unaware of the continental matching problems, such as those presented by the Azores region. In fact, he regarded the large Azorean massif as a real obstacle to a complete ‘closing-up’ of the Central Atlantic – prior to his alleged continental separation. He argued that geological information from these islands indicated that they had a continental substratum. From the work of Hartung (1860), it had been known for decades that pebbles of a variety of continental rocks – including granite, mica-schist, quartzite, sandstone and limestone – were scattered throughout the archipelago. In addition, the 1883 Talisman expedition, through deep sea dredging in a wide region just north of the islands, had obtained a multitude of continental metamorphics and Lower Palaeozoic fossiliferous sediments, so the accumulated rock evidence was indeed suggestive of a wider area around the Azorean archipelago being continental in origin. However, due to the revival of Wegenerian drift in the 1950s palaeomagnetic data combined with a complete closure of the Atlantic were in the forefront of attention; by now, taking up geological counter arguments were not in vogue and therefore not debated; one didn’t bother wasting time on ‘old-fashioned’ arguments. Wegener’s classical 1929-book was cited in practically every palaeomagnetic paper, though hardly anybody had read it (see below). If the book, along with relevant palaeoclimate evidence from Antarctica – readily available in the scientific archives – had been carefully studied, the fundamental problems with Wegener’s Gondwana configuration would most likely have been unveiled.

Palaeomagnetism embraces Wegener’s hypothesis – but its basic problems ignored

In Britain, the work of P.M.S. Blackett and S.K. Runcorn had led to the supposition that the origin of the main component of the geomagnetic field was closely tied to the Earth’s rotation; it was hypothesized that the main field, after averaging out transient (secular) variations, originated from a dipole aligned co-axially with the Earth’s rotation axis. Runcorn (1954 and 1955) explained the coincidence of the mean geomagnetic and rotational axes as due to the dominance of the Coriolis Effect in the outer fluid core. Granted the validity of the geocentric dipole field assumption throughout geological time, presuming that secular variation had effectively been averaged out during rock consolidation, a palaeomagnetic pole position would correspond to the location of the relative geographic pole for the geological epoch concerned. On this
basis, the swiftly increasing palaeomagnetic database of the 1950s showed time-progressive polar paths for individual continents – in addition to discrepancies between the polar curves for separate continents. These results brought continental drift and polar wandering back into the limelight of geophysical discussion. First of all, it seemed that the Earth’s body had changed its spatial orientation through geological time – for which quantitative reasoning by Gold (1955) endorsed additional support. But despite an unsatisfactory scatter in palaeomagnetic data the incongruity of the various polar paths suggested that some kind of relative continental motions had also taken place.

However, at around 1955 the Runcorn group was hesitant to accepting drift (Runcorn 1955). Thus, the view of Creer et al. (1957) “was that polar wander was the simpler hypothesis, involving fewer degrees of freedom than continental drift, and the one easier to reconcile with the present model of the Earth”. But data from other continents (than Europe) were complicating the picture. Comparison of palaeomagnetic poles from Britain and India suggested to some workers that significant relative drift had taken place (Blackett, 1956; Clegg et al., 1954 and 1956; Deutsch et al., 1958; Deutsch, 1958). A large-scale northward drift of India (away from Antarctica, cf. Fig. 1) was a significant element in Wegener’s hypothesized
southern continental assembly – Gondwana, and all of a sudden Wegener’s drift system was in the forefront of the discussion. In this conversion process, Creer et al. (1957) alluded to the possibility that Europe and North America might have been geographically closer in pre-Jurassic time than they are today. Furthermore, in a synthesis of global palaeomagnetic poles, Irving (1956) concluded that “prior to Tertiary times the pole has not only shifted its positions with respect to certain land-masses, but also that these land-masses have moved relative to one another”. In terms of continental mobility, only Wegener’s lateral drift system was not considered (see also below).

In other words, palaeomagnetism was beginning to sustain Wegener’s twin-hypothesis of continental drift and polar wandering. However, right from the early start there were problems in matching pre-drift segments of the polar curves – for attaining Wegener’s pre-drift unification of the land masses. But faced with the growing palaeomagnetic matching problems, why did not the leading British geophysicists of the 1950s consider alternative models for continental mobility? For example, the latitude-dependent inertia forces, associated with changes in Earth rotation, might well have affected the global lithosphere through a system of lithospheric torsions, thereby producing the observed discrepancies in polar wander paths, but this ready alternative was not thought of (K.M. Creer, personal communication March 1996). The
main reason for this lack of alternative thinking was most likely the competition between the Blackett (London) and Runcorn (Cambridge) groups; from their study in India, the Blackett group favoured large-scale continental drift, an idea that soon also the Runcorn group (from 1956 located in Newcastle) wholeheartedly endorsed. But this change of opinion, in favour of Wegener’s hypothesis, had a very shaky foundation – filled with unsettled problems.

Despite their preference for drift, the Blackett group was aware of an important ambiguity in their data (see below). If India had changed its azimuthal orientation by a significant in situ clockwise rotation, after cooling of the investigated Deccan lava sequence, an original downward pointing magnetization (in the Eurasian frame) would attain the present discrepant southward and upward pointing magnetization – thereby giving the wrong impression of having originally had a relatively high latitude southern hemisphere origin (see Storetvedt, 1997 and 2003 for detailed discussion). In May 1992, I had the opportunity to discuss the internal discussions and attitude within the Imperial College group with one of the principal investigators of the India project – the late Ernst (Ernie) Deutsch of Memorial University, St. John’s. What Ernie told me fits extremely well with the range of psychosocial and political factors in science described by Paul Davies – cited in the opening paragraph of the present paper. In my 2003-book Global Wrench Tectonics, I summarized my discussion with Ernie as follows:

Why, I had asked, had they preferred an extravagant northward translation of India instead of a simple in situ rotation, the latter requiring by far the least driving force? Ernie readily admitted that, in retrospect, he could clearly see that their preferred conclusion had been coloured by a combination of ignorance (on the geological part), prejudice, rivalry with the Newcastle group (which originally had been more inclined towards polar wander than drift) and geopolitics. He said that the workers in the Imperial College investigation were all physicists, and none of them had had anything but superficial insight into the geological arguments that Wegener and Du Toit had used in advancing their hypotheses of a large-scale
northward drifting of India. They just had taken for granted that the geological arguments applied were correct, he said. In any case, it was safer to associate the Deccan palaeomagnetic data with an old idea, albeit a controversial one, than to link them to the rotation alternative which most likely would only add further confusion and, hence, confer no credit on the workers involved. In his honest and thought- provoking remarks, Ernie went on to explain that in the mid -1950s palaeomagnetism was a new and minor geophysical discipline in need of a wider exposure. To associate rock magnetism with a revolutionary and long-debated scientific idea, perhaps triggering news headlines such as New Physical Measurements Confirm Controversial Geological Theory, would create just that boost the young research field so badly
needed.

The palaeomagnetic conclusion in favour of a mega-scale northward drift of India clearly had a shaky foundation, and yet it soon came to play a most decisive role in the trendy mobilistic edifice which poured ahead, and which during the next decade completely transformed the Earth sciences. Above all, it is important to acknowledge that palaeomagnetism did not provide conclusive evidence for large-scale translation of India as commonly alleged. Nevertheless, palaeomagnetism had turned global tectonics into a mobilistic business – broadly consistent with Wegener’s proposal. Hence, a formerly practically disbanded idea had, for falsified and misunderstood reasons, been imbued with new life and vitality. The traditional geological counterarguments, disfavouring a close pre-drift continental amalgamation, were as relevant as
before, but by now they had been bypassed and superseded by a new and supposedly more trustworthy geophysical discipline – palaeomagnetism. In hindsight, it is easy to see that by the late 1950s a new global tectonic paradigm was forcing its way forward; nothing could apparently stop it! The primus motor in this transformation process was S.K. Runcorn who with his ceaseless lecture tours around the world managed to gradually break down the traditional opposition to mobile continents, and in 1962 he edited a landmark volume entitled Continental Drift. This multi-authored book opened with Keith Runcorn’s major article ‘Palaeomagnetic Evidence for Continental Drift and its Geophysical Cause’, followed by one on ‘Palaeoclimatology and Continental Drift’ by Neil Opdyke. A simplified version of Runcorn’s polar trails is shown in Fig. 2.

Fig. 2. Schematic polar wandering curves simplified after Runcorn (1962).

To reconcile the global palaeomagnetic results, relative movements between the land masses were clearly needed. And because the majority of available data came from Europe and North America, the discrepancy between their polar paths was in the forefront of the discussion. Runcorn (1962) concluded that the data seemed broadly concurrent with Wegener’s hypothesis. However, there were problems in adapting the North Atlantic polar curves by simple lateral (E-W) continental adjustment. There was a tendency of corresponding polar estimates for the two continents, notably for the Palaeozoic, to having a latitudinal offset – North American poles were displaced to the south relative to the European ones. In hindsight, there was an indication a certain amount of relative clockwise ‘in situ’ rotation of North America relative to Europe; in that perspective, the southward fanning-out shape of the North Atlantic basin could be accounted for, but nobody seems to have noted this solution.

A decade later, Roy (1972) did a more systematic study of the polar curves for the North Atlantic continents confirming the early indication of North America having been subjected to a clockwise rotational swing relative to Europe. But by then lateral continental drift had become an unquestioned dogma, so when I, during discussions with colleagues, mentioned the alarming outcome of Jean Roy’s analysis the reaction was just a shrug of shoulders. Nevertheless, matching of other continental polar trails, according to the Wegener’s model, soon turned out to be a general problem. It was obvious that at least some continents had been subjected to ‘in situ’ rotation. For example, in addition to the favoured major northward translation of India, the London palaeomagnetists had to include a certain azimuthal rotation, and Australia had apparently undergone substantial (counter clockwise) rotation (Irving, 1957, 1958 and 1964). During a memorial symposium for Runcorn in Newcastle in February 1996, I discussed the many well-known physiographic and palaeomagnetic complications, versus classical continental drift, with Professor Ken Creer. He confirmed that palaeomagnetists had not explored alternative mobilistic schemes (other than Wegener’s) at the critical time in the late 50s and early 60s.

Mobility constraints disregarded

In the early 1960s there was apparently too much inertia in lateral continental drift to take notice of contrary observations; the opportunity of reopening the debate was, in many ways, over. Nevertheless, a number of authors had unveiled a series of kinematic paradoxes and other problems with conventional drift, without having captured other than superficial attention. Indeed, it is a paradox that once an idea has gained foothold in a science community, fundamental problems pertaining to that idea are either ignored or not seen at all. Nevertheless, Heezen (1959), referring to principal kinematic problems related to lateral drift, argued that “The locations of the Mid-Oceanic Ridge, oft cited as a remnant of the original continental rift, opposes continental drift since it seems to acquire that the continents drift in several directions at the same time. A possible way out of this dilemma is to postulate an expanding Earth”.

Earth expansion had also been the way out for Carey (1958); both Carey and Heezen referred to cases of major difficulty in the context of conventional drift. The supreme example was the geometric-kinematic riddles that arise for Antarctica; this continent is surrounded by oceanic ridges but is without any signs of undergoing compression. The kinematic-tectonic problems become even more acute for Africa and Antarctica combined – sometimes being referred to as the Africa/Antarctica paradox. However, if the Earth was undergoing expansion while remaining constant in size, newly added crust would have to be found in the oceans – most likely along mid-ocean rift valleys. If so, these ridges ought to constitute ridge-parallel dyke-in-dyke complexes, but subsequent deep sea drilling revealed that such two-dimensional intrusive complexes are practically absent in the deep sea crust (cf. Storetvedt, 1997 and 2003).

In his reconfiguration of the Earth’s crust, Wegener had put great emphasis on selected palaeoclimate observations. Thus, in connection with the rapidly growing palaeomagnetic data base, Runcorn convened a palaeoclimate conference in Newcastle, January 1963. From this conference I remember in particular arguments against a close pre-drift arrangement of the adjoining Atlantic continents brought up by the Norwegian palaeontologist and zoo geographer Professor Niels Spjeldnæs. By compiling the global distribution of Ordovician biological provinces he had concluded that the various fauna provinces basically represented climatic zones (Spjeldnæs, 1961). At the Newcastle conference he pointed out that Wegener’s pre-drift juxtaposition of North America and NW Africa was inconsistent with respect to palaeoclimate evidence: it did not make sense to place the Lower-Middle Palaeozoic tropical conditions of eastern North America in close contact with the corresponding palaeo-polar and glaciated region of NW Africa. In his demonstrated close relationship between global Ordovician fauna and climate, one of the Ordovician Polar Regions was in Central Sahara and the other in the north-central Pacific (Fig. 3). In his discussion, Spjeldnæs referred to a palaeomagnetic paper by Runcorn (1959) which ‘advertised’ continental drift, and in his Atlantic configuration he seems to some extent to have been influenced by that article. For example, in his relative position of South America he included NE Brazil within the polar region while the limited
glacial evidence came from the south-eastern Atlantic margin of the continent (cf. Fig. 1). Thus, a somewhat wider South Atlantic, but without its present southward fanning-out shape, would therefore be consistent with Spjeldnæs’ reconstruction.

Fig. 3. Reconstruction of the global palaeoclimatic system based on Upper Ordovician fauna provinces – simplified
after Spjeldnæs (1961). Filled symbols correspond to inferred tropical fauna. Note that the Atlantic continents (upper
figure) are placed closer together than in the current geographical situation although the biological data did not justify
it (see text). The corresponding palaeoclimate situation for the ‘Pacific hemisphere’ is depicted in the lower figure.
These palaeoclimatic reconstructions gave further support to the palaeomagnetic evidence for polar wander. Thus,
during the Middle Palaeozoic the relative geographical poles were located near the present equator – at around
Southwest Africa and in the Central Pacific respectively. At that time the tropical belt passed the present Arctic and
Antarctic regions. In other words, Spjeldnæs’ data indicated that after the Middle Palaeozoic the Earth’s body had
changed its spatial orientation (relative to the astronomical rotation axis) in the order of 90˚ of latitude – being broadly
consistent with the palaeomagnetic ‘N-S’ trending polar tracks for the Atlantic continents which had been devised for
the actual time range.

Spjeldnæs’ evidence was clearly at variance with the idea of a significant closure of the Atlantic; by placing the palaeo-pole in Sahara, and accepting a relatively wide Central Atlantic, the corresponding palaeo-equator would pass along eastern North America and in close proximity to the present north polar region and Central Asia for which there were ample evidence for hot-warm conditions during the Palaeozoic. Though Spjeldnæs’ global compilation of Ordovician fossil data was clearly at variance with the close pre- drift match of North America and NW Africa, I cannot remember that this important aspect triggered anything but casual comments; even at this non-tectonic conference, continental drift apparently formed such an important backdrop at the conference so bringing up contradictory opinion was apparently not appropriate. In fact, Spjeldnæs’ data could easily be interpreted in terms of a quasi-stable continental arrangement; moderate relative motions could have taken place, but, above all, his data clearly supported the classical phenomenon of polar wander. Since Ordovician time an accumulated 70°-90° of spatial shift of the whole planetary body (or perhaps some outer brittle layer) had taken place relative to the axis of rotation. In other words, Spjeldnæs’ data were consistent with a statement of Creer et al. (1957) a few years earlier that “polar wander was the simpler hypothesis, involving fewer degrees of freedom than continental drift, and the one easier to reconcile with the present model of the Earth”.

By the early 1960s, convection driven continental drift had become a hot topic and, in fact, this dual-hypothesis concept was in the initial stages of becoming a geoscientific urban legend. Looking into the question of driving forces, MacDonald (1964), using a combination of heat flow and surface gravity data, velocity information from surface waves, and observations of artificial Earth satellites, presented the first tentative evidence of regional differences between continental and oceanic upper mantles. MacDonald concluded: “The deep structure of continents places heavy restrictions on any theory of continental drift. A relative [lateral] motion of the continents must involve the mantle to depths of several hundred kilometres; it is no longer possible to imagine this continental blocks sailing over a fluid mantle”.

According to MacDonald, the idea of mantle convection, to most mobilistic geophysicists the only plausible cause of drift had run into problems. The continents were too thick and the mantle too rigid to permit lateral motions. The early suggestion that continents have deep roots into the upper mantle has later been amply demonstrated by modern mantle tomography. Taken at its face value, the presence of deep continental roots, and shallow oceanic anti-roots, put the very idea of mantle convection into doubt. The deep-rooted land masses could be interpreted in favour of being remains of an original pan-global continental crust while oceanic basins, with their corresponding mantle anti-roots, represented later crustal thinning and chemical transformation instigated by unevenly distributed mantle processes. The quasi-stationary continental arrangement arrived at by Spjeldnæs’ palaeo-climatic evaluation would fit rather well with such an evolutionary scheme.

Fictitious supporting evidence

In the early 1960s, continental drift was much more acceptable in Britain than in the USA, and under the auspices of the Royal Society a Symposium on Continental Drift was held on 19-20 March 1964 – convened by P.M.S. Blackett (London), E. Bullard (Cambridge) and S.K. Runcorn (Newcastle). Perhaps the most conspicuous contribution at this meeting was the paper entitled The fit of the continents around the Atlantic – by Edward Bullard and associates. During Runcorn’s NATO symposium in Newcastle a couple of weeks later, stubborn rumours would have it that the Cambridge group had “proven” that Wegener had been right by his pre-drift closure of the Atlantic. The fact that Wegener had considered the vast Azores Plateau as a barrier to a complete closure of the North Atlantic was not mentioned; in retrospect, this neglect is not difficult to understand for, as mentioned above, at that time hardly anyone had seen or bothered to study Wegener’s book.

The contributions from the meeting of the Royal Society were published in book form a year later (Phil. Trans. Roy. Soc., 258, 1965). The article of Bullard and associates attracted enormous attention; all of a sudden traditional geological counter arguments became overshadowed. The great interest was primarily because the ‘least squares’ fit of Bullard et al. gave the impression of being a solid scientific product. The study was based on objective arithmetic methods, Euler’s theorem for motions on a sphere, in addition to using a digital computer which at that time was the latest thing. The legendary reconstruction of the Cambridge Group is shown in Fig. 4. However, if we take a critical look at the procedures of the Bullard- group, the physical meaning of their pre-drift Atlantic is effectively eliminated by two essential facts:

a) Land masses and topographic contours which fitted with the wishes of the drift sympathizers were retained, while topographic features that did not ‘fit in’ were excluded; in other words, they had removed what they ‘disliked’. For example, the major Iceland Plateau, the shallow Faeroe-Iceland-Greenland Ridge, Jan Mayen Ridge, Walvis Ridge etc. were disregarded because they were obstacles to the drift model. The large Rockall Plateau was however accepted because it could be adapted. In other words, the popular ‘mathematical configuration’ of Bullard et al. was based on a combination of an arbitrary data selection and wishful thinking.

b) For the ‘sake of clarity’, continental excrescences which stood in the way of any trans-Atlantic adaptation were disregarded. The vital question of what Wegener, and now also the Bullard group, had done with regions such as the larger part of Mexico, all of Central America, and the Caribbean was as unclear as before.

Fig. 4. The famous ‘computer fit’ of the adjoining Atlantic margins was based on questionable procedures and was completely without scientific value. The alleged continental fit is based on Bullard et al. (1965), for which black between-continent segments denote overlaps and gaps. See text for discussion.

Science historian Homer Le Grand writes that Bullard and co-workers “had massaged the data and though the term ‘garbage in, garbage out’ had not been coined, charge could be made against their use of a computer” (Le Grand, 1988, p. 204). What one had witnessed was nothing more than a game for the gallery. The attention surrounding new technology and a semblance of objectivity, using mathematical methods, had taken the concentration away from the real issues that remained unresolved just as before. However, the attacks on the work of the Cambridge group died away very quickly. The popularity of continental drift drew to ever greater heights, and because hardly anyone would take the chance of ending up on the outside, critics soon got self-imposed muzzle.

One may wonder why a prominent geophysicist like Edward Bullard could let himself enmesh in a survey that completely lacked sobriety with regard to the data base. Was it all staged to win easily acquired publicity? In any case, it was very easy to see that the investigation had absolutely no scientific value – in fact, it was a clear abuse of mathematics. In spite of this, the ‘Bullard fit’ became famous and for years it was one of the most frequently cited works in the geoscience literature. In his scientific autobiography, The Ocean of Truth, Henry Menard gives a few short recollections from his research stay at the institute in Cambridge (located on Madingley Rise) – just before Bullard’s sudden interest in continental drift. Menard (1988, p. 208) writes: “Everyone met and talked twice daily at tea and coffee, so anything noteworthy was known widely at once at Madingley. I do not remember anyone, let alone Teddy [Edward Bullard] making
any interpretation in which continental drift was important
”.

According to Menard, neither Walter Munk, who at the time had close cooperation with Bullard, can remember that continental drift was featured. The question now arising is whether the continental drift epidemic – a mentality change that was about to conquer the collective imagination among British geophysicists – had turned Bullard to become a ‘weathercock’ scientist? Researchers will gain honour and power by defending and supporting popular intellectual fashions; this was probably the case with Edward Bullard’s sudden attraction to continental drift. Often we become silent witnesses to how ‘turning-the coat-for-the-wind’ maneuvers can have a decisive influence on the general adherence to an idea – especially if people with professional standing are leading the way.

During the prosperity of the drift hypothesis in the early 60s, another prominent personality entered the arena: John Tuzo Wilson, professor of geophysics at the University of Toronto. With his early academic background and later work as a geologist in the Precambrian of Canada, he had got a fairly static view of global geology – totally in keeping with the traditional way of thinking among the majority of North American professionals: continental drift was not the answer to geological problems! As late as 1959 Tuzo Wilson supported the American version of the contraction theory (Wilson, 1959). He rejected mantle convection and continental drift as realistic mechanisms to account for the new structures which, during and after World War II, had been discovered on the seabed. With regard to the mid-ocean ridge system, he
wrote (Wilson, 1959, p. 5): “The lack of abandoned ridges and slow rate of the ridge’s volcanism suggest that it has been in its present position for a very long time, perhaps most of the Earth’s history”.

However, the continental drift epidemic pushed on, and just a year later Wilson (1960) sympathized with a mobilistic scheme: the theory of Earth expansion. Within a short time, he had abandoned the traditional continental fixity in favour of mobile land masses. Not only that, after another two years he had been ‘influenced’ by the theory of sea floor spreading (Wilson 1963), in addition to accepting Runcorn’s convection model as the actual driving mechanism for drift/spreading. The mid-ocean ridges had now become young, not old as he had suggested a couple of years earlier. If the crust of these ridges were newly formed and therefore relatively warm, their topographic elevation was naturally a product of Archimedes’ principle. It is a fact that Tuzo Wilson, at an age of more than 50, suddenly gave up his traditional scientific platform, and in just three years emerged with two new conceptual outfits. Would such a major shift of emphasis be a natural thing for a scientist of his age, or did he just turn towards the most promising wind direction?

Tuzo Wilson had a prominent position within the geological sciences. During the years 1957-63 he had been President and Past President of IUGG, received many awards and was also well-established as an international lecturer. With his prominent status and freedom of speech he had unusually large ‘elbowroom’. In the first half of the 60s he published several interesting articles related to the model of sea floor spreading. As a young and inexperienced researcher I found Wilson’s articles highly welcome contributions for my teaching. It was only many years later that I became aware that his conclusions might have been ruled by a combination of wishful thinking and the reinforcement syndrome.

One of Wilson’s articles was about the age of volcanic islands versus the sea floor spreading model; their ages were supposed to increase with increasing distance from the crestal zone of mid-ocean ridges (Wilson, 1963). He concluded that his study had shown that this prediction was correct. He proclaimed: “Convection […] introduces horizontal motion. This would slowly move volcanoes away from their sources on the mid-ocean ridge, so that those volcanoes would cease to be active. The sources would remain fixed over the vertical currents and would produce fresh volcanoes. Thus, in time, each source on the mid-ocean ridge would produce a chain of progressively older extinct volcanoes, or perhaps two such chains, one on either side of the mid-ocean ridge” (Wilson, 1963, p. 537). All of a sudden, he seems to have set aside both earlier and new objections against continental drift. Wilson had entered the mobilism train, and soon an ever growing number of Earth scientists jumped after him. For many years Wilson’s mid-ocean ridge/islands article became one of the most cited in the geoscientific literature, but how many had really taken time to critically evaluate his paper?

Wilson’s conclusion was simple and elegant on paper, but the reality is totally different. His assertion that young volcanoes are a characteristic feature along the central axis of mid-ocean ridges was obviously wrong. With the exception of Iceland and Jan Mayen there are hardly active volcanic islands along these ridge axes; in the Pacific for example, active volcanism may occur anywhere within the deep sea basins. Furthermore, even since the time of Charles Darwin it has been known that St. Paul’s Rock, located on the Central Atlantic Ridge, is non-volcanic – consisting mostly of mantle rocks brought to the surface in sold state and driven by tectonic forces. Nevertheless, St. Paul’s Rock was included in the selective data
base – only because these mid-ocean cliffs had to be young, according to Wilson. Menard in his autobiography wrote that Wilson’s plea that “the farther an ocean island is from a mid-ocean ridge, the older it is likely to be” was both vexatious and wrong. Menard further added that Wilson made a number of refutable errors, but despite of the many glaring problems he (Menard) admits of having been “unprofitably distracted into publishing a refutation of the significance of his data” (Menard 1986, p. 194).

Concluding perspective: Buying a pig in a poke

Since the late 1980s, I have argued that the continental drift/plate tectonic revolution in the late 1960s was little more than a quasi-scientific cultural invasion; to a large extent this relatively rapid unorthodox transition was driven by a combination of wishful thinking, the pressure to conform, the repetition principle, rubbing shoulders, and authoritarianism (cf. Storetvedt, 1997, 2003, 2010 and 2014). Critical evaluation of facts versus the hypothesis was in surprisingly short supply. In the late 50s, palaeomagnetism was in need of some kind of between-continent motion, in order to make sense of the discrepant polar wander paths. As Wegener’s lateral drift model was ready at hand, this long-held anarchic hypothesis was soon announced, by the leading British palaeomagnetists, as a major breakthrough for understanding Earth’s physiographic
and structural development. By the early 60s, the drift idea drove forward like an epidemic; in this upturn, researchers persistently ignored observational inconsistencies and behaved as if the foundation of the new belief system was entirely secure – though hardly anybody had a reasonable professional grip on the state of the art. The 1929-book of Wegener’s was frequently referred to but, as it later turned out hardly anybody had seen the book. In order of getting firsthand information on the academic laziness with reading primary basic publications, I took a trip to Newcastle and Edinburgh, in late November 2003, to meet David Collinson and Ken Creer – both working under Keith Runcorn in the critical years of the 50s, first in Cambridge and then in Newcastle.

I first went to see David Collinson in Newcastle, and we spent a full day talking about the early history of palaeomagnetism and its connection with Wegener’s hypothesis. He told that he had not seen Wegener’s book, neither in Cambridge nor in Newcastle, before it was reprinted in 1966. He thought the same had been the case with Keith Runcorn. I further asked David to read Ernie Deutsch story about the Deccan study of the Imperial College group and their controversy with Cambridge/Newcastle team (summarized above). I was glad to hear that David had no objection to what Ernie had told me. The following day I went to see Ken Creer in Edinburgh to discuss the heroic 50s in palaeomagnetic research. Ken argued that he and Eduard Irving had accepted Wegener’s continental drift at least a couple of years before Keith Runcorn. But he admitted that it was first of all Keith who, with his ceaseless worldwide travelling, reopened the debate on drift in Europe and North America and practically singlehandedly managed to break down the traditional resistance to the hypothesis. To my question whether he had read Wegener’s book in the 50’s, Ken denied it and remarked: “We did not have time chasing that old book in the library”. Ken further stated that they had primarily been interested in Wegener’s classical figures which they had found in the books of Arthur Holmes and Alex du Toit. In other words, they had frequently referred to Wegener’s book but without having read it; in doing so they had in fact committed an academic lie – by having plagiarized Wegener’s
hypothesis from the Holmes/du Toit books. In other words, the leading British palaeomagnetists had consulted only secondary literature written by careless drift sympathizers (notably with respect to the ignorance of the true palaeoclimate history of Antarctica) – and thereby having spread an ‘adorned’ but utterly false scientific message.

If one plagiarizes, but write in one’s own words, a claim that turns out to be completely lunatic, one gets not only a pig in a poke. But the lack of actual citations and consulted references implies that other readers will find seemingly independent texts which mutually confirm and reinforce each other. In that way, errors and fantasies may float around in academic publications building up tenacious urban legends. “A good reason for avoiding the use of secondary sources in academia is that messages that pass through several links have the unfortunate tendency to become modified or altered along the way, as in the whisper game […] Hypothesis, assumptions, and suggestions can in such a way be transformed into knowledge and scientific facts” (Rekdal 2014). The current state of affairs is probably worse than ever; publication pressure and competition for resources have clearly led to a careless and sloppy working culture (see Bauerlein et al. 2010). For example, today many Earth science publications are overloaded with references – sometimes filling up whole paragraphs of journal articles, but it often does not take much effort to unveil that the majority of the cited papers have not been read. Exaggerated references are actually a form of academic littering unless one, on the basis of these references, brings something new into the discussion.

Nevertheless, during the 1960s Wegener’s undocumented continental drift, and then its successor plate tectonics, became a global tectonic pop-wave sweeping across the geosciences. Fairly soon it was meritorious only to claim support of the new model, and from the 1970s onwards critics were frequently sat aside and ridiculed. The PT bandwagon became the guarantor of geoscientific modernity and soon was so captivating that it became shielded from refutation by even the strongest counter evidence (cf. Storetvedt, 1997 and 2003). Testing the theory’s explanatory capability and its capacity to link diverse geological phenomena did not turn out as expected, but even so the increasing load of disheartening complications was sidelined. Actually, when a new set of basic rules and mechanisms (a paradigm) achieves to become widely accepted, it will eventually become inextricably associated with the complicated network of socio-politics, pressure to conform, power, honour, and professional identity – integrated components in any intellectual activity. Numerous research careers may even get to stand and fall with the theory’s validity.

For these reasons, even the worst of ruling theory get the cat’s nine lives because few will have the courage to openly admit that they have backed the wrong horse – and thereupon having bought a pig in a poke. For that reason the geoscientific community at large has, for decades, been restrained by a kind of ‘Frankenstein Monster’ who refuses to give up its dominance. Therefore, it will be up to a new generation of scientists to clean up the mess that has accumulated by empirical data having been manipulated into contexts in which they do not belong. Such an utter confusion is readily demonstrated by the uncoordinated ad hoc situation characterizing present day PT-dominated research activity (Fig. 5). Despite the major geo-technological advances during the last few decades true phenomenological understanding has not been achieved; all prominent questions in global geology have remained incoherent and unsettled.

Fig. 5. Guided by a faulty maxi-theory, real scientific progress – in terms of coordinated phenomenological understanding – may come to a complete halt. Thus, during the reign of plate tectonics Earth science books and journals have been filled up with a confusing mix of true and theory-laden observations, from which interpretations and conclusions have given rise to a heavy load of non-contiguous and baffling ad hoc propositions.

A functional (realistic) theory make the information flow manageable, meaning that it serve to create unity from a phenomenological diversity. Without such an operating thought construction the flow of new data may become impossible to handle, and the whole science in question may suffer serious setbacks. Regarding global geology, both space geodetic techniques and palaeomagnetism give strong evidence in favour of some kind of between-continent motion. But after decades of unsuccessful data adjustment, time is long overdue to conclude that the lateral drift system is not working. In fact, it was Wegener’s wishful thinking about his preferred Gondwana continental assembly that eventually brought global tectonics astray; if he had taken the fossil and rock evidence for long-term tropical to subtropical climate in Antarctica at their face value, his southern hemisphere continental unification would have got a powerful shot across the bow. In that case, it is likely that the British geophysicists would have looked for an alternative mobilistic scheme, than lateral drift, to account for the observed discrepancies between polar wander curves of individual continents. And if so, global tectonics would most likely have got a completely different development. The moral is: honesty, directness and simplicity are the true way forward for science.

References

  • Bauerlein, M. et al., 2010. We Must Stop the Avalanche of Low-Quality Research. The Chronical of Higher Education, June 13, 2010.
  • Blackett, P.M.S., 1956. Lectures on Rock Magnetism, Jerusalem, Weissman Science Press, 131p.
  • Clegg, J.A., Almond, M. and Griffiths, D.H., 1954. Some recent studies of the pre-history of the Earth’s magnetic field. J. Geomag. and Geoelec., v. 6, p. 194-199.
  • Clegg, J.A., Deutsch, E.R. and Griffiths, D.H., 1956. Rock magnetism in India. Phil. Mag., v. 1, p. 419-431.
  • Creer, K.M., Irving, E. and Runcorn, S.K., 1954. The direction of the geomagnetic field in remote epochs in Great Britain. J. Geomag. Geoelec., v. 6, p. 163-168.
  • Creer, K.M., Irving, E. and Runcorn, S.K., 1957. Geophysical interpretation of palaeomagnetic directions from Great Britain. Phil. Trans. Roy. Soc. London, v. A250, p. 144-156.
  • Davies, P., 1995. From Special Introduction to R.P. Feynman: The Fundamentals of Physics Explained, 3rd Ed. London, Penguin Books, 146p.
  • Deutsch, E.R., 1058. Recent palaeomagnetic evidence for the northward movement of India. J. Alberta Soc. Petrol. Geol., v. 6, p. 155-162.
  • Deutsch, E.R., Radakrishnamurty, C. and Sahasrabudhe, P.W., 1958. The remanent magnetism of some lavas in the Deccan Traps. Phil. Mag., v. 3, p. 170-184.
  • Du Toit, A. and Reed, F.R.C, 1927. A geological comparison of South America with Africa. Publ. of the Carnegie Institution of Washington, no. 381.
  • Du Toit, A., 1937. Our Wandering Continents. Edinburgh, Oliver & Boyd, 366p.
  • Feyerabend, P.K, 1988. Against Method. New York, Verso, 296p.
  • Gold, T., 1955. Instability of the Earth’s axis of rotation. Nature, v. 175, p. 526-529.
  • Hartung, G., 1860. Die Azoren in ihrer ausseren Erscheinung und nach ihrer geognostisher Natur. Leipzig, Engelman, 329p.
  • Havlíček, V., 1974. Some problems of the Ordovician in the Mediterranean region. Vestnik Ustred. Ust. Geol., v. 49, p. 343-348.
  • Holmes, A., 1944-54. Principles of Physical Geology. Edinburgh, Thomas Nelson Ltd, 532p.
  • Irving, E., 1956. Palaeomagnetic and Palaeoclimatological aspects of polar wandering. Geofis. Pura et Applicata, v 33, p. 23-41.
  • Popper, K., 1990. A World of Propensities. Bristol, Thoemmes Press, 51p.
  • Rekdal, O.B., 2014. Academic urban legends. Social Studies of Science, v. 44, p. 638-654.
  • Runcorn, S.K., 1954. The Earth’s core. Trans. Am. Geophys. Un., v. 35, p. 49-63.
  • Runcorn, S.K., 1955. Rock magnetism – geophysical aspects. Advances in Physics, v. 4, p. 244-291.
  • Runcorn, S.K.., 1962 (editor). Continental Drift. London, Academic Press, 338p.
  • Spjeldnæs, N., 1961. Ordovician climatic zones. Norsk Geol. Tidsskrift, v. 41, p. 45-77.
  • Storetvedt, K.M., 1997. Our Evolving Planet. Bergen, Alma Mater, 456 p.
  • Storetvedt, K.M., 2003. Global Wrench Tectonics. Bergen, Fagbokforlaget, 397p.
  • Storetvedt, K.M., 2010. Falling Plate Tectonics – Rising new Paradigm: Salient Historical Facts and the Current Situation. NCGT Newsletter, no. 55, p. 4-34.
  • Storetvedt, K.M., 2014. When global tectonics became a ‘pathological science’. NCGT Journal, v. 2, p. 106-121.
  • Wegener, A., 1929. The Origin of Continents and Oceans. Reprinted 1966. Northampton, Dover Publications, 248p.

other articles

Når jorden skjelver

Når jorden skjelver

Jordskjelvkatastrofen i Italia 6 april 2009 minnet oss på nytt om de sterke latente krefter i jordens indre. Men hva denne energien egentlig består i – hvordan den bygges opp og utløses som plutselige kaskader av elastiske bølger, er ikke...

read more
Kritikk av lærebøkenes feilinformasjon om jorden

Kritikk av lærebøkenes feilinformasjon om jorden

Læreplanen i geografi (studieforberedende) fremhever at undervisningen skal være motiverende og stimulerende og ...medvirke til forståelse og mestring av den store mengden av informasjon i vår tid. Mer kunnskap motiverer og stimulerer evnen til...

read more