Skip to main content

Main menu

  • Home
    • Series home
    • Lyell Collection home
    • Geological Society home
  • Content
    • Online First
    • Current volume
    • All volumes
    • All collections
    • Supplementary publications
    • Open Access
  • Subscribe
    • GSL fellows
    • Institutions
    • Corporate
    • Other member types
  • Info
    • Authors
    • Librarians
    • Readers
    • Access for GSL Fellows
    • Access for other member types
    • Press office
    • Accessibility
    • Help
  • Alert sign up
    • RSS feeds
    • Newsletters
  • Propose
  • Geological Society of London Publications
    • Engineering Geology Special Publications
    • Geochemistry: Exploration, Environment, Analysis
    • Journal of Micropalaeontology
    • Journal of the Geological Society
    • Lyell Collection home
    • Memoirs
    • Petroleum Geology Conference Series
    • Petroleum Geoscience
    • Proceedings of the Yorkshire Geological Society
    • Quarterly Journal of Engineering Geology and Hydrogeology
    • Quarterly Journal of the Geological Society
    • Scottish Journal of Geology
    • Special Publications
    • Transactions of the Edinburgh Geological Society
    • Transactions of the Geological Society of Glasgow
    • Transactions of the Geological Society of London

User menu

  • My alerts
  • Log in
  • Log out
  • My Cart

Search

  • Advanced search
Geological Society, London, Special Publications
  • Geological Society of London Publications
    • Engineering Geology Special Publications
    • Geochemistry: Exploration, Environment, Analysis
    • Journal of Micropalaeontology
    • Journal of the Geological Society
    • Lyell Collection home
    • Memoirs
    • Petroleum Geology Conference Series
    • Petroleum Geoscience
    • Proceedings of the Yorkshire Geological Society
    • Quarterly Journal of Engineering Geology and Hydrogeology
    • Quarterly Journal of the Geological Society
    • Scottish Journal of Geology
    • Special Publications
    • Transactions of the Edinburgh Geological Society
    • Transactions of the Geological Society of Glasgow
    • Transactions of the Geological Society of London
  • My alerts
  • Log in
  • Log out
  • My Cart
  • Follow gsl on Twitter
  • Visit gsl on Facebook
  • Visit gsl on Youtube
  • Visit gsl on Linkedin
Geological Society, London, Special Publications

Advanced search

  • Home
    • Series home
    • Lyell Collection home
    • Geological Society home
  • Content
    • Online First
    • Current volume
    • All volumes
    • All collections
    • Supplementary publications
    • Open Access
  • Subscribe
    • GSL fellows
    • Institutions
    • Corporate
    • Other member types
  • Info
    • Authors
    • Librarians
    • Readers
    • Access for GSL Fellows
    • Access for other member types
    • Press office
    • Accessibility
    • Help
  • Alert sign up
    • RSS feeds
    • Newsletters
  • Propose

The Lewisian Complex: insights into deep crustal evolution

J. Wheeler, R. G. Park, H. R. Rollinson and A. Beach
Geological Society, London, Special Publications, 335, 51-79, 1 January 2010, https://doi.org/10.1144/SP335.4
J. Wheeler
1Department of Earth and Ocean Sciences, Jane Herdman Building, Liverpool University, Liverpool L69 3GP, UK
  • Find this author on Google Scholar
  • Search for this author on this site
  • For correspondence: [email protected]
R. G. Park
212 Provost Ferguson Drive, Tain, Ross-shire, IV19 1RE, UK
  • Find this author on Google Scholar
  • Search for this author on this site
H. R. Rollinson
3Department of Geographical, Earth and Environmental Sciences, University of Derby, Kedleston Road, Derby DE22 1GB, UK
  • Find this author on Google Scholar
  • Search for this author on this site
A. Beach
4Exploration Outcomes, 1 Huntly Gardens, Glasgow G12 9AS, UK
  • Find this author on Google Scholar
  • Search for this author on this site
PreviousNext
  • Article
  • Figures & Data
  • Info & Metrics
  • PDF
Loading

Abstract

The Lewisian Complex is an Archaean/Proterozoic craton fragment found in NW Scotland and throughout the Outer Hebrides. The 1907 memoir recognized, simply from field relationships and petrographic observation, key features of Lewisian evolution. The bulk of the Lewisian is an old, deformed complex consisting mainly of acid igneous rocks, with some basics, ultrabasics and metasediments. In the Central District of the mainland these are pyroxene bearing (now recognized as granulite facies). The Lewisian Complex was intruded by a suite of basic and ultrabasic dykes which show variable states of later deformation, the intensity of strain being correlated with the development of hornblende schist in the dykes and amphibolite facies assemblages in the country rocks. In the Northern and Southern Districts, this deformation is pervasive and the dykes become concordant hornblende schist sheets. The new foliation with transposed dykes and metasediment sheets is then folded around NW–SE axes. Today there is no single agreed model for the evolution of the complex but an outline is as follows. In the pre-dyke (Scourian) history, subduction led to melting of oceanic crust which provided vast volumes of tonalite-trondhjemite-granodiorite in the period 3100–2700 Ma. Ages show geographic variations but it is not proven whether that implies large displacements between pieces of crust or whether it represents intrusions into other intrusions. The subcontinental lithospheric mantle dates from c. 3000 Ma. K, U and other large ion lithophile elements are depleted in the Central District of the mainland; this is due to depletion in the downgoing oceanic slab which in turn is a result of dehydration prior to melting. Other areas are not depleted in such elements, so various tectonic settings were involved. Remnants of metabasic material in the Lewisian may be relics of oceanic crust. Granulite facies metamorphism with, in places, P>10 kb and T>1000 °C occurred a considerable time after intrusion so is not necessarily linked to igneous events. This ‘Badcallian’ episode affected mainly the Central District and a part of the southern Outer Hebrides; other areas show only amphibolite facies. Zircon dating indicates two high-grade events at 2500 and 2700 Ma. During the ‘Inverian’ episode a series of wide amphibolite-facies shear zones affected the granulite-facies Scourian gneiss prior to the intrusion of the Scourie dykes. The Scourie dykes were intruded from 2400–2000 Ma and are largely quartz tholeiites derived from enriched subcontinental lithospheric mantle; there are some picrites which yield the oldest ages but are also seen to crosscut basic dykes. The dykes trend NW–SE and are steep where not affected by later deformation except where they intrude along, and are controlled by, Inverian fabrics. Post-dyke (Laxfordian) history involves the development of calc-alkaline igneous rocks in the Outer Hebrides and mainland (c. 1900 Ma). Volcanics associated with sediments younger than 2000 Ma comprise an accretionary complex formed in a subduction setting; they are now intercalated between slabs of Archaean basement indicating that the complex was involved in collision with continental crust. Huge strains transposing dykes and country rocks affected almost all of the Outer Hebrides and the mainland except for the Central District. The NW–SE trending lineation indicates the collision direction; the metasediments on the mainland and the South Harris Igneous Complex may mark a folded suture between two continents. Metamorphism was amphibolite facies almost everywhere; in South Harris it was granulite facies at c. 1880 Ma. At 1750–1675 Ma, a distinct event, called late Laxfordian but much younger than earlier Laxfordian metamorphism and with a distinct tectonic setting, caused folding of the previous structures along NW–SE axes, migmatization and renewed amphibolite facies metamorphism.

In wandering over the region of the Lewisian gneiss in the north-west of Scotland, it is impossible to avoid being forcibly impressed with the resemblance between the architectural features of the rocks and the forms and dispositions of the foam-flecks on the pools of comparatively still water below falls and rapids. In the spaces where little or no movement is going on, irregular and rounded masses of foam are separated by dark areas of still water. On the margins of these spaces, where movement is in progress, they are drawn out first into lenticles and then into thin streaks which may remain parallel or be bent, by subsequent movements, into serpentine folds or complicated convolutions which defy analysis, and resemble the damascening of old sword-blades and gun-barrels. (Peach et al. 1907, p. 71)

This passage is not only the most poetic, but also quite possibly the first summary we have of the effects of overprinting on structures. It refers to the heterogeneous reworking of a suite of Proterozoic dykes, as seen in all scales from outcrop to 100 km, one of the many features of the Lewisian which have been recognized as widespread phenomena in Earth's evolution. The Lewisian Complex is a fragment of the Precambrian Laurentian craton and is made of Archaean to Proterozoic metagranitoids, metabasites and subordinate metasediments. Its small size relative to other Precambrian cratons belies its significant influence on research into basement evolution and crust formation in general. The influential 1907 memoir edited by Peach et al. documented the setting of the Lewisian rocks: late Proterozoic arkoses (the Torridonian) lie above the Lewisian on a profound undulating unconformity. Above the Torridonian, again unconformable, is a Cambro-Ordovician shelf sequence. All three rock units are caught up in the Caledonian Moine Thrust Zone, which carries the Neoproterozoic metasediments of the Moine Supergroup in its hanging wall. These other rock units are discussed elsewhere in this Special Publication.

Figures 1 and 2 show the main outcrops of the Lewisian, excluding only the southernmost islands of the Outer Hebrides (including Barra) and Coll and Tiree further south. The inliers in the Caledonian are not reviewed here. They have broadly the same composition and age as the Lewisian Gneiss (Friend et al. 2008), with the exception of the Glenelg-Attadale inliers (Storey 2008) which are anomalous in terms of metamorphic grade (eclogite facies) and age (as young as Grenville in part).

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Geography and outline of structural and metamorphic state of the Lewisian Complex. The southernmost outcrops on Barra, Coll and Tiree are omitted. Younger rocks include the late Precambrian Torridonian; the Moine and the Cambro-Ordovician shelf sequence. Islands are named in capitals, other geographic features in lower case and geological features in italics. Lakes marked in blue.

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Protoliths and selected protolith and metamorphic ages. Protolith ages are in italics and are for TTG gneisses unless indicated specifically by G, distinct granite body; SD, Scourie dyke; B, metabasic; MS, metasediment; D, diorite; or A, anorthosite. Metamorphic ages in blue upright lettering. Ages are U–Pb (mainly on zircon) except where indicated – see Table 1 for details.

In this contribution we will examine how the memoir presented the Lewisian Complex, pursue the development of ideas since then (so the reader may appreciate the context of our present day knowledge), summarize the evolution, and finally offer opinions on key problems and ways forward in Lewisian research. We give much attention to field and structural aspects since these are prominent in the memoir and are pivotal to present day understanding. Our review of the huge geochemical and geochronological progress made in the latter part of the 20th century will be relatively brief but we will highlight gaps and overlaps between the views provided by these methods and those based on structural and tectonic ideas. Other reviews of the Lewisian are available in two books edited by Park & Tarney (1973, 1987) and in Park et al. (2002), the igneous aspects in MacDonald & Fettes (2007), the Laxford Shear Zone (Fig. 1) in Goodenough et al. (2010).

The Lewisian Complex as understood in the Peach et al. 1907 memoir

The Survey geologists relied on field relationships to establish relative timings. There were no radiometric dates available to them although the idea of using radioactive products to date rocks had been proposed recently (Rutherford 1905). The memoir recognized the existence of very old rocks with a complicated history predating unmetamorphosed Torridonian sediments. For example in relation to the Stoer area (Fig. 1). This district furnishes abundant evidence that the pre-Torridonian movements which led to the reconstruction of the gneiss and dykes were completed before the deposition of the Torridon sandstone. For example, at Stoer, where there is an important outlier of that formation, the numerous shear zones can be traced up to its margin, while pebbles of deformed gneiss and hornblende-schist occur in the bands of conglomerate which there appear. (p. 170)

The ‘reconstruction’ referred to, which would now be described as the Laxfordian event or events, was itself identified purely via field relationships seen within the Lewisian Complex. We choose five themes which show the way the Survey worked and set the scene for later developments.

Protoliths

These were diagnosed from the chemical and mineralogical affinities of Lewisian rocks, but also from the shapes and relationships of bodies of rock in the field. A ‘Fundamental Complex’ was intruded by a ‘great series’ of dykes and sills. All the major rock types that we know in the Lewisian today were recognized. In the Fundamental Complex, ‘Rocks of probably Plutonic Origin’ (Peach et al. 1907, p. 33) comprise: ultrabasic rocks, such as pyroxenites; pyroxene gneisses with some feldspar and quartz; hornblende rocks with feldspar almost always present; quartz-feldspar-biotite gneisses; and muscovite-bearing gneisses. Basic rocks are usually older than acidic ones in outcrops showing diagnostic relationships (Peach et al. 1907, p. 73). Still in the Fundamental Complex, ‘Altered sediments’ include garnet schists, graphitic schists and calcareous rocks. The ‘Later Intrusions (Dykes and Sills)’ include ultrabasic through acidic material. In some regions of the Lewisian, the Survey found ‘ … the ultrabasic and basic intrusions preserve their dyke-like character, showing clearly that they rose along vertical or highly inclined fissures, after the development of early banding or foliation in the gneisses of the Fundamental Complex’ (Peach et al. 1907, p. 36). Thus, the petrology and shape of these bodies together pointed to their igneous origin. Cross-cutting relationships again give timing, and the ultrabasic dykes are usually later than the basic ones (e.g. Peach et al. 1907, map on p. 163). These ‘Later Intrusions’ are now referred to as the Scourie dykes.

Foliations and lineations

In the memoir banding was commonly interpreted as the result of deformation, though sometimes in a partially molten state, motivated by comparisons with plutons such as Criffel in the Southern Uplands of Scotland. ‘Basic inclusions are here extremely common, and they have been drawn out into lenticles and bands by differential movement which has affected also the later acid veins’ (Peach et al. 1907, p. 73) – the Lewisian commonly looks similar.

The memoir contains many references to lineations. For example, in a granite the lineation ‘is often more or less parallel to the foliation of the gneiss’ (Peach et al. 1907, p. 108); in Scourie dykes ‘a linear foliation or rod-structure replaces the plane-parallel foliation’ (Peach et al. 1907, p. 120); in gneisses in the Laxford area, The minerals on the foliation planes of the granulitic gneisses are often arranged with their long axes parallel. The direction of elongation, or stretching, varies in different localities, but near the same thrust is persistent for long distances. Over the greater part of the district, when the observer looks north towards the foliation planes, the lines of elongation appear diagonal, about half way between the directions of strike and dip of the planes, and with their lower ends on his right hand. (Peach et al. 1907, p. 138; our italics)

The last quote shows a sharp awareness of the three-dimensional (3D) nature of lineation (though the lack of a succinct way of expressing it), and also implies that the origin of such lineations is through stretching – a modern interpretation, though at the time there was no precise way in which to pursue the implications. In places on the published one-inch maps, the horizontal projections of the lineations are marked, with a note indicating that these should be used together with marked dips to determine the 3D orientation (Peach et al. 1907, p. 208).

A final quote shows that the possibility of solid-state deformation was appreciated. In relation to rocks near Gairloch (Peach et al. 1907, p. 98), These facts prove that the rock was either intruded during movement and consolidated in its present form, or that a previously formed rock was entirely recrystallised under the influence of the stresses which produced the ‘rodded’ or ‘mullion-structure’ of the district.

Quantitative structural geology, as we know it, was not available to assist in interpreting such data – but the next two sections show how deep insight was gained from the general patterns of deformation.

Dykes as time markers: small-scale relationships

Perhaps the most significant contribution that the memoir makes in regard of the Lewisian relates to timing of deformation. A single sentence captures that contribution. The detailed mapping [of the Lewisian] has shown that, after the eruption of the ultra-basic, basic and intermediate dykes, the whole area was in pre-Torridonian time subjected to earth movements which, affecting both the various members of the Fundamental Complex and also these later intrusions, gave rise to rapid plication [folding] of these strata and to lines of disruption or shear-zones, accompanied or followed by re-crystallisation of the original constituents, development of foliation, and occasional mylonitisation of the rocks. (Peach et al. 1907, p. 36)

There are several perfectly recognizable modern concepts here. It is crucial to be aware that this reworking is heterogeneous on all scales, and therefore is commonly visible within a single dyke. This theme appears many times in the memoir. For example, ‘Transitions from massive epidiorite [hornblende-plagioclase rock] to foliated hornblende-schist may take place within the limits of a hand specimen or even a microscopic section’ (Peach et al. 1907, p. 89; see also Teall 1885). On a still local, but larger, scale, we have (Peach et al. 1907, p. 150), A mile and a third E.S.E. of Ben Auskaird [itself 7 km SE of Scourie], the gradual bending, thinning and production of foliation in a broad dyke near a line of disruption can be clearly seen. About 20 to 30 yards on either side of the thrust, the dyke averages 50 yards in width, but in the deflected portions it is only three or four feet.

Note that 1 yard=0.9 m, 1 foot=0.3 m. Such deformation also affects the country rock: ‘gneisses next to the intrusive rock have been involved in the changes of dip and structure’ in some places (Peach et al. 1907, p. 150). Elsewhere, for example in what we now call the Canisp Shear Zone, reworking within gneisses is pictured (Fig. 3), and ‘ … between the lines of disruption, the folia of the older gneiss are sharply contorted and dragged into parallelism with the disruption planes’ (Peach et al. 1907, p. 167).

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Sketch of fabrics from near Achmelvich, from the memoir. The original caption reads ‘Relation of newer to older planes of foliation in pyroxenic and hornblendic gneiss’. (Peach et al. 1907, fig. 8 on p. 1670).

Dykes as time markers: reworking and large scale structure

These small-scale observations must have helped promote the idea that bodies made entirely of hornblende schist, perhaps with gneissic banding parallel to their margins, were once intrusions. This interpretation leads, then, to a picture of heterogeneous deformation of the whole Lewisian outcrop. In terms of large-scale structure, this modification is not seen everywhere – thus a central district, from near Scourie to Loch Broom, is described as unmodified, with easily recognizable dykes. In the northern district between Laxford and Cape Wrath, ‘The basic dykes … appear in the form of bands of hornblende-gneiss, which can be followed for only short distances’ (Peach et al. 1907, p. 37). In the southern district, from Gruinard Bay to Loch Torridon and Raasay, basic dykes appear ‘either in the massive form or as hornblende-schists’ (Peach et al. 1907, p. 38). There seems to have been little doubt that sheets of hornblende schist could be correlated with dykes of the central district. These three districts remain a simple and useful division of the mainland and we will use them, in capitals to indicate they are the original definitions.

Large scale structures were also recognized, particularly in the Southern District, by tracing sheets of metasediment and/or hornblende schists around folds. Since those schists themselves represent deformed dykes, and those dykes themselves intruded sometimes banded rock of the Fundamental Complex, there are effectively at least three stages of deformation documented in the memoir.

Metamorphism

There was, at the time of the memoir, no theoretical framework for interpreting metamorphic mineral assemblages. Although the theory of thermodynamics had been developed in the 1870s (collected works in Gibbs 1906) it would be decades before it was applied in geology and even the facies concept was yet to appear (Eskola 1915). The account shows that the metamorphism was understood in terms of the rearrangement of material within individual rocks and, in some cases, bulk addition and removal. Note that the term ‘granulitic’ is used in a purely textural sense in the memoir, so when read in the quotes below it should not be confused with the modern concept of granulite facies.

Many types of gneiss that are characterized by pyroxene or hornblende correspond very roughly to granulite and amphibolite facies respectively. Quartz, when present in the granulite facies rocks, shows a distinctive blue or opalescent colour and in thin section has inclusions as ‘rows of minute dots’ and ‘extremely thin hairs’ (Peach et al. 1907, p. 54). Teall, doubtless prompted by his earlier observations (Teall 1885) notes that The rocks with hornblende occurring in fibrous or other aggregates … are common in the central zone in which the pyroxene-gneisses abound, and they are connected with these in such a manner as to suggest that the hornblende is in many, if not in all cases a secondary product after pyroxene.

This theme is reiterated in many places and is a description of amphibolite facies overprinting granulite facies minerals. So, on p. 138 (Peach et al. 1907) we find … the less altered gneisses near most of the pre-Torridonian thrusts are chiefly those with pyroxene or with aggregates of hornblende and biotite, which have probably replaced the pyroxene, and from these most of the granulitic biotite gneisses have evidently been derived. The biotite must have been chiefly formed from pyroxene or hornblende, and the white mica from feldspar.

Such changes may be associated with shearing: in the Canisp shear zone (Peach et al. 1907, p. 167) ‘… the granules of quartz lose their opalescence, are elongated and become granulitic, biotite is developed from the hornblende, and the feldspars are granulitised.’ Changes are documented in gneisses and dykes, both basic and ultrabasic. Thus, regarding a picrite dyke near Lochinver, ‘A series of specimens … show progressive alteration into a perfect schist composed of chlorite, pale hornblende and talc’ (Peach et al. 1907, p. 170).

Many observations made in passing are clear enough to be interpreted today. For example, in basic rocks north of Loch Laxford ‘Garnets … frequently occur in small aggregates surrounded by a thin white rim of feldspar granules’ (Peach et al. 1907, p. 120, see also p. 210 etc.). These are reaction rims and could relate to decompression. In the authors' experiences such textures are widespread in the Scourie dykes, though a detailed study is yet to be made. In gneiss from Ard Ialltaig, near Gairloch, a dark fine-grained rock shows ‘white spots, about the size of peas, with curved tail-like processes’ (Peach et al. 1907, p. 198): a tantalizing note given the general absence of kinematic indicators in the Lewisian. Even grain shape and crystallographic fabrics are documented: in a Scourie dyke from near Ben Stack, hornblendes are ‘arranged with their longest diameters parallel to the direction of stretching’ (Peach et al. 1907, p. 94). Thin sections were cut in three perpendicular orientations. The preparation at right angles to the direction of stretching abounds in sections [through individual hornblende grains] which show the characteristic cleavages meeting at angles of 124°, whereas such sections are rare in the other two preparations.

The hornblendes thus have a crystallographic preferred orientation with their c-axes (which are parallel to the cleavage intersections) aligned parallel to the stretching direction.

In terms of the large-scale distribution of minerals, in the district between Laxford and Cape Wrath, pyroxene-gneiss with blue quartz is absent (Peach et al. 1907, p. 37). In the district stretching from Gruinard Bay to Loch Torridon and Raasay, there is an ‘absence or comparative absence of pyroxene-gneiss with blue quartz’. Granulite facies assemblages, then, are only well preserved in the Central District. Together with the previous section we see that the memoir documents a Central District with granulite facies and comparatively little post-dyke deformation, with Northern and Southern Districts characterized by more intense post-dyke deformation and amphibolite facies assemblages.

Summary of memoir view of Lewisian evolution

The Fundamental Complex was made dominantly of igneous rocks, with much banding related perhaps to magmatic flow or solid state deformation. Some patches of basic rocks were intruded by the more acid varieties which form the bulk of the Lewisian; there were also some garnet mica schists interpreted as altered sediments.

After much deformation, a suite of basic dykes was intruded into the Fundamental Complex, cross-cutting earlier fabrics. Ultrabasic dykes are less common than basic ones and crosscut them.

Later, the dykes and their surroundings were deformed, but much less so in the Central District than in those to the north and south. Deformation involved the ‘drawing out’ of patches of rock, or ‘stretching’ to produce linear structures. Bending, thinning and foliation of dykes were all correlated, as was the metamorphism of the original igneous minerals to hornblende-plagioclase mineralogies. Consequently, foliated basic sheets with entirely that mineralogy, and parallel to foliation in the country rocks, were interpreted as original dykes. Such ‘hornblende schist’ sheets were folded, together with layers of metasediment, around large late folds. The Central District was characterized by pyroxene gneisses and these were transformed into hornblende bearing rocks, much as the dykes were. In this metamorphism biotite was commonly formed, and the blueish quartz often found in the pyroxene gneisses lost its distinctive colour. Ultrabasic rocks were metamorphosed to amphibole-talc assemblages. The Northern and Southern Districts have pervasive post-dyke deformation, and also generally lacked pyroxene and blue quartz in the country rocks.

This broad picture remains correct today.

The Lewisian Complex a century after the memoir

In this section we select just a few papers which mark the evolution of ideas in the Lewisian. As well as the themes we highlight from the memoir, entirely new approaches became available via geochemistry, geophysics and geochronology, and research expanded to cover all parts of the Lewisian including the Outer Hebrides and isolated inliers. We cover the different aspects of geology under separate headings, and then synthesize Lewisian evolution in a subsequent section.

Dykes as time markers: reworking and large scale structure

Sutton & Watson (1951) formalized the structural history of the Lewisian with the aid of new nomenclature. On the assumption that the dykes were of a single age they distinguished Scourian (pre-dyke) and Laxfordian (post-dyke) metamorphism and deformation, with a Central District relatively unaffected by Laxfordian reworking (Fig. 1). In that paper, the word ‘Scourian’ is defined as a metamorphic episode but is also used to describe protoliths which are older than the Scourie dykes: thus the term covers pre-dyke events and protoliths (Fig. 4). Dykes were proposed to have a single age or, more precisely, there were no deformation events during the period of dyke emplacement. In contrast, Bowes (1969) suggested four stages of dyke emplacement separated by deformation. Neither idea can be proved conclusively without absolute ages for deformation and intrusion. The first hypothesis has been generally adopted, yet in a modified form, recognizing that there was some ductile deformation during dyke emplacement (Tarney 1973). In the 1960s there was much development of structural ideas as the relationship of foliations and lineations to strain became clearer (Flinn 1965) and multiple stages of deformation were recognized.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Regional cross sections Modified from previous work (Coward 1984; Coward et al. 1970; Coward & Park 1987; Graham 1980; Park 2005), these are perpendicular to the general Laxfordian lineation, so any shear movements related to that lineation will be into or out from the page. The ‘faint’ colours and lines indicate extrapolation above present surface. (a) Outer Hebrides. The possibility that the South Harris Igneous Complex is in a synform is indicated, as is the hypothesis that the Scourian granulites of South Uist were once above the amphibolite facies rocks; (b) Mainland. The geometry as shown is well established, except for the northernmost continuation of the Loch Maree Group. This is shown arching over an antiform and passing beneath the Scourian block of the Central District as suggested by Park (2005); (c) and (d) Speculative sketches showing geometry along lines (a) and (b) after removing effects of late Laxfordian folding. The speculative Laxfordian D2 suture is shown, which may or may not have had metasediments and metaigneous rocks caught up along it. The sites of some ‘future’ D3 structures are shown – see text for details.

It is hazardous to correlate structures on the basis of style and orientation but we still find it useful for general description to use a sequence of D numbers for the Laxfordian deformation episodes as first described by Coward (1973) for the southern Outer Hebrides and subsequently by Coward & Park (1987) for the whole Lewisian. In this scheme, D2 is the main fabric-forming event, producing fabrics in the dykes and transposing them parallel to foliations in the country rock, although D2 does not appear everywhere. D3 is the event during which (usually) NW–SE-trending folds and shear-zones were formed that deformed the D2 fabrics, and D4 is a localized deformation producing minor structures associated with retrogressive metamorphism.

The Outer Hebrides were recognized as being dominated by Laxfordian deformation (Coward et al. 1970), with amphibolite sheets parallel to gneissic foliations almost everywhere (D2), and those foliations were affected by later stages of D3 folding around NW–SE axes (Fig. 4a) (Coward 1973). The effects of superimposed deformations were invoked to explain complicated structural patterns, using the (then) new idea that rock fabrics were the finite results of superimposed strains. For example, dominantly linear fabrics could be explained as the result of two superimposed plane strain deformations (Coward 1973).

Similar structures exist on the mainland in the Northern and Southern Districts (Fig. 4b); in particular NW–SE folds affect an earlier fabric and will be referred to as D3. In our account we make the simplest (but unproven) assumption that D3 folds are of roughly similar age throughout the Lewisian. On the mainland it was recognized that, where they are best preserved, the Scourian rocks themselves had a complex history prior to dyke intrusion. In particular amphibolite facies fabrics can be shown to be pre-dyke in parts of the Central District, despite similarities to Laxfordian tectonites in grade and NW–SE orientation. This retrogressive event, which must have involved a large input of water (of unknown origin), is defined as the ‘Inverian’ (Park 1964; Evans 1965). The older granulite-facies metamorphism and deformation in the Scourian of the Central District was named as ‘Badcallian’ (Park 1970). Field identification of Inverian structure is difficult, especially in the Southern District where the Lewisian is entirely amphibolite facies. The earlier Scourian though, can be identified in the field in terms of fabrics which are demonstrably early and do not trend NW–SE (Fig. 5a, b). Systematic mapping identified a substantial component of deformation as Inverian, based on, for example, subtle low angle discordances between Scourie dykes and wall rock fabrics (Park & Cresswell 1973; Fig. 5c). The dyke intrusion was argued to be controlled by these pre-existing NW–SE fabrics, not just in terms of orientation but also in terms of more abundant and thinner dykes in zones of high Inverian strain (Park & Cresswell 1973). The Inverian is now widely accepted as an important stage of Lewisian deformation (Coward & Park 1987; Wheeler 2007). In zones lacking Inverian fabrics, dykes are subvertical and the simplest explanation is that those zones have not rotated around horizontal axes since dyke emplacement (this assumption is used in the discussion of Laxfordian events, below).

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

Photographs illustrating key aspects of the Lewisian Complex: (a) Migmatitic early Scourian gneiss, south shore of Loch Torridon; (b) Early Scourian banding cut by thin undeformed Scourie dyke (a larger dyke forms the top of the hillock). North side of Loch Torridon; (c) Scourie dyke cross-cutting systematic foliation in gneisses at a low angle. The dyke is deformed but the angular relationship suggests that the country rock fabric is pre-dyke (Inverian). North side of Loch Torridon; (d) View NW of thick Scourie dyke on right (north) side of headland is parallel to (likely Inverian) foliation in country rock on left. North of Scourie, approaching the highest strain part of the Laxford Shear Zone; (e) Shear zone (Laxfordian) in amphibolitized Scourie dyke. NW coast of North Uist; (f) Scourie dyke, pervasively amphibolitized, transposed into parallelism with foliation in country rocks and folded during late Laxfordian deformation. NW coast of North Uist.

In these works, the various deformations were not attributed to an overall tectonic driving force or regime – but there would soon be renewed impetus for ideas to evolve. First, plate tectonics arrived. Second, Ramsay & Graham (1970) pointed out the importance of shear zones as a style of deformation (Fig. 5e). The Lewisian provided examples of outcrop scale shear zones (from a heterogeneously deformed Scourie dyke in North Uist) to fuel the discussion. Simple shear quickly became established as a model for large scale, as well as, outcrop scale structure, partly because it readily explains heterogeneous strain. One clear example of a strain gradient is the northern boundary of the Central District on the mainland, the Laxford Shear Zone (Fig. 1). This was interpreted in a model involving simple shear, with Inverian south side up reverse sense shear explaining the juxtaposition of the Central District granulites with the Northern District which appears never to have been affected by granulite facies (Beach et al. 1974). Later Laxfordian discrete sinistral normal sense shear zones formed just to the south. The curved shapes of these shears suggest they may be rotated listric structures (Wynn 1995). The evolution of the Laxford Shear Zone is discussed in detail elsewhere in this volume (Goodenough et al. 2010). Simple shear with dextral sense was also invoked to explain the kilometre-scale pattern of foliations in South Harris (Graham & Coward 1973), and is a major component of deformation in the Gairloch (Odling 1984) and Canisp (Attfield 1987) Shear Zones.

Laxfordian strain gradients on the kilometre scale are seen in the Southern District, for example, in the Gruinard area the strain gradually increases southwards (Figs 1 & 4b) in a manner broadly analogous to the Laxford Shear Zone. There is more complexity here though because of the folded metasediments of the Loch Maree Group and also because of the low-strain block, several kilometres across, in the Torridon area (the Ruadh Mheallan block, Fig. 4b). Passing south from this block, Laxfordian strain again increases but with a quite different style. Instead of a rather smooth increase in strain as the transition region is traversed, an anastomosing set of shear zones separating low-strain lenses is present (Wheeler et al. 1987; Wheeler 2007). Passing SW the abundance of low-strain lenses decreases, until the deformation is uniformly high strain and resembles that of the Northern District and the Outer Hebrides.

Simple shear on a still larger scale is a possible explanation for those high-strain fabrics which are folded but have flat enveloping surfaces in Laxfordian areas such as the Outer Hebrides (Coward 1975); such an idea was developed simultaneously in other Precambrian regions (Bridgwater et al. 1973). Coward alluded to the possibility that the lineation indicated overall transport direction, and that this could be related to plate movements: though he was cautious about both, this remains a ‘modern’ interpretation of this area (Park 2005) and is a widely applicable tectonic model (Shackleton & Ries 1984). The hinges of the D3 folds of the mainland and Outer Hebrides (Fig. 4) are broadly parallel to the NW–SE lineation, but they were formed in a later event that may relate to a separate collision, much younger than D2, described in the synthesis section later.

The Outer Hebrides exhibit, running along their eastern sides, a complicated zone of greenschist facies ductile fabrics, brittle fault rocks and pseudotachylites: the Outer Hebrides Fault Zone. These SE dipping fabrics and faults crosscut Laxfordian amphibolite facies structures and are later. They show a spectrum of plastically and brittle deformed rocks with somewhat similar kinematics, leading Sibson (1977) to propose that they are snapshots of a kinematically linked brittle fault – plastic shear zone system. This model underlines the role of shear zones as the plastic equivalent of faults on crustal scale, an idea widely applied since. Although the Outer Hebrides Fault Zone is not Laxfordian, the idea of linked brittle-ductile shear zone systems is as likely to have applied to the Laxfordian as to younger belts where it can be more easily demonstrated – there are no known remnants of major Laxfordian faults to which the shear zones coupled, although there are brittle structures which might, in principle, be Laxfordian. By the 1980s knowledge of thrust tectonics had evolved to include the geometries and consequences of ramp-flat geometries and fault-related folds (Boyer & Elliott 1982). These concepts, together with the idea that ductile shear zones are kinematically equivalent to faults, led to a vision of lower crustal deformation in which linked shear zones could form flats and ramps and could provide detachments for folds above them: a vision elegantly summarized by Coward (1984).

Current understanding of the largest scale Laxfordian features then is as follows. In the southern Outer Hebrides there is no doubt that later folds have affected what may well have been a subhorizontal shear zone, which was many kilometres thick since its upper and lower margins are not revealed in the cores of later antiforms or synforms (Fig. 4a). These D3 folds have distinctive form, with open antiforms and tight synforms and will have detached on a shear zone at depth. There are only small low-strain zones. On the mainland, the geometry of the Loch Maree Group leaves no doubt that there has been folding of a large shear zone. Here, though, the orientation of dykes in low strain zones can, if they are assumed vertical prior to the Laxfordian, provide a guide to possible ramp geometries. Thus, on Figure 4d the early Laxfordian fabrics are shown as ramping down beneath the low-strain zone near Gruinard. A band of metasediments, too thin to illustrate, occurs in this ramp region, and is possibly correlated with the Loch Maree Group. Ramps are also suggested beneath the Ruadh Mheallan low-strain zone at Torridon and beneath the Central District in the Laxford Shear Zone.

Concepts of crustal-scale shear zones led to a reassessment of the Inverian structural evolution, which was interpreted in terms of dextral transpression on discrete steep shear zones such as Canisp as well as on subhorizontal structures (Coward & Park 1987). The Inverian foliations and lineations are often coplanar with the later Laxfordian fabrics and, in places, collinear. One explanation is that intense Laxfordian strain will inevitably rotate earlier structures into parallelism. In addition the Scourie dykes appear to have been, at some times, relatively weak during the Laxfordian. At Torridon, Inverian shear zones anastomose and, to a first approximation, they share a common lineation. Scourie dyke intrusion was controlled by these shear zones. If slip is localized on these dykes, then, movement is ‘guided’ by dyke intersections and is thus parallel to the Inverian lineation (Wheeler et al. 1987).

The Outer Hebrides Fault Zone was itself thought to have a thrust sense of movement, partly because it brings granulite facies rocks above amphibolite facies in South Uist, but there is much evidence for normal movement in addition (White & Glasser 1987). Consequently, it appears that the same zone has accommodated different movements and is a persistent weak zone: several episodes of reactivation have been proposed (Butler et al. 1995).

Nomenclature

In this contribution we use the term ‘Scourian’ for any pre-Scourie dyke protolith or metamorphism (with or without deformation), and the term ‘Laxfordian’ for any post-Scourie dyke protolith or metamorphism (with or without deformation), ‘Badcallian’ means Scourian granulite facies metamorphism (with or without deformation) and ‘Inverian’ means Scourian amphibolite facies metamorphism (with or without deformation) which, in the Central District, is seen to postdate the Badcallian. In the Southern District, most fabrics are amphibolite facies but have a variety of ages and orientations. Here, the latest Scourian fabrics run NW–SE and are referred to as Inverian because they are relatively late in the Scourian history and are parallel to Inverian structures in the Central District. Our terms are based on field relationships. Kinny et al. (2005) propose a large number of new names for Lewisian events, based on zircon dating coupled to a proposal that the Lewisian is divided into numerous terranes (an idea discussed later). Currently we cannot relate this nomenclature scheme to field relationships or to the Lewisian literature that we review here, so we adhere to the traditional terms.

Protoliths and geochemistry

Geochemistry provides a precise basis for diagnosing the origins and/or metasomatic history of protoliths. The bulk of the Lewisian was formed in the late Archaean; we split this section into three according to the ages of protoliths as established in the field, though for one or two we deploy geochronological data which are discussed in detail in a later section.

Scourian protoliths: geochemical variability between the North, Central and Southern Districts

The memoir was correct in the identification of the main rock types: the dominant quartzofeldspathic gneisses are meta-igneous and Rollinson & Windley (1980) were the first to identify them as of the tonalite – trondhjemite – granodiorite (TTG) association, typical of Archaean cratons worldwide which have grown by the addition of silicic TTG magmas. In detail the igneous lithologies of the Lewisian complex form a bimodal association, in which older mafic rocks comprising amphibolites and layered mafic ultramafic complexes are enclosed in the more abundant TTG gneisses.

Major and trace element compositions of the TTG gneisses in the North, Central and Southern Districts show spatial variability, mainly in the proportions of tonalite to trondhjemite in the different parts of the Lewisian (Rollinson 1996). So for example, tonalites are most common in the Central District but form a much smaller proportion of the gneisses in the north and south, where granodiorite is more typical. Mantle-normalized trace element plots show important differences in the concentrations of the large ion lithophile (LIL) elements K, Rb, U and Th, but great similarities in immobile element concentrations. The concentrations of the LIL elements in the Central District are much lower than in the Northern District where K-feldspar is commonly observed, an obvious indication that LIL abundances are higher (Sheraton et al. 1973; Weaver & Tarney 1981). It is often proposed that LIL elements are ‘fluid mobile’. Consequently a long-standing interpretation of this geochemical difference is to correlate it with the granulite facies metamorphism of the Central District and infer that the TTG crust was depleted in fluid mobile LIL elements during granulite facies metamorphism (Moorbath et al. 1969) through partial melting and the loss of a silicate melt. However, loss of melt should give rise to a positive Eu anomaly in the restite and a negative Eu anomaly in the more silicic melts whereas in the Scourian the most silicic melts have a positive Eu anomaly (Tarney & Weaver 1987a). More recently Rollinson & Tarney (2005) proposed that the low concentrations of LIL elements in the TTG gneisses of the Central District are primary and reflects a ‘depletion’ process which took place in the TTG magma source, rather than being the product of later granulite facies metamorphism.

Scourian crust generation

The felsic rocks of the TTG association are the product of the partial melting of a basaltic precursor (Rollinson 2006, 2007) and cannot be a direct melt of the mantle. Thus the presence of old mafic rocks in the TTG gneisses is important, because they may represent TTG source rock. This hypothesis has been tested with the basaltic amphibolites found enclosed in TTG gneisses at Gruinard Bay in the southern part of the Lewisian. These basaltic amphibolites are geochemically variable: some show light-rare earth elements (REE) depleted REE patterns, similar to modern N-type mid-ocean ridge basalt (MORB), whereas others show a light-REE enriched pattern (Rollinson 1987). Modelling of the origin of the Lewisian tonalitic and trondhjemitic gneisses at Gruinard Bay through the partial melting of a light-REE enriched basalt shows that their trace element patterns can be reproduced if the melting residue contains a small amount of garnet and plagioclase and larger amounts of clinopyroxene and hornblende (Rollinson & Fowler 1987; Rollinson 1996). Tonalites represent melting at about 18 kb and 800 °C and the trondhjemites melting at slightly higher pressures and temperatures (>22 kb, >800 °C). These results show that the felsic crust was formed through partial melting of a hydrated mafic source at a depth of 60–70 km. In terms of tectonic setting such a deep source is compatible with (though does not prove) melting in the basaltic portion of a subducting slab, so that the source of TTGs was former oceanic crust and they are analogous to modern adakites (Rollinson & Fowler 1987). A point not emphasized in Rollinson & Fowler (1987) is that the basaltic lithology at Gruinard Bay which looks most like modern MORB is the light-REE depleted amphibolites, whilst a light-REE enriched basaltic composition is required to source the observed TTG. Such enrichments are found today in ‘enriched-MORB’, but may also occur in other tectonic settings, not necessarily MORB. Regardless of the explanation for REE enrichment, LIL depletion could relate to dehydration (not melting) in the downgoing Archaean oceanic slab, prior to melting (Rollinson & Tarney 2005). Fluid released at this stage could carry LIL elements with it as it escapes. This model has something in common with the original explanation, because in both models fluids lost during metamorphism carry LIL elements with them. However, in the new model metamorphism and dehydration occur in the source rock prior to melting, rather than in the product TTG after solidification.

Variations in concentrations of LIL elements in TTG across the Lewisian could, extending this argument, reflect chemical differences in the basaltic precursor to the TTG magma. For example in the Central District the low levels of K, Rb, U and Th would indicate that the basaltic parent was dehydrated prior to melting and that those elements were removed in fluid. In the north and south however, the basaltic parent was not dehydrated prior to melting, and the LIL elements contributed to the melt. This difference could, in turn, indicate a difference in the dip of the subducting slab, such that a slab in which dehydration takes place prior to melting dips more steeply than a slab which experiences hydrous melting.

The geochemistry of the Northern District is sufficiently different from the Central that it cannot have been derived simply by hydrating ‘Central District’ material. The two types of TTG could have occupied different crustal levels, or comprised large but distinct batholithic complexes side-by-side within one piece of crust. Then subsequent vertical and/or horizontal movements on the Laxford Shear Zone – which is clearly a zone of high strain – might juxtapose such geochemically distinct protoliths at the present outcrop. Alternatively, geochemical and age differences (see below) have been together used to support much larger movements, with the Laxford Shear Zone being a terrane boundary (Friend & Kinny 2001; Goodenough et al. 2010).

The South Harris Igneous Complex, now entirely metamorphic, comprises original anorthosites, gabbros, norites and diorites (Fettes et al. 1992). These igneous rocks, once thought to be comagmatic, are now known to be very different in age. The anorthosite is Scourian (Palaeoproterozoic) but the other bodies are younger (see below).

Scourie dykes

Scourie dykes comprise quartz tholeiites (dominant), bronzite picrites, olivine gabbros and norites. They were largely derived by melting of enriched subcontinental lithospheric mantle (SCLM) (Tarney & Weaver 1987b) and represent NE–SW crustal extension, albeit over an extended time period (see below). In places fresh dykes cut amphibolitized ones (Tarney 1963) as if the water influx of the Inverian were continuing whilst dykes were intruded.

Most dykes have an enriched trace element signature, which could indicate contamination from the adjacent felsic Scourian gneisses. However, Tarney & Weaver (1987b) show that this is inconsistent with the measured trace element abundances and propose that the enrichment is inherited from a subcontinental mantle source. Thus the emplacement of the Scourie dykes reflects a process of the growth and evolution of the SCLM beneath the Archaean craton. Waters et al. (1990) reported Nd-isotope evidence to suggest that this lithospheric component may be as old as 3 Ga, indicating that there was an enriched subcontinental mantle, isolated from the convecting mantle more than 1 Ga before the time of dyke emplacement. This is consistent with observations from other Archaean cratons that the SCLM formed at broadly the same time as the felsic rocks of the craton and implies a coupled process.

Post-Scourie dyke protoliths

The gabbros, norites and diorites of the South Harris Igneous Complex formed much later than the Scourian anorthosite despite its proximity. Geochemistry shows a calc-alkaline affinity for the major bodies, and an olivine tholeiite affinity for some layered gabbros and ultrabasics (Fettes et al. 1992). This, together with parallels with the geology of the Kohistan Arc in the Himalayas (Coward et al. 1982), suggests that the South Harris Igneous Complex may have been an island arc and was tectonically incorporated in the surrounding gneisses – its boundaries are highly sheared.

Metasediments and metavolcanics occur at Gairloch and north of Loch Maree nearby (Park 1964). Together called the Loch Maree Group, they consist of two contrasting assemblages: c. 2000 Ma-old clastic sediments considered to have formed in a trench or back-arc setting, and a shallow-marine sequence including limestones, banded-iron-formation and graphite schists associated with oceanic plateau-type volcanic rocks. This assemblage is interpreted as an accretionary complex formed at a subduction zone (Park et al. 2001) and is intercalated between slabs of Archaean basement indicating that the complex has been involved in collision with continental crust. Other metasediments form the Langavat and Leverburgh belts of South Harris (Fettes et al. 1992).

The Ard gneiss, a metagranite associated with the Loch Maree Group is post-Scourie dyke (Table 1). Younger Laxfordian granites are present in significant volumes (420 km2) in the Uig Hills – Harris Granite Complex (Myers 1971) immediately north of the South Harris Igneous Complex (Fig. 1) and in sheets in the Laxford Shear Zone.

View this table:
  • View inline
  • View popup
Table 1.

Some key dates from the Lewisian. Horizontal lines indicate 100 Ma intervals, to highlight the abundance of dates in each interval; alternate 100 Ma intervals are shown in black on the left of the table

Metamorphism

There are three developments we highlight here. First, the advancement of metamorphic techniques allowing precise characterization of conditions; second the link between metamorphism and large-scale tectonics and third the recognition of complicated feedbacks between deformation and metamorphism.

Scourian metamorphic conditions

Sutton & Watson (1951) discussed the metamorphism in terms of amphibolite facies and a charnockitic facies: the latter would now be referred to as granulite facies. They recognized that the addition of water to the granulite facies assemblages was required during Laxfordian overprinting. Experimental petrology developments in the 1960s enabled quantitative P–T (pressure–temperature) estimates for Scourian granulites of 15 kb, ±3 kb and 1150±100 °C (O'Hara 1977; O'Hara & Yarwood 1978), implying a high Archaean geothermal gradient. Estimates vary: for example 7–8 kb and 750–800 °C is representative for the mainland Scourian as a whole (Sills & Rollinson 1987) but metasediments near Stoer yield 920–980 °C and >11 kb (Cartwright & Barnicoat 1987). A temperature of 1000 °C (Cartwright & Barnicoat 1989) means that the metamorphism is classified as ultra high temperature (UHT). The meaning of these estimates is difficult to determine for some are based upon former igneous rocks (Rollinson 1982), and so might reflect magmatic temperatures, whereas others are from undoubted metasediments (Cartwright & Barnicoat 1989), and so indicate the deep burial of rocks which originally formed at the Earth's surface. Such UHT metamorphism was more common in the Precambrian and may in general record closure and thickening of backarcs with their characteristic high heat flow (Brown 2007), though there is no direct evidence for that model here. High metamorphic temperatures are consistent with a crust which is built through a process of magmatic accretion but timing information (see below) does not necessarily support this.

Evidence of partial melting in the Central District granulite facies rocks is sparse, which is problematic given the very high temperature estimates which surely make melting likely in many lithologies. Felsic sills in the Badcall and Scourie areas, originally thought to be partial melts of the local felsic gneisses are instead likely to be from much deeper (Rollinson 1994), though there is debate (Cartwright & Rollinson 1995). The only rock types where partial melting is unequivocal are the rather rare metasediments such as those found at Stoer (Cartwright & Barnicoat 1987).

No P–T estimates are available for the amphibolites facies Scourian of the Southern District. One ultrabasic body at Torridon contains assemblages which are granulite facies but may be relict igneous (Cresswell & Park 1973). The basic gneisses above the Outer Hebrides Fault Zone in South Uist and Barra are Scourian granulite facies (Whitehouse 1993).

The Inverian amphibolite facies yields few assemblages which give good P–T data but T is near 600 °C (Sills & Rollinson 1987) or 500–625 °C at 3–6 kb (Cartwright et al. 1985). There is no clear evidence for progradation, although there are syntectonic migmatites reported from the Southern District (Cresswell & Park 1973).

Scourie dyke metamorphic conditions

By definition any Scourie dyke metamorphism is Laxfordian – nevertheless we distinguish very early recrystallization here, which may have preceded any Laxfordian tectonics. The Scourie dykes where undeformed may show margins of hornblende and plagioclase±garnet which O'Hara (1961) suggests formed by ‘autometamorphism’ when the rocks were still hot (300–500 °C). Therefore we are presently seeing a deep level of dyke intrusion. This is confirmed by observations of dykes deformed and metamorphosed at amphibolites facies being cut by undeformed ones (Tarney 1973) implying that, unless we postulate cyclic unroofing and burial, the country rocks were at amphibolites facies. Though the country rocks were hot, they were still much cooler than the magma and chilled margins formed in many dykes.

In the Outer Hebrides, dyke centres commonly show recrystallized textures with two pyroxenes, plagioclase and sometimes garnet, which on mineralogical grounds is granulite facies and is widespread, see Fettes et al. (1992, fig. 12). This is problematic given the general absence of any evidence for granulite facies in the country rocks. However, their temperature estimates of 620–640 °C using coexisting garnet and clinopyroxene (on an assumption of 4–7 kb pressure) from metadolerites are more compatible with the upper amphibolite facies (they also quote unpublished work giving 700 °C). Since the dykes were originally anhydrous, it may be that the dykes recrystallized under conditions of reduced water activity at pressures and temperatures that would, in the presence of excess water, give assemblages dominated by amphibole.

Laxfordian and later metamorphic conditions

Precise Laxfordian data are available from the Loch Maree Group (Droop et al. 1999): 630± 30 °C and 6.5±1.5 kbar from metapelites; 530±20 °C from other rock types. Most of the Laxfordian is deduced to be amphibolite facies simply from the dominance of hornblende and plagioclase in deformed and recrystallized Scourie dykes (Teall 1885). The P–T estimates from the Outer Hebrides Scourie dykes mentioned above may be earlier than Laxfordian deformation and are not necessarily the same grade as that deformation. Apart from that and the problem of geographical separation, completely recrystallized mainland dykes do not commonly contain garnet and clinopyroxene so caution should be applied in extrapolating those conditions. Though they do not provide quantitative P–T data, Cresswell & Park (1973) describe mineralogical evolution in the Scourie dykes of the Southern District. Hornblende and plagioclase persist once the igneous assemblages are replaced, but after garnet coronas (autometamorphic) are destroyed and a foliation developed, garnet porphyroblasts grow statically. This garnet is then replaced by epidote, plagioclase and hornblende±biotite and ores. Later metamorphism also involved crystallization of K-feldspar which was proposed as related to alkali metasomatism. There are hints here of a retrograde P–T path but there is clearly a need for quantitative chemical data to advance our understanding.

Uniquely for the Laxfordian, South Harris reached granulite facies as recorded in the South Harris Igneous Complex and with specific P–T estimates from the Leverburgh metasediments of 800 °C and 13–14 kb (Baba 1998). A more detailed P–T path shows peak T of 955±45 °C at 10.0±1.5 kb followed by peak P of 12.5± 0.8 kb at 905±25 °C (Hollis et al. 2006). The granulite facies metamorphism was originally thought to be Scourian, because of the grade, but is now known to be Laxfordian, affecting both Scourian and post-Scourian protoliths (see below for details of dates). Remarkably, then, the Lewisian shows two UHT regions of quite different ages.

Inliers within the Moine Supergroup are diagnosed as Lewisian in the Memoir and these show some distinctive assemblages, namely eclogite facies in the Glenelg-Attadale inliers (Teall 1891) with P–T estimates of 20 kb and 750–780 °C (Storey et al. 2005).

Metamorphism and large-scale tectonics

The P–T evolution of the Lewisian seems, at first sight, to indicate gradual unroofing and cooling throughout the Proterozoic (O'Hara 1977), though this picture is biased by the amount of data from the Scourie-Assynt area. O'Hara proposed that the Badcallian granulites were not far above the Moho when they formed, so that many kilometres of crustal material must have been added, by what he called a ‘non-tectonic’ (i.e. presumably igneous) underplating process. However, igneous underplating cannot explain the differential uplift shown by the juxtaposition of rocks with different pressure–temperature–time (P–T–t) histories.

Instead, tectonics and erosion together form the basis of viable interpretations. Overthrusting means exactly the same thing as tectonic underplating. Tectonics allows explanation of how rocks with different P–T–t histories can be juxtaposed. A brief summary in the context of diagnosing extensional structures is given by Wheeler & Butler (1994) – the logic is basically the same for other structures. Even post-metamorphic tectonics is not essential to explain high grade rocks at the Earth's surface. If the Badcallian crust was, for example, 70 km thick, with the present erosion level 30 km above the Moho, it could have been unroofed slowly by erosion without necessarily requiring material to be added beneath by any mechanism (Sutton & Watson 1987). However, once it is appreciated that the Loch Maree Group was deposited long after the Scourian metamorphism and is now at amphibolite facies, a Laxfordian cycle of burial and unroofing must be hypothesized. Moreover, unless the tectonic emplacement of the Loch Maree Group onto its present footwall (Fig. 4) was very late in the Laxfordian, whatever buried the Loch Maree Group would also have buried those footwall rocks. Similar arguments for relative movements can be deployed in South Harris. If it is the case that the TTG gneisses around this area have not experienced granulite facies in the Laxfordian, and the South Harris Igneous Complex clearly has, then there must have been relative movements to juxtapose the two different metamorphic grades. A history of slow, monotonic unroofing for the whole Lewisian is not viable; some form of differential tectonic thickening during the Proterozoic must have occurred.

Relationships between deformation and metamorphism

Teall recognized that the transformation of Scourie dyke to amphibolite was spatially associated with deformation. The relationship – or feedback – between deformation and metamorphism has since been recognized as of huge importance in crustal (and likely mantle) behaviour. ‘Transformation modified deformability’ (Brodie & Rutter 1985) is often manifest as (possibly transient) softening, for example due to the reduction in grain size caused by growth of new minerals. Smaller grain sizes allow easier deformation by diffusion creep. Complementary to this, deformation can promote reaction or enhance reaction rates by, for example, generation of permeability during deformation (allowing fluid ingress and hydration), and by increasing the plastic strain energy stored within grains and hence decreasing their chemical stability. The positive feedbacks of deformation on reaction and of reaction on deformation provide an explanation for the observed association of strain and metamorphic changes.

The amphibolitization of anhydrous Scourie dyke mineralogies obviously involved the addition of water. In fact, the volumes of water involved were larger than this alone would imply. Metasomatic changes recorded in shear zones show that they were fluid conduits (because of deformation-enhanced permeability), not just fluid sinks. Beach & Fyfe (1972) shows that potassium is more abundant in Laxfordian shear zones near Scourie than in the adjacent wall rocks; calcium is depleted (Beach 1973). Fluids were proposed to originate from the Laxfordian granites found to the north. Metasomatism is widespread in the Lewisian, for example at Torridon shearing involved loss of potassium and gain of sodium (Beach 1976); this suggests that, on the scale of the whole Lewisian, deformation may have promoted fluid flow, retrogression and metasomatism.

Geochronology

Field based and petrographic observations can, at best, constrain the relative order of events. Absolute dating is essential to understand the Lewisian, but we need to ask: what exactly does a radiometric age mean in a metamorphic region (Cliff 1985) – the formation of the protolith (igneous, in the case of the quartzofeldspathic gneisses and Scourie dykes), the metamorphism, the deformation, a fluid influx event, or the cooling? Indeed in crust, which has grown through magmatic accretion, can we distinguish completely between igneous and metamorphic events, especially in the case of magmas added directly to the deeper crust? This issue arises when interpreting not just the Badcallian but also the earliest recrystallization of the Scourie dykes, and the South Harris Igneous Complex metamorphism. We need to appreciate that age data are most valuable when tightly linked to field observations and the petrographic setting of individual dated minerals. At the time of writing, geochronology is the most controversial aspect of Lewisian research. In the account we give, note that the early work may have been re-interpreted or superceded: and this may eventually happen to recently obtained data too. We will not review all dates but give a basic outline, with examples selected to highlight current debates in interpretation.

Development of geochronological work

The suspected great age of the Lewisian was confirmed by Rb–Sr and K–Ar ages of 2460 Ma for a Scourian pegmatite and 1600 Ma for Laxfordian granites and pegmatites (Giletti et al. 1961). Soon after, the Inverian episode was defined on the basis of a pre-dyke amphibolite-facies event overprinting granulite facies rocks in the Lochinver area and dated at 2200 Ma by K–Ar and Rb–Sr (Evans 1965). As explained previously, this event has now been widely identified in the Lewisian, though using structural rather than geochronological information. Such dates may not relate simply to particular geological events because of the ease with which Ar can be gained and lost from minerals and because Rb–Sr ages may register cooling rather than crystallization. The U–Pb system is less prone to those problems. The U depletion event (described in the geochemistry section) was dated at 2900±100 Ma (Moorbath et al. 1969), and consequently the magmatic protoliths of the Lewisian had separated from the mantle by then. U depletion also occurred in South Uist, Benbecula and Barra (Moorbath et al. 1975) so was not restricted to those areas lacking Laxfordian overprinting. Because of problems with open-system behaviour during cooling and reheating, attention was paid to minerals with high retention of parent and daughter isotopes: Pidgeon & Bowes (1972) dated growth of zircons from granulite facies rocks using U–Pb as 2660±20 Ma. Sm–Nd is the final isotopic system we mention here; it has the advantage that it can be used on minerals such as garnet which are part of assemblages indicative of grade. The first use of this in the Lewisian gave 2490 Ma for a granulite facies rock (Humphries & Cliff 1982); although Sm–Nd has a quite a high blocking temperature this is interpreted as a cooling age: hence it could be reconciled with older zircon ages. Burton et al. (1994) reported a Sm–Nd mineral isochron age of 3300 Ma for amphibolites at Gruinard, candidate source rocks for the TTG intrusions (see geochemistry section). This date was interpreted as being shortly after differentiation from a depleted mantle source. Being very old, this date stimulated further work which contradicts it. Using whole rock data, which they argue are more robust than mineral isochrons, Whitehouse et al. (1996) obtained a Sm–Nd age of 2943 Ma for an amphibolite suite and 2846 Ma for a hornblendite-metagabbro suite, both from Gruinard. These ages were interpreted as magmatic.

Key protolith ages

The previous section introduced the potential and the challenges of Lewisian geochronology. Table 1 shows some key ages, with brief notes on the authors' interpretations. There are some dates that are relatively straightforward to interpret on a local level (although tectonic interpretations may vary, we discuss these separately). For example oscillatory zoning in zircons is interpreted as igneous and thus a U–Pb date for the oscillatory zoned portion of a zircon, obtained by a variety of techniques and with a variety of caveats (Ireland & Williams 2003; Parrish & Noble 2003), will relate to intrusion. Figure 2 shows the positions of protolith ages from Table 1. Scourian protolith ages for TTG rocks of the Lewisian cover a range from 3135 Ma (Loch Torridon) and 3125 Ma (Isle of Lewis) through to 2680 Ma (Northern District) – a span of 455 Ma, representing the formation of the bulk of the Lewisian that we see today. The present spatial distribution of ages represents some combination of (a) igneous rocks intruding older igneous rocks; and (b) relative movements of pieces of crust.

The Scourie dykes yield two ages: c. 2400 Ma from baddeleyite, and a Rb–Sr whole rock age (Chapman 1979); and c. 2000 Ma from baddeleyite, and from Sm–Nd mineral isochrons (Waters et al. 1990). In general field evidence shows that the ultrabasic dykes are younger than the basic suite, but, oddly, a picrite dyke yields the oldest age known. Whether intrusion continued throughout this 400 Ma period (which is almost as long as the span of Archaean TTG ages) or was episodic is unknown.

The South Harris Igneous Complex has anorthosites intruded at 2491 Ma with diorites, norites and tonalites at 1890–1870 Ma, see also Whitehouse & Bridgewater (2001). Clearly the anorthosites cannot be genetically associated with the other igneous rocks. From the point of view of field relationships, the distinctive lithologies and very high strain contacts (Graham 1980) are quite in keeping with the large displacements required to juxtapose diorites and norites of a ‘young’ calc-alkaline arc (the South Harris Igneous Complex) against ‘old’ continental material. However Mason et al. (2004) argue that the diorites and norites must have intruded continental crust.

On the mainland the Loch Maree Group metasediments (an accretionary complex, see above) contain detrital zircons dated at 2200–2000 Ma so are younger than those ages; the Ard gneiss, metagranodiorite intrusive in the Loch Maree Group, is 1900 Ma. Granite sheets at Loch Laxford are dated at 1885 Ma but a sheet on South Harris is 1675 Ma, markedly younger. These are the youngest significant protolith ages.

Metamorphic ages

Igneous ages may be used to date, or constrain dates for, metamorphism if it is demonstrable in the field that the igneous body is contemporaneous with, or has a definite time relationship to, the metamorphism. For example a trondhjemitic sheet near Scourie dated at 2720 Ma (Corfu et al. 1994) is interpreted as syn-granulite facies (Cartwright 1988).

Of more general applicability is the idea that if a mineral shows a growth stage which is interpreted to be below its closure temperature and can be linked to a particular metamorphic event (via textures) then that event can be dated. Zircons commonly show distinctive growth textures related to metamorphic events, and have a very high closure temperature. U–Pb concordia diagrams may also show discordant ages which are related to lead loss due to heating during metamorphism. Thus, zircon may be used, with care, to date metamorphism (Parrish 2001) but a key issue is the ambiguity in relating growth and/or lead loss to specific metamorphic events (Parrish & Noble 2003).

Scourian metamorphic ages include recrystallized zircon rims from Gruinard Bay showing c. 2730 Ma, interpreted as granulite facies, and monazites (enclosed in garnets) showing 2760 Ma near Scourie (Table 1). Monazites from less than 1.5 km from the 2760 Ma sample show 2526 Ma; new growth of zircon in a quite fresh granulite is dated at 2490 Ma; and zircon from a pyroxenite grew at 2470±30 Ma (Kinny & Friend 1997). The latter age is referred to as ‘Inverian’ but in its original usage this was intended to refer to an amphibolite facies overprint. The last two dates are from quite fresh granulites and so it is argued must relate to a granulite facies event. However it is certainly the case that zircon can grow at amphibolite facies (Parrish & Noble 2003). The granulite facies region as a whole exhibits much heterogeneous amphibolite facies overprinting, that heterogeneity being determined by the supply of water not by variations in temperature. How sure can one be that zircon growth was not triggered at amphibolite facies temperatures (the rock as a whole retaining its granulite facies assemblage due to lack of water)? Note also that the c. 2500 Ma ages are not registered in all rock types, even those which are so close together that they must have shared the same P–T history (an example being the trondhjemitic sheet from Badcall dated at 2720 Ma). Thus, in terms of minerals dating metamorphic events, ‘absence of evidence is not evidence for absence’.

Laxfordian metamorphic ages cover a considerable span. Cooling after granulite facies metamorphism in the South Harris Igneous Complex anorthosites was dated by Sm–Nd as 1870± 40 Ma; this agrees with zircon metamorphic overgrowths of 1880 Ma (Friend & Kinny 2001). Because the metamorphic ages are close to the intrusive ages here, it is plausible that the UHT conditions recorded represent residual magmatic heat (Baba 2004). The granite/migmatite complex is much later (1675 Ma).

The exclusively amphibolite facies Laxfordian on the mainland appears much younger than the granulite facies in the South Harris Igneous Complex, despite the existence of an 1855 Ma granite north of the Laxford Shear Zone. Monazites are interpreted to have grown at amphibolite facies at 1752 Ma. Titanites grew at 1760 Ma and 1670 Ma in the Scourie–Laxford area (U–Pb) and were linked to hydrothermal activity related to granites (Corfu et al. 1994) although no comparable dates are known from nearby granites. Rutile from a Scourie dyke shows a formation or reset U–Pb age of 1689 Ma. Note that the closure temperatures of these minerals are not well known so there is a chance that these are closure not crystallization ages. Pegmatites in the Gairloch area synchronous with Laxfordian D3 deformation are 1700 Ma (Park et al. 2001): this is the only Laxfordian date that can be directly linked to structures.

Disregarding low closure temperature cooling ages, these are the youngest metamorphic events in the main part of the Lewisian, though pseudotachylites are common in the Lewisian & Sherlock et al. (2008) find Grenville 40Ar/39Ar ages from the Gairloch region (1024–980 Ma). These are actually younger than the oldest age for unmetamorphosed Torridonian sediments, 1200 Ma (Kinnaird et al. 2007). The eclogites of the eastern part of the Glenelg-Attadale inlier are Grenville in age (Sanders et al. 1984; Storey et al. 2005). Remarkably, in view of their proximity and similar grade, the eclogites in the western part of the Glenelg-Attadale inlier are much older, c. 1700 Ma (Storey 2002).

We noted in the metamorphic discussion that the Lewisian has been described as undergoing slow cooling through its history. However the Loch Maree Group metasediments undoubtedly show a burial cycle in the Laxfordian – the adjacent gneisses may have experienced this too. Moreover, the younger Scourian metamorphic ages described in this section have been attributed to a discrete heating event (Corfu et al. 1994). We cannot yet be sure whether the gaps between recorded events relate to a smooth or cyclic P–T evolution. Figure 6 relates metamorphic conditions to dates for three parts of the Lewisian.

Fig. 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 6.

Metamorphic P–T histories and dates in Ga for three parts of the Lewisian. It is important to understand that these paths are based on just a few points, and potentially have as yet unrecognized P–T excursions and reburial episodes. Scourie path from Cartwright & Barnicoat (1989) except Scourie dyke ages span 2.0–2.4 Ga and exhumation to the surface was complete by 1.2 Ga; other dates are not adjusted from the published figure in view of current uncertainty about possible multiple granulite facies events in the Scourian. South Harris path from Baba (1998) with the metamorphic age marked by us (speculatively) as relating to peak pressure. Loch Maree Group shown as a single PT estimate (Droop et al. 1999) with a speculative age, plus an indication that those sediments were deposited later than 2.2 Ga, to emphasize the difference to Scourie.

Alternative tectonometamorphic interpretations

The great variety of protolith ages together with the heterogeneous pattern of metamorphic ages realized in recent years has led to the proposal that several disparate tectonic units, ‘terranes’ (Coney et al. 1980), have been assembled to form the Lewisian, the terranes being ‘diverse packages of lithologically and geochronologically distinctive rocks’ (Kinny et al. 2005). One argument used is that different ranges of protolith age must represent different regions, though one must consider the possibility that different generations of igneous rock have intruded one region (Park 2005). For example one terrane boundary runs through the Assynt area and separates the 2960–3030 Ma ages of the ‘Assynt Terrane’ from 2840 Ma protolith ages of the ‘Gruinard Terrane’. To the field geologist these two terranes look rather similar. Both have quartzofeldspathic gneisses, with granulite facies metamorphism (less prevalent in the Gruinard area). A particular lineament, the ‘Strathan Line’ (Kinny et al. 2005) is proposed as the terrane boundary but there are no indications in the field that this is a major shear zone. In contrast shear zones within a single terrane (with, presumably, less displacement) are spectacular in the field for example Canisp. An alternative explanation for age differences is simply that there was more than one age of intrusion into a single piece of crust during the Archaean: the implied contacts between different suites of TTG rocks now being transposed or obscured by later deformation.

Now consider the metamorphic ages: the 2490 Ma event in the Assynt terrane (described in the previous subsection) contrasts with the 2730 Ma metamorphic age in the Gruinard Terrane – the absence of a 2490 Ma record implies a large separation between the terranes at this time (Kinny et al. 2005). Regional metamorphism, by definition, is governed by thermal evolution at crust/orogen scale, and one expects variations in P–T history only over large distances, unless there are major tectonic contacts. However, the previous section gave examples of rocks near to each other, with no possibility of a terrane boundary between them, which register different events. The trondhjemitic sheet from Badcall dated at 2720 Ma via zircon (see above) does not register the 2490 Ma event seen in adjacent rocks (Table 1); the monazites from near Scourie dated at 2760 Ma do not register the 2526 Ma event recorded in nearby monazites. Consequently the apparent absence of a metamorphic event cannot unequivocally be used to deduce that that event did not happen.

Alternative models with fewer disparate tectonic units can be proposed (Park 2005). There is no doubt that large relative movements are involved during the evolution of the Lewisian – the intense strain in the Laxfordian is itself evidence for such movements. The question is whether, to explain and accord with geochronological, structural and metamorphic data, really large displacements are required along rather specific boundaries. Such displacements are easiest to justify when very different rock types are involved, such as the Loch Maree Group metasediments above TTG gneisses. The South Harris Igneous Complex has very distinct character, and is described as an accreted terrane in Park's deliberately conservative model; yet even here an alternative model exists in which the diorites and norites are intruded into existing crust rather than accreted to it (Mason et al. 2004). The possibility that the Laxford Shear Zone is a terrane boundary is discussed in detail in this volume (Goodenough et al. 2010).

We conclude by reiterating the fact that the behaviour of minerals used in dating is complicated and not fully understood; this reinforces the need to take field and petrographic data into account when evaluating the significance of dates in the Lewisian and elsewhere.

Geophysics

Insights into crustal structure

Refraction surveys show mid-crustal P-wave (check) velocities of 6.55 m s−1 and lower crust 6.7 m s−1 (Powell & Sinha 1987; Kelly et al. 2007). This suggests that the crust becomes more basic downwards, though obviously these are average values, and seismic anisotropy will influence interpretation (Barruol & Mainprice 1993; Rudnick & Fountain 1995). In one Cenozoic dyke on Lewis, lower crustal xenoliths are brought up (Menzies et al. 1986; Macdonald & Fettes 2007): they comprise mafic garnet-free granulites and mantle peridotites. The former are compatible with the lower crustal velocities. Reflection profiles show a rather featureless upper crust but a lower crust (from 12 km to the Moho at 26–29 km) replete with seismic reflectors (Smythe 1987). The Outer Hebrides Fault Zone can be traced as an east-dipping set of reflectors down into the middle crust. Gravity and magnetic data allow the shallow-depth extrapolation of major lithological boundaries in the Lewisian (Hall 1987). All of these works give insight into the present day geometry of the Lewisian crust: they do not provide firm evidence allowing us to distinguish between models for Lewisian evolution invoking originally thick Scourian crust or igneous underplating through the Proterozoic.

Large scale movements and plate tectonic setting

Palaeomagnetism may in principle give information on crustal movements if contrasting remanence vectors of equivalent ages are found. For example Piper (1992) shows how such data, together with matching of aeromagnetic anomalies, support c. 95 km of (presumed Caledonian) sinistral strike slip in the Minches. Scourie dyke trends provide a comparable estimate (Lisle 1993). Remanence is often acquired during cooling so the directions recorded relate mainly to cooling after the Laxfordian event: they vary regionally but, apart from the above result, variations relate to the complex details of magnetic carrier mineralogy.

On a much larger scale, palaeomagnetism contributes towards reconstructions of the positions of major cratons. A plausible palaeomagnetic reconstruction at c. 1350 Ma (Buchan et al. 2000) places the Proterozoic continent of Baltica against the eastern margin of Laurentia, close to the position of the Lewisian fragment (Fig. 7) but in an orientation c. 90° anticlockwise from its mid-Phanerozoic position. Thus Palaeoproterozoic events and structures in Greenland, Labrador and Baltica have a direct relevance to the Lewisian (Table 2). In particular there are deformed belts formed at c. 1.91–1.84 Ga in Greenland (the western and eastern Nagssugtoqidian) and in Baltica (the Lapland-Kola) as shown on Figure 7, and in Labrador (the Torngat), It has in the past been proposed that these belts were ‘intracratonic’ in that they formed by localized shortening within a contiguous region. It is now recognized that they were collisional orogens hosting subduction-related calc-alkaline plutons (Kalsbeek et al. 1993; Park 1994). During the same period, major subduction-related crustal growth occurred in the Svecofennian belt on the southeastern side of Baltica.

Fig. 7.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 7.

Palaeogeographic evolution of North Atlantic area during the Laxfordian. Arrows indicate convergence directions at each stage. Abbreviations: Lew, Lewisian; NS, north Scotland; NI, north Ireland; for others see Table 2. Colours relate to the age of last main pervasive deformation (Table 2), except for green which indicates material in island arcs etc. which is juvenile in relation to each time illustrated. Dots indicate calc-alkaline igneous activity. Modified after Park (2005) and references therein.

View this table:
  • View inline
  • View popup
Table 2.

Tectonic units in the North Atlantic area (see also Fig. 6) with broad ages of Palaeoproterozoic deformation and magmatism (Park 1995, 2005)

Following on from the period of cratonic amalgamation, a major phase of subduction-related crustal growth took place in both Laurentia and Baltica at the margins of the combined continental mass from c. 1830 Ma to c. 1600 Ma in the Ketilidian & Gothian belts (Fig. 7) and also in the Labradorian and Makkovik belts (Park 1991, 1994; Gower & Krogh 2002).

We should expect the Lewisian to fall into this general framework. The Loch Maree Group and South Harris Igneous Complex then represent igneous activity and sedimentation related to a closing ocean, with igneous intrusion and metamorphism close together at c. 1900 Ma. The ‘main’ Laxfordian deformation, D2, associated with flat-lying fabrics and a NW–SE lineation has not been dated but plausibly that relates to the same oceanic closure. Material both above and below the suture is heterogeneously deformed during the Laxfordian. The Laxford Shear Zone is not likely to have accommodated huge movements during the Laxfordian, so does not represent a Laxfordian suture – it is probably Inverian (Goodenough et al. 2010) – consequently the difference between the Laxfordian of the Central and Northern Districts is due to heterogeneous strain in what was a single body of continental material. Similarly, the strain gradient at Loch Torridon does not, in the field, appear to include a major suture, and the area exemplifies strain heterogeneity on all scales [e.g. fig. 2 of Wheeler (2007)].

The late Laxfordian D3 deformation seems to relate to NE–SW shortening and is substantially younger (c. 1700 Ma). The structures may be explained by an approximately north–south-orientated compressive stress. Park (1994, 2005) relates this tectonothermal event to the formation and subsequent collision of a marginal volcanic arc co-extensive with the contemporary Gothian belt of Baltica (Fig. 6) and the Labrador belt in Laurentia. The Laxfordian D3 event did not form new fabrics on the large scale, hence in Figure 6 the Lewisian is shown as outside the ‘main’ part of the Gothian-Ketilidian belt.

Whatever the precise assignment of subareas to the lower and upper Laxfordian plates, it seems that both experienced the Inverian, both experienced intrusion of a dyke swarm, and where undeformed those dykes in the final assembly have a rather consistent NW–SE trend.

Summary of Lewisian evolution

We offer an outline only, and it should be recalled that ‘the Lewisian’ as we see it today may have been assembled in disparate localities as well as at different times.

Early Scourian

The earliest rocks are metabasites with associated metasediments – some of which could be fragments of Archaean oceanic crust. This crust was metasomatized during subduction, losing U, Th and K. It melted at c. 20 kb, creating voluminous TTG magmas (at 3000–2900 Ma), which themselves had low U, Rb, Th and K and are found in the Central District and southern Outer Hebrides. At other times and/or in other places, crust was subducted with a variety of dip angles, and was not always depleted prior to melting. Similar magmas, though not depleted in those LIL elements, intruded at 3100 Ma in the Outer Hebrides and Southern Districts and at c. 2800 Ma in the Northern District, though all districts have a span of TTG ages. Associated with this early crust-forming event was the formation of the subcontinental lithospheric mantle. Structural arguments that the crust deformed by subhorizontal shear in the Laxfordian, and general analogies with Phanerozoic orogens, indicate the possibility of large horizontal movements which can displace crust sideways from the lithospheric mantle it was originally above. Nevertheless it seems that the Scourie dykes (which provide isotopic evidence for the age of their source) formed above lithospheric mantle which has the same age as the earliest Scourian crust.

Deformation forming early (Badcallian) banding and fabrics accompanied the peak metamorphic granulite facies metamorphism, which is substantially later than TTG formation in all parts of the Lewisian where it is seen. In the Gruinard area mafic rocks crystallized at 2800 Ma or earlier but granulite facies metamorphism occurred at 2700 Ma. In the Central District (Scourie region) a 2760 Ma granulite facies event is documented in monazites. There are several metamorphic dates near 2500 Ma from the Central District, which are suggested to relate to a later granulite facies event. In the Torridon area and further south, there is no firm evidence for Scourian granulite facies but it is present in the Outer Hebrides above the Outer Hebrides Fault Zone.

Pressure estimates indicate that the granulite facies rocks we see now at the surface were once at >30 km depth, and there is another 30 km or so of crust beneath them now, so it is possible that there was a 70 km-thick crust in the Badcallian, with the present exposure level in the middle. As these rocks are mostly very low in heat-producing elements, radiogenic heating is unlikely to account for the metamorphism. Magma addition could produce not only thickening but also heat – however it is worth remembering the time gap between the youngest TTG ages and the granulite-facies ages in all Scourian regions. Alternatively, as in Phanerozoic orogens, the high grade rocks found at the surface are the result of thrust-related uplift (which could equally be referred to as tectonic underplating) coupled to unroofing by erosion and/or tectonic extension. The widespread occurrence of granulites in Archaean cratons led O'Hara (1977) to prefer ‘non-tectonic’ (i.e. igneous) underplating. He argued that the granulites were initially not far above the Moho: as a consequence most of the present thickness of the crust would have been added later. We do not dismiss the possibility of magmatic underplating but as discussed in the metamorphic section, there are limits to what it can explain. It cannot produce the juxtaposition of rocks with different P–T–t histories, nor can it explain repeated uplift and reburial; we therefore prefer the tectonic explanation.

Later Scourian

There is limited petrographic or field evidence for two granulite facies events in the Central District. Instead, there is abundant evidence for amphibolite facies overprinting which preceded the Scourie dykes – the ‘Inverian’ event as diagnosed from field criteria. Whether the term should be redefined as a time period, including the postulated second granulite facies event, or restricted to refer to amphibolite facies episode, is a matter of nomenclature.

The major question is how c. 2500 Ma dates in granulite-facies rocks can be reconciled with field and petrographic information that suggest an amphibolite-facies event prior to Scourie Dyke intrusion, which started at 2400 Ma. As defined on field criteria, the Inverian has a systematic structural style: steep NW–SE shear zones formed at amphibolite facies, retrogressing anhydrous rocks if they were the local precursors. Associated is the static retrogression of granulites to amphibolites in the Central District, so large amounts of water were added to the present exposure level at this time. Pressure is substantially less than 10 kb, a result of the first stage of gradual unroofing of the Badcallian complex. The Inverian itself could plausibly have caused the Badcallian granulites to be thrust over lower-grade rocks on a ductile ramp-flat system (Coward 1984): tectonic underplating assisting the uplift and unroofing of the granulites. Inverian shear zones include the Canisp Shear Zone which is within what is usually considered a coherent piece of crust, and should be contrasted with the Laxford Shear Zone, which separates quite different pieces of crust and may relate to docking of those different terranes late in the Scourian (Goodenough et al. 2010).

The Inverian may be as important as the Laxfordian in terms of deformation, but its effects are partly masked by that later stage and so are difficult to unravel. Where fabrics in quartzofeldspathic gneisses are parallel to completely foliated Scourie dykes, as in the Northern District and Outer Hebrides, there are no field criteria allowing the two events to be separated.

Scourie dykes

The Scourie dyke swarm is NW–SE trending and dominated by quartz tholeiites derived by melting of enriched subcontinental lithospheric mantle. It represents NE–SW crustal extension and span 2400–2000 Ma; ultrabasic dykes arrived late in the sequence according to field criteria but yield an early date. Intrusion geometry was commonly controlled by NW–SE Inverian fabrics where those were present: dykes appear to have been thinner, and to anastomose more, in such areas, but roughly the same dyke trends exist regardless of wall rock structures. In areas of control by steep Inverian structures, the dykes are subvertical (i.e. sill-like) but are also subvertical when seen far from areas of Inverian deformation.

Dykes often have a static amphibolite facies overprint, but those may be cut by fresh dykes, as if the Inverian episode of water addition was continuing. Other dykes have autometamorphic garnet or two pyroxene metamorphic assemblages indicating intrusion into hot country rock: overall then we are looking at considerable depths and wall rock temperatures despite the obvious brittle failure involved in intrusion. There is no evidence for any ambient pressure change from the earliest to the latest dykes, so there must be a limit on the amount of erosion that occurred during this 400 Ma time interval.

Early Laxfordian

Just after the youngest known Scourie dyke, the Loch Maree Group metasediments and calc-alkaline rocks of South Harris formed and were incorporated into a Laxfordian orogen. Figure 4c, d show a simple (two plate) model for the early Laxfordian structures. Park (2005) argues that a SE plate was subducted under a NW plate. NW–SE directed ductile overthrusting created crustal-scale shear zones throughout much of the Lewisian.

The Loch Maree Group lies beneath basement rocks that must have been thrust over it during the Laxfordian and the simplest explanation for its metamorphism is burial and conductive/radioactive heating. Thus, the early Laxfordian involved burial and progradation in the style of a Phanerozoic orogeny. It is rather hard to envisage the footwall rocks to the Loch Maree Group being at amphibolite facies when the Loch Maree Group was emplaced into its present position. The Loch Maree Group, an accretionary complex, is likely to have been structurally higher than the middle-lower crust of the Lewisian continental margin over which it was emplaced. Instead, it is more likely that the footwall had been exhumed to high crustal levels before the Laxfordian, and was then tectonically buried, heated and deformed just as was the Loch Maree Group above it. Figure 4c shows most deformation occurring in the footwall to this main suture; deformation occurred at amphibolite facies throughout the levels exposed today on the mainland, and Scourie dykes were transposed parallel to Laxfordian wall rock foliations. A single granite sheet north of the Laxford Shear Zone is known to be of this age.

The main Laxfordian fabrics on the mainland are on a large scale closely related to the Inverian ones. This is due to the transposition of Inverian fabrics during Laxfordian strain, and to the behaviour of dykes as easy-slip horizons leading to a coincidence in lineation directions (Wheeler et al. 1987).

In the Outer Hebrides, the calc-alkaline nature of the South Harris Igneous Complex suggests that it belongs to the upper plate, regardless of whether it was an island arc or was intruded into adjacent rocks. It experienced granulite facies metamorphism at 1880 Ma; there is no evidence for this in the TTG gneisses and, if they really did not experience it, there must be large relative movements between the South Harris Igneous Complex and the TTG gneisses. Figure 4c interprets the South Harris Igneous Complex as part of a larger nappe, thrust over pervasively deformed lower plate rocks [but carrying the South Uist granulites above it (Park 2005)]. The remaining outcrop is interpreted as synformal as a result of D3 (the antiformal closure of the anorthosite body may have been established earlier, and hence is not diagnostic of the overall South Harris Igneous Complex geometry). Coward (1984) linked the overthrusting of a hot granulite facies nappe to the formation of the migmatite complex in its footwall. New dates show that those migmatites are 200 Ma younger and cannot be related in that fashion. However emplacement of such a nappe would undoubtedly have led to temperature rise in its footwall, by downwards heat conduction and by providing thermal insulation. Pursuing that argument, the footwall rocks would have been colder than amphibolite facies prior to the overthrusting, and experienced a prograde path, though there is no direct evidence for this. Moreover, the footwall rocks did not experience granulite facies so the South Harris Igneous Complex nappe must have been partly unroofed prior to emplacement: a common phenomenon in Phanerozoic orogens.

Later Laxfordian

Late Laxfordian NW–SE trending folds and dextral shear zones (D3) that overprint the main Laxfordian fabrics (D2) are associated with syntectonic pegmatites, dated at 1700 Ma, considerably post-dating South Harris Igneous Complex metamorphism (at c. 1850 Ma) suggesting a separate event. The D3 structures involved roughly north–south shortening, in places partitioned into steep dextral shear zones. There is no D3 suture in the main outcrop of the Lewisian. However, the scale of folding shown in D3 (Fig. 4) is quite considerable, and there must have been ductile shear zones at depth on which these folds detached, so significant thickening of the crust is implied. It has been suggested by Park (1994) that these D3 structures may be related to collision with an arc or small terrane further south (Fig. 7). Monazites record a 1750 Ma amphibolite facies event near Scourie and migmatization occurred in the Uig Hills – Harris Granite Complex. The causes of amphibolites facies metamorphism and migmatization remain to be clarified but significant crustal thickening is expected to have had thermal consequences.

On the mainland, NW–SE folds are present in the Northern and Southern Districts. The low strain zone at Torridon presents a problem in interpretation. The Loch Maree Group is clearly folded (Fig. 4b) as is its immediate footwall. The Gairloch geometry suggests an isoclinal fold. If the original base of the Loch Maree Group nappe was subhorizontal then it, and the adjacent rocks to the south, would have been rotated 90°. However the Scourie dykes just to the south are still subvertical. For this reason we propose a D3 shear zone juxtaposing the Ruadh Mheallan block (in its original attitude) next to the Loch Maree Group (tightly folded). This is just one of several options for the structural evolution of this area.

These events established the main Lewisian features in the form we see them today, apart from the removal of another 15–20 km of material to exhume the Laxfordian amphibolite facies rocks, a process complete when the earliest Torridonian rocks were deposited prior to 1200 Ma (Kinnaird et al. 2007). Younger than that, isolated Grenville ages relate to pseudotachylites along certain faults, and to eclogites in the Glenelg-Attadale Inliers. Caledonian and later deformation affected the Outer Hebrides Fault Zone and led to 80–90 km of sinistral offset along the Minches.

Some unanswered questions concerning Lewisian evolution

Several questions posed 20 years ago by Sutton & Watson (1987) remain unanswered. A survey of literature since then shows a dominance of new radiometric dates, with a relatively small contribution from structural and metamorphic studies. There is a limit to the extent to which these scientific strands are integrated with each other and we are optimistic that further advances can be made if this integration is addressed.

Clearly the new dates from the Lewisian have stimulated much thinking on its evolution. However, there are few direct ways to date a structure: cross-cutting dateable intrusions are most useful but not always available. Much recent dating has not had a link to structures as defined on field criteria. Consequently, we still have no direct date for the single most major event to affect the Lewisian – the widespread early Laxfordian event giving rise to huge strains across much of the mainland and almost all the Outer Hebrides, metamorphosing and transposing the Scourie dykes. The 1880 Ma metamorphic age from the South Harris Igneous Complex is key, but relates to a small area of anomalous lithology, structure and metamorphic grade. An improved understanding of how deformation affects isotopic systems would assist progress in this area.

The same issue arises with linking dates to metamorphic events. Multiple events are postulated even within small areas, such as the two granulite facies episodes near Scourie (based on monazite and zircon ages). How can these episodes be identified on a scale larger than a thin section? Are the ‘dominant’ granulite facies assemblages of the Central District, documented a century ago in the Memoir, related mainly to the first or second event?

Part of the problem is that in the past, P–T work was mainly carried out on mafic rocks, which are not amenable for dating (U-poor) whilst most dating has been done on felsic rocks, whose metamorphism is harder to constrain. There is almost no published work on the metamorphism of the dominant quartzofeldspathic rock type nor, indeed, on the amphibolite facies Scourie dykes. The advances in metamorphic modelling of equilibrium assemblages, in understanding how major element zoning in minerals may be linked to process and timescales, and in how isotopes behave during metamorphism, which have been made in the last 20 years are surely of value to future Lewisian research – and could assist specifically with these dominant lithologies. The role of fluids in Lewisian metamorphism should not be ignored. Great amounts of water were added to the Central District during the Inverian episode, and everywhere during the Laxfordian to hydrate the dykes and cause pervasive amphibolitization in areas of previously granulite-facies gneisses. The source of this water is still problematic.

Finally, in terms of crustal growth, we know now that TTG magmatism occurred from 3100 through to 2700 Ma, but was this episodic and did it affect disparate regions not originally close together? We do not know the geometries of magmatic bodies and whether they underplated earlier crustal rocks. The SCLM has been probed via Scourie dyke geochemistry but the tectonic picture of the Laxfordian implies that the Central District may be allochthonous, and displaced from the SCLM it was above at the time of dyke formation. There is much scope for geochemical and tectonic views of the Lewisian to become more integrated.

Acknowledgments

We thank M. Krabbendam and R. Strachan for their reviews, and R. Graham for comments.

  • © The Geological Society of London 2010

References

  1. ↵
    1. Attfield P.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, The structural history of the Canisp Shear Zone, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 165–173.
    OpenUrlAbstract/FREE Full Text
  2. ↵
    1. Baba S.
    (1998) Proterozoic anticlockwise P–T path of the Lewisian Complex of South Harris, Outer Hebrides, NW Scotland. Journal of Metamorphic Geology 16:819–841.
    OpenUrlCrossRefWeb of Science
  3. ↵
    1. Baba S.
    (2004) Palaeoproterozoic UHT metamorphism in the Lewisian complex and North Atlantic region. Journal of Mineralogical and Petrological Sciences 99:202–212.
    OpenUrlCrossRefWeb of Science
  4. ↵
    1. Barruol G.,
    2. Mainprice D.
    (1993) A quantitative evaluation of the contribution of crustal rocks to the shear-wave splitting of teleseismic SKS waves. Physics of the Earth and Planetary Interiors 78:281–300.
    OpenUrlCrossRefWeb of Science
  5. ↵
    1. Beach A.
    (1973) The mineralogy of high temperature shear zones at Scourie, N.W. Scotland. Journal of Petrology 14:231–248.
    OpenUrlCrossRefWeb of Science
  6. ↵
    1. Beach A.
    (1976) The interrelations of fluid transport, deformation, geochemistry and heat flow in early Proterozoic shear zones in the Lewisian complex. Philosophical Transactions of the Royal Society of London A 280:569–604.
    OpenUrlCrossRef
  7. ↵
    1. Beach A.,
    2. Fyfe W. S.
    (1972) Fluid transport and shear zones at Scourie, Sutherland – evidence of overthrusting. Contributions to Mineralogy and Petrology 36:175.
    OpenUrlCrossRefWeb of Science
  8. ↵
    1. Beach A.,
    2. Coward M. P.,
    3. Graham R. H.
    (1974) An interpretation of the structural evolution of the Laxford front, north–west Scotland. Scottish Journal of Geology 9:297–308.
    OpenUrl
  9. ↵
    1. Bowes D. R.
    (1969) in North Atlantic Geology and Continental Drift – A Symposium, The Lewisian of the North–West Highlands of Scotland, American Association of Petroleum Geologists Memoir, ed Kay M. pp 575–594.
  10. ↵
    1. Boyer S. E.,
    2. Elliott D.
    (1982) Thrust systems. The American Association of Petroleum Geologists Bulletin 66:1196–1230.
    OpenUrlAbstract
  11. ↵
    1. Bridgwater D.,
    2. Escher A.,
    3. Watterson J.
    (1973) Tectonic displacements and thermal activity in 2 contrasting proterozoic mobile belts from Greenland. Philosophical Transactions of the Royal Society of London Series A – Mathematical Physical and Engineering Sciences 273:513–533.
    OpenUrlCrossRef
  12. ↵
    1. Brodie K. H.,
    2. Rutter E. H.
    (1985) in Metamorphic Reactions: Kinetics, Textures and Deformation, On the relationship between deformation and metamorphism, with special reference to the behaviour of basic rocks, eds Thompson A. B., Rubie D. C. (Springer-Verlag, Holland) Advances in Physical Geochemistry, pp 138–179.
  13. ↵
    1. Brown M.
    (2007) Metamorphic conditions in orogenic belts: a record of secular change. International Geology Review 49:193–234.
    OpenUrlWeb of Science
  14. ↵
    1. Buchan K. L.,
    2. Mertanen S.,
    3. Park R. G.,
    4. Pesonen L. J.,
    5. Elming S. A.,
    6. Abrahamsen N.,
    7. Bylund G.
    (2000) Comparing the drift of Laurentia and Baltica in the Proterozoic: the importance of key palaeomagnetic poles. Tectonophysics 319:167–198.
    OpenUrlCrossRefWeb of Science
  15. ↵
    1. Burton K. W.,
    2. Cohen A. S.,
    3. O'Nions R. K.,
    4. O'Hara M. J.
    (1994) Archean Crustal Development in the Lewisian Complex of Northwest Scotland. Nature 370:552–555.
    OpenUrlCrossRefWeb of Science
  16. ↵
    1. Butler C. A.,
    2. Holdsworth R. E.,
    3. Strachan R. A.
    (1995) Evidence for Caledonian sinistral strike-slip motion and associated fault zone weakening, Outer Hebrides Fault Zone, NW Scotland. Journal of the Geological Society 152:743–746.
    OpenUrlAbstract/FREE Full Text
  17. ↵
    1. Cartwright I.
    (1988) Crystallization of melts, pegmatite intrusion and the Inverian retrogression of the Scourian complex, Northwest Scotland. Journal of Metamorphic Geology 6:77–93.
    OpenUrlWeb of Science
  18. ↵
    1. Cartwright I.,
    2. Barnicoat A. C.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, Petrology of Scourian supracrustal rocks and orthogneisses from Stoer, NW Scotland: implications for the geological evolution of the Lewisian complex, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 93–107.
    OpenUrlAbstract/FREE Full Text
  19. ↵
    1. Cartwright I.,
    2. Barnicoat A. C.
    (1989) in Evolution of Metamorphic Belts, Evolution of the Scourian complex, Geological Society, London, Special Publications, eds Daly J. S., Cliff R. A., Yardley B. W. D. 27, pp 297–301.
    OpenUrlCrossRef
  20. ↵
    1. Cartwright I.,
    2. Rollinson H. R.
    (1995) Origin of felsic sheets in the Scourian granulites – new evidence from rare-earth elements – discussion. Scottish Journal of Geology 31:91–94.
    OpenUrlFREE Full Text
  21. ↵
    1. Cartwright I.,
    2. Fitches W. R.,
    3. O‘Hara M. J.,
    4. Barnicoat A. C.,
    5. O'Hara S.
    (1985) Archean Supracrustal Rocks from the Lewisian near Stoer, Sutherland. Scottish Journal of Geology 21:187–196.
    OpenUrlAbstract/FREE Full Text
  22. ↵
    1. Chapman H. J.
    (1979) 2 390 Myr Rb–Sr Whole-Rock for the Scourie Dykes of Northwest Scotland. Nature 277:642–643.
    OpenUrlCrossRefWeb of Science
  23. ↵
    1. Cliff R. A.
    (1985) Isotopic dating in metamorphic belts. Journal of the Geological Society, London 142:97–110.
    OpenUrlAbstract/FREE Full Text
    1. Cliff R. A.,
    2. Gray C. M.,
    3. Huhma H.
    (1983) A Sm–Nd Isotopic Study of the South Harris Igneous Complex, The Outer Hebrides. Contributions to Mineralogy and Petrology 82:91–98.
    OpenUrlCrossRefWeb of Science
  24. ↵
    1. Coney P. J.,
    2. Jones D. L.,
    3. Monger J. W. H.
    (1980) Cordilleran suspect terranes. Nature 288:329–333.
    OpenUrlCrossRefWeb of Science
  25. ↵
    1. Corfu F.,
    2. Heaman L. M.,
    3. Rogers G.
    (1994) Polymetamorphic Evolution of the Lewisian Complex, NW Scotland, as Recorded by U–Pb Isotopic Compositions of Zircon, Titanite and Rutile. Contributions to Mineralogy and Petrology 117:215–228.
    OpenUrlCrossRefWeb of Science
  26. ↵
    1. Coward M. P.
    (1973) Structure and Origin of Areas of Anomalously Low-Intensity Finite Deformation in Basement Gneiss Complex of Outer Hebrides. Tectonophysics 16:117–140.
    OpenUrlCrossRefWeb of Science
  27. ↵
    1. Coward M. P.
    (1975) Flat-lying structures within the Lewisian basement gneiss complex of NW Scotland. Proceedings of the Geologists’ Association 85:459–472.
    OpenUrl
  28. ↵
    1. Coward M. P.
    (1984) in Precambrian Tectonics Illustrated, Major shear zones in the Precambrian crust; examples from NW Scotland and southern Africa and their significance, eds Kröner A., Greiling R. (E. Schweizerbart'sche Verlagsbuchhandlung, Stuttgart, Germany), pp 207–235.
  29. ↵
    1. Coward M. P.,
    2. Park R. G.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, The role of mid-crustal shear zones in the Early Proterozoic evolution of the Lewisian, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 127–138.
    OpenUrlAbstract/FREE Full Text
  30. ↵
    1. Coward M. P.,
    2. Francis P. W.,
    3. Graham R. H.,
    4. Watson J.
    (1970) Large-scale Laxfordian structures of Outer Hebrides in relation to those of Scottish Mainland. Tectonophysics 10:425–435.
    OpenUrlCrossRefWeb of Science
  31. ↵
    1. Coward M. P.,
    2. Jan M. Q.,
    3. Rex D.,
    4. Tarney J.,
    5. Thirlwall M.,
    6. Windley B. F.
    (1982) Structural evolution of a crustal section in the Western Himalaya. Nature 295:22–24.
    OpenUrlCrossRef
  32. ↵
    1. Cresswell D.,
    2. Park R. G.
    (1973) in The Early Precambrian of Scotland and Related Rocks of Greenland, The metamorphic history of the Lewisian rocks of the Torridon area in relation to that of the remainder of the southern Laxfordian belt, eds Park R. G., Tarney J. (University of Keele, Keele, UK), pp 105–118.
  33. ↵
    1. Droop G. T. R.,
    2. Fernandes L. A. D.,
    3. Shaw S.
    (1999) Laxfordian metamorphic conditions of the Palaeoproterozoic Loch Maree Group, Lewisian Complex, NW Scotland. Scottish Journal of Geology 35:31–50.
    OpenUrlAbstract/FREE Full Text
  34. ↵
    1. Eskola P.
    (1915) On the relations between the chemical and mineralogical composition in the metamorphic rocks of the Orijarvi region. Bulletin de la Commission géologique de Finlande 44:1–107, (in Finnish) (in English) 109–145.
    OpenUrl
  35. ↵
    1. Evans C. R.
    (1965) Geochronology of the Lewisian basement near Lochinver, Sutherland. Nature 207:54–55.
    OpenUrlCrossRefWeb of Science
  36. ↵
    1. Fettes D. J.,
    2. Mendum J. R.,
    3. Smith D. I.,
    4. Watson J. V.
    (1992) Geology of the Outer Hebrides (British Geological Survey Memoir, London, HMSO).
  37. ↵
    1. Flinn D.
    (1965) On the symmetry principle in the deformation ellipsoid. Geological Magazine 102:36–45.
    OpenUrlAbstract
    1. Friend C. R. L.,
    2. Kinny P. D.
    (1995) New evidence for protolith ages of Lewisian granulites, Northwest Scotland. Geology 23:1027–1030.
    OpenUrlAbstract/FREE Full Text
  38. ↵
    1. Friend C. R. L.,
    2. Kinny P. D.
    (2001) A reappraisal of the Lewisian Gneiss Complex: geochronological evidence for its tectonic assembly from disparate terranes in the Proterozoic. Contributions to Mineralogy and Petrology 142:198–218.
    OpenUrlCrossRefWeb of Science
  39. ↵
    1. Friend C. R. L.,
    2. Strachan R. A.,
    3. Kinny P. D.
    (2008) U–Pb zircon dating of basement inliers within the Moine Supergroup, Scottish Caledonides: implications of Archaean protolith ages. Journal of the Geological Society 165:807–815.
    OpenUrlAbstract/FREE Full Text
  40. ↵
    1. Giletti B. J.,
    2. Moorbath S.,
    3. Lambert R. S.
    (1961) A geochronological study of the metamorphic complexes of the Scottish Highlands. Quarterly Journal of the Geological Society, London 117:233–264.
    OpenUrlCrossRef
  41. ↵
    1. Goodenough K. M.,
    2. Park R. G.,
    3. et al.
    (2010) in Continental Tectonics and Mountain Building: The Legacy of Peach and Horne, The Laxford Shear Zone: an end-Archaean terrane boundary? Geological Society, London, Special Publications, eds Law R. D., Butler R. W. H., Holdsworth R. E., Krabbendam M., Strachan R. A. 335, pp 103–120.
    OpenUrlAbstract/FREE Full Text
  42. ↵
    1. Gower C. F.,
    2. Krogh T. E.
    (2002) A U–Pb geochronological review of the Proterozoic history of the eastern Grenville Province. Canadian Journal of Earth Sciences 39:795–829.
    OpenUrl
  43. ↵
    1. Graham R. H.
    (1980) The Role of Shear Belts in the Structural Evolution of the South Harris Igneous Complex. Journal of Structural Geology 2:29–37.
    OpenUrlCrossRefWeb of Science
    1. Graham R. H.,
    2. Coward M. P.
    (1973) in The Early Precambrian of Scotland and Related Rocks of Greenland, The Laxfordian of the Outer Hebrides, eds Park R. G., Tarney J. (University of Keele, Keele, UK), pp 85–93.
  44. ↵
    1. Hall J.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, Physical properties of Lewisian rocks: implications for deep crustal structure, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 185–192.
    OpenUrlAbstract/FREE Full Text
    1. Heaman L. M.,
    2. & Tarney J.
    (1989) U–Pb Baddeleyite ages for the Scourie Dyke Swarm, Scotland – evidence for 2 distinct intrusion events. Nature 340:705–708.
    OpenUrlCrossRefWeb of Science
  45. ↵
    1. Hollis J. A.,
    2. Harley S. L.,
    3. White R. W.,
    4. Clarke G. L.
    (2006) Preservation of evidence for prograde metamorphism in ultrahigh-temperature, high-pressure kyanite-bearing granulites, South Harris, Scotland. Journal of Metamorphic Geology 24:263–279.
    OpenUrlCrossRefWeb of Science
  46. ↵
    1. Humphries F. J.,
    2. Cliff R. A.
    (1982) Sm–Nd dating and cooling history of Scourian granulites, Sutherland. Nature 295:515–517.
    OpenUrlCrossRefWeb of Science
  47. ↵
    1. Ireland T. R.,
    2. Williams I. S.
    (2003) Considerations in zircon geochronology by SIMS. Reviews in Mineralogy & Geochemistry 35:215–241.
    OpenUrl
  48. ↵
    1. Kalsbeek F.,
    2. Austrheim H.,
    3. Bridgwater D.,
    4. Hansen B. T.,
    5. Pedersen S.,
    6. Taylor P. N.
    (1993) Geochronology of Archean and Proterozoic events in the Ammassalik area, south–east Greenland, and comparisons with the Lewisian of Scotland and the Nagssugtoqidian of west Greenland. Precambrian Research 62:239–270.
    OpenUrlCrossRefWeb of Science
  49. ↵
    1. Kelly A.,
    2. England R. W.,
    3. Maguire P. K. H.
    (2007) A crustal seismic velocity model for the UK, Ireland and surrounding seas. Geophysical Journal International 171:1172–1184.
    OpenUrlCrossRef
  50. ↵
    1. Kinnaird T. C.,
    2. Prave A. R.,
    3. Kirkland C. L.,
    4. Horstwood M.,
    5. Parrish R.,
    6. Batchelor R. A.
    (2007) The late Mesoproterozoic-early Neoproterozoic tectonostratigraphic evolution of NW Scotland: the Torridonian revisited. Journal of the Geological Society 164:541–551.
    OpenUrlAbstract/FREE Full Text
  51. ↵
    1. Kinny P. D.,
    2. Friend C. R. L.
    (1997) U–Pb isotopic evidence for the accretion of different crustal blocks to form the Lewisian Complex of northwest Scotland. Contributions to Mineralogy and Petrology 129:326–340.
    OpenUrlCrossRefWeb of Science
  52. ↵
    1. Kinny P. D.,
    2. Friend C. R. L.,
    3. Love G. J.
    (2005) Proposal for a terrane-based nomenclature for the Lewisian Gneiss Complex of NW Scotland. Journal of the Geological Society 162:175–186.
    OpenUrlAbstract/FREE Full Text
  53. ↵
    1. Lisle R. J.
    (1993) Strike-slip motion in the Minches, NW Scotland, deduced from the trends of the Scourie dyke swarm. Journal of the Geological Society 150:653–656.
    OpenUrlAbstract/FREE Full Text
    1. Love G. J.,
    2. Kinny P. D.,
    3. Friend C. R. L.
    (2004) Timing of magmatism and metamorphism in the Gruinard Bay area of the Lewisian Gneiss Complex: comparisons with the Assynt Terrane and implications for terrane accretion. Contributions to Mineralogy and Petrology 146:620–636.
    OpenUrlCrossRefWeb of Science
  54. ↵
    1. MacDonald R.,
    2. Fettes D. J.
    (2007) The tectonomagmatic evolution of Scotland. Transactions of the Royal Society of Edinburgh-Earth Sciences, 97:213–295.
  55. ↵
    1. Mason A. J.,
    2. Parrish R. R.,
    3. Brewer T. S.
    (2004) U–Pb geochronology of Lewisian orthogneisses in the Outer Hebrides, Scotland: implications for the tectonic setting and correlation of the South Harris Complex. Journal of the Geological Society 161:45–54.
    OpenUrlAbstract/FREE Full Text
  56. ↵
    1. Menzies M. A.,
    2. Halliday A. N.,
    3. Hunter R. H.,
    4. MacIntyre R. M.,
    5. Upton B. G. J.
    (1986) in Kimberlites and Related Rocks, The age, composition and significance of a xenolith-bearing monchiquite dike, Lewis, Scotland, Geological Society, Australia, Special Publication, eds Ross J., Jaques A. L., Ferguson J., Green D. H., O'Reilly S. Y., Danchin R. V., Janse A. J. A. 14, pp 843–852.
    OpenUrl
  57. ↵
    1. Moorbath S.,
    2. Powell J. L.,
    3. Taylor P. N.
    (1975) Isotopic evidence for the age and origin of the ‘grey gneiss’ complex of the southern Outer Hebrides, Scotland. Journal of the Geological Society 131:213–222.
    OpenUrlAbstract/FREE Full Text
  58. ↵
    1. Moorbath S.,
    2. Welke H.,
    3. Gale N. H.
    (1969) Significance of lead isotope studies in ancient, high-grade metamorphic basement complexes, as exemplified by Lewisian rocks of Northwest Scotland. Earth and Planetary Science Letters 6, 256.
  59. ↵
    1. Myers J. S.
    (1971) The Late Laxfordian granite-migmatite complex of western Harris, Outer Hebrides. Scottish Journal of Geology 7:254–284.
    OpenUrlAbstract/FREE Full Text
  60. ↵
    1. O‘Hara M. J.
    (1961) Petrology of the Scourie dyke, Sutherland. Mineralogical Magazine 32:848–865.
    OpenUrlCrossRef
  61. ↵
    1. O'Hara M. J.
    (1977) Thermal history of excavation of Archaean gneisses from the base of the continental crust. Journal of the Geological Society, London 134:185–200.
    OpenUrlAbstract/FREE Full Text
  62. ↵
    1. O'Hara M. J.,
    2. Yarwood G.
    (1978) High pressure-temperature point on an Archaean geotherm, implied magma genesis by crustal anatexis, and consequences for garnet-pyroxene thermometry and barometry. Philosophical Transactions of the Royal Society of London Series a-Mathematical Physical and Engineering Sciences 288:441–453.
    OpenUrlCrossRef
  63. ↵
    1. Odling N. E.
    (1984) Strain analysis and strain path modeling in the Loch Tollie Gneisses, Gairloch, NW Scotland. Journal of Structural Geology 6:543–562.
    OpenUrlCrossRefWeb of Science
  64. ↵
    1. Park A. F.
    (1991) Continental growth by accretion – a tectonostratigraphic terrane analysis of the evolution of the western and central Baltic shield, 2.50 to 1.75 Ga. Geological Society of America Bulletin 103:522–537.
    OpenUrlAbstract/FREE Full Text
  65. ↵
    1. Park R. G.
    (1964) The structural history of the Lewisian rocks of Gairloch, Wester Ross. Quarterly Journal of the Geological Society, London 120:397–434.
    OpenUrl
  66. ↵
    1. Park R. G.
    (1970) Observations on Lewisian chronology. Scottish Journal of Geology 6:379–399.
    OpenUrlAbstract/FREE Full Text
  67. ↵
    1. Park R. G.
    (1994) Early Proterozoic tectonic overview of the northern British Isles and neighboring terrains in Laurentia and Baltica. Precambrian Research 68:65–79.
    OpenUrlCrossRefWeb of Science
  68. ↵
    1. Park R. G.
    (1995) in Early Precambrian Processes, Palaeoproterozoic Laurentia–Baltica relationships: a view from the Lewisian, Geological Society, London, Special Publications, eds Coward M. P., Ries A. C. 95, pp 211–224.
    OpenUrlAbstract/FREE Full Text
  69. ↵
    1. Park R. G.
    (2005) The Lewisian terrane model: a review. Scottish Journal of Geology 41:105–118.
    OpenUrlAbstract/FREE Full Text
  70. ↵
    1. Park R. G.,
    2. Cresswell D.
    (1973) in The Early Precambrian of Scotland and Related Rocks of Greenland, The dykes of the Laxfordian belt, eds Park R. G., Tarney J. (University of Keele, Keele, UK), pp 119–130.
  71. ↵
    1. Park R. G.,
    2. Tarney J.
    (1973) The Early Precambrian of Scotland and Related Rocks of Greenland (University of Keele, Keele, UK).
  72. ↵
    1. Park R. G.,
    2. Tarney J.
    (1987) Evolution of the Lewisian and Comparable Precambrian High Grade Terrains (Geological Society, London).
  73. ↵
    1. Park R. G.,
    2. Stewart A. D.,
    3. Wright D. T.
    (2002) in The Geology of Scotland, The Hebridean terrane, ed Trewin N. H. (Geological Society, London), pp 45–80.
  74. ↵
    1. Park R. G.,
    2. Tarney J.,
    3. Connelly J. N.
    (2001) The Loch Maree Group: Palaeoproterozoic subduction-accretion complex in the Lewisian of NW Scotland. Precambrian Research 105:205–226.
    OpenUrlCrossRefWeb of Science
  75. ↵
    1. Parrish R. R.
    (2001) in Continental Reactivation and Reworking, The response of mineral chronometers to metamorphism and deformation in orogenic belts, Geological Society, London, Special Publications, eds Miller J. A., Holdsworth R. E., Buick I. S., Hand M. 184, pp 289–301.
    OpenUrlAbstract/FREE Full Text
  76. ↵
    1. Parrish R. R.,
    2. Noble S. R.
    (2003) Zircon U–Th–Pb geochronology by isotope dilution – thermal ionization mass spectrometry (ID–TIMS). Reviews in Mineralogy & Geochemistry 53:183–213.
    OpenUrlFREE Full Text
  77. ↵
    1. Peach B. N.,
    2. Horne J.,
    3. Gunn W.,
    4. Clough C. T.,
    5. Hinxman L. W.,
    6. Teall J. J. H.
    (1907) The Geological Structure of the North–West Highlands of Scotland (Memoirs of the Geological Survey, London) Her Majesty's Stationery Office.
  78. ↵
    1. Pidgeon R. T.,
    2. Bowes D. R.
    (1972) Zircon U–Pb ages of granulites from central region of Lewisian, Northwestern Scotland. Geological Magazine 109, 258.
  79. ↵
    1. Piper J. D. A.
    (1992) Post-Laxfordian magnetic imprint in the Lewisian metamorphic complex and strike-slip motion in the Minches, NW Scotland. Journal of the Geological Society 149:127–137.
    OpenUrlAbstract/FREE Full Text
  80. ↵
    1. Powell C. M. R.,
    2. Sinha M. C.
    (1987) The PUMA experiment west of Lewis, UK. Geophysical Journal of the Royal Astronomical Society 89:259–264.
    OpenUrlWeb of Science
  81. ↵
    1. Ramsay J. G.,
    2. Graham R. H.
    (1970) Strain variation in shear belts. Canadian Journal of Earth Sciences 7:786–813.
    OpenUrlAbstract
  82. ↵
    1. Rollinson H.
    (1994) Origin of felsic sheets in the Scourian granulites – new evidence from rare-earth elements. Scottish Journal of Geology 30:121–129.
    OpenUrlAbstract/FREE Full Text
  83. ↵
    1. Rollinson H. R.
    (1982) Evidence from feldspar compositions of high-temperatures in granite sheets in the Scourian complex, NW Scotland. Mineralogical Magazine 46:73–76.
    OpenUrlCrossRefWeb of Science
  84. ↵
    1. Rollinson H. R.
    (1987) Early basic magmatism in the evolution of Archean high-grade gneiss terrains – an example from the Lewisian of NW Scotland. Mineralogical Magazine 51:345–355.
    OpenUrlCrossRefWeb of Science
  85. ↵
    1. Rollinson H. R.
    (1996) in Precambrian Crustal Evolution in the North Atlantic Region, Tonalite-trondhjemite-granodiorite magmatism and the genesis of Lewisian crust during the Archaean, Geological Society, London, Special Publications, ed Brewer T. S. 112, pp 25–42.
    OpenUrlAbstract/FREE Full Text
  86. ↵
    1. Rollinson H. R.
    (2006) in Evolution of the Continental Crust, Archaean Crustal Evolution, eds Brown M., Rushmer T. (Cambridge University Press, Cambridge), pp 174–231.
  87. ↵
    1. Rollinson H. R.
    (2007) Early Earth Systems: A Geochemical Approach (Blackwell Publishing, Oxford, UK).
  88. ↵
    1. Rollinson H. R.,
    2. Fowler M. B.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, The magmatic evolution of the Scourian complex at Gruinard Bay, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 57–71.
    OpenUrlAbstract/FREE Full Text
  89. ↵
    1. Rollinson H. R.,
    2. Tarney J.
    (2005) Adakites – The key to understanding LILE depletion in granulites. Lithos 79:61–81.
    OpenUrlCrossRefWeb of Science
  90. ↵
    1. Rollinson H. R.,
    2. Windley B. F.
    (1980) An Archean granulite-grade tonalite-trondhjemite-granite suite from Scourie, NW Scotland – geochemistry and origin. Contributions to Mineralogy and Petrology 72:265–281.
    OpenUrlCrossRefWeb of Science
  91. ↵
    1. Rudnick R. L.,
    2. Fountain D. M.
    (1995) Nature and composition of the continental crust – a lower crustal perspective. Reviews of Geophysics 33:267–309.
    OpenUrlCrossRefWeb of Science
  92. ↵
    1. Rutherford E.
    (1905) Present problems in radioactivity. Popular Science Monthly 57:1–34.
    OpenUrl
  93. ↵
    1. Sanders I. S.,
    2. Vancalsteren P. W. C.,
    3. Hawkesworth C. J.
    (1984) A Grenville Sm–Nd age for the Glenelg eclogite in Northwest Scotland. Nature 312:439–440.
    OpenUrlCrossRefWeb of Science
  94. ↵
    1. Shackleton R. M.,
    2. Ries A. C.
    (1984) The relation between regionally consistent stretching lineations and plate motions. Journal of Structural Geology 6:111–117.
    OpenUrlCrossRefWeb of Science
  95. ↵
    1. Sheraton J. W.,
    2. Tarney J.,
    3. Wright A. E.
    (1973) in The Early Precambrian of Scotland and Related Rocks of Greenland, The geochemistry of the Scourian gneisses of the Assynt district, eds Park R. G., Tarney J. (University of Keele, Keele, UK), pp 13–30.
  96. ↵
    1. Sherlock S. C.,
    2. Jones K. A.,
    3. Park R. G.
    (2008) Grenville age pseudotachylyte in the Lewisian: laserprobe 40Ar/39Ar ages from the Gairloch region of Scotland (UK). Journal of the Geological Society 165:73–84.
    OpenUrlAbstract/FREE Full Text
  97. ↵
    1. Sibson R. H.
    (1977) Fault rocks and fault mechanisms. Journal of the Geological Society, London 133:191–213.
    OpenUrlAbstract/FREE Full Text
  98. ↵
    1. Sills J. D.,
    2. Rollinson H. R.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, Metamorphic evolution of the mainland Lewisian complex, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 81–92.
    OpenUrlAbstract/FREE Full Text
  99. ↵
    1. Smythe D. K.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, Deep seismic reflection profiling of the Lewsian foreland, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 193–203.
    OpenUrlAbstract/FREE Full Text
  100. ↵
    1. Storey C.
    (2008) The Glenelg-Attadale Inlier, NW Scotland, with emphasis on the Precambrian high-pressure metamorphic history and subsequent retrogression: an introduction and review. Scottish Journal of Geology 44:1–16.
    OpenUrlAbstract/FREE Full Text
  101. ↵
    1. Storey C. D.
    (2002) Tectono-metamorphic evolution of the Glenelg-Attadale Inlier, northwest Scotland. Unpublished PhD thesis (University of Leicester).
  102. ↵
    1. Storey C. D.,
    2. Brewer T. S.,
    3. Temperley S.
    (2005) P–T conditions of Grenville-age eclogite facies metamorphism and amphibolite facies retrogression of the Glenelg-Attadale Inlier, NW Scotland. Geological Magazine 142:605–615.
    OpenUrlAbstract/FREE Full Text
  103. ↵
    1. Sutton J.,
    2. Watson J.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, The Lewisian Complex: questions for the future, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 7–12.
    OpenUrlAbstract/FREE Full Text
  104. ↵
    1. Sutton J.,
    2. Watson J. V.
    (1951) The pre-Torridonian metamorphic history of the Loch Torridon and Scourie areas in the NW Highlands and its bearing on the chronological classification of the Lewisian. Journal of the Geological Society, London 106:241–308.
    OpenUrl
  105. ↵
    1. Tarney J.
    (1963) Assynt dykes and their metamorphism. Nature 199:672–674.
    OpenUrlCrossRefWeb of Science
  106. ↵
    1. Tarney J.
    (1973) in The Early Precambrian of Scotland and Related Rocks of Greenland, The Scourie dyke suite and the nature of the Inverian event in Assynt, eds Park R. G., Tarney J. (University of Keele, Keele, UK), pp 105–118.
  107. ↵
    1. Tarney J.,
    2. Weaver B. L.
    (1987a) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, Geochemistry of the Scourian Complex: petrogenesis and tectonic models, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 45–56.
    OpenUrlAbstract/FREE Full Text
  108. ↵
    1. Tarney J.,
    2. Weaver B. L.
    (1987b) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, Mineralogy, petrology and geochemistry of the Scourie dykes: petrogenesis and crystallization processes in dykes intruded at depth, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 217–233.
    OpenUrlAbstract/FREE Full Text
  109. ↵
    1. Teall J. J. H.
    (1885) The metamorphism of a dolerite dyke into hornblende schist. Quarterly Journal of the Geological Society, London 41:133–144.
    OpenUrl
  110. ↵
    1. Teall J. J. H.
    (1891) On an eclogite from Loch Duich. Mineralogical Magazine 9:217–218.
    OpenUrlCrossRef
  111. ↵
    1. Waters F. G.,
    2. Cohen A. S.,
    3. O'Nions R. K.,
    4. O'Hara M. J.
    (1990) Development of Archean lithosphere deduced from chronology and isotope chemistry of Scourie dykes. Earth and Planetary Science Letters 97:241–255.
    OpenUrlCrossRefWeb of Science
  112. ↵
    1. Weaver B. L.,
    2. Tarney J.
    (1981) Lewisian gneiss geochemistry and Archaean crustal development models. Earth and Planetary Science Letters 55:171–180.
    OpenUrlCrossRefWeb of Science
  113. ↵
    1. Wheeler J.
    (2007) in Deformation of the Continental Crust: The Legacy of Mike Coward, A major high strain zone in the Lewisian Complex in the Loch Torridon area, NW Scotland: insights into deep crustal deformation, Geological Society, London, Special Publications, eds Ries A. C., Butler R. W. H., Graham R. H. 272, pp 27–45.
    OpenUrlAbstract/FREE Full Text
  114. ↵
    1. Wheeler J.,
    2. Butler R. W. H.
    (1994) Criteria for identifying structures related to true crustal extension in orogens. Journal of Structural Geology 16:1023–1027.
    OpenUrlCrossRefWeb of Science
  115. ↵
    1. Wheeler J.,
    2. Windley B. F.,
    3. Davies F. B.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, Internal evolution of the major Precambrian shear belt at Torridon, NW Scotland, Geological Society, London, Special Publications, eds Park R. G., Tarney J. 27, pp 153–164.
    OpenUrlAbstract/FREE Full Text
  116. ↵
    1. White S. H.,
    2. Glasser J.
    (1987) in Evolution of the Lewisian and Comparable Precambrian High Grade Terrains, The Outer Hebrides Fault Zone: evidence for normal movements, eds Park R. G., Tarney J. (Geological Society, London), 27, pp 175–183.
    OpenUrl
  117. ↵
    1. Whitehouse M. J.
    (1993) Age of the Corodale gneisses, South Uist. Scottish Journal of Geology 29:1–7.
    OpenUrlAbstract/FREE Full Text
  118. ↵
    1. Whitehouse M. J.,
    2. Bridgwater D.
    (2001) Geochronological constraints on Paleoproterozoic crustal evolution and regional correlations of the northern Outer Hebridean Lewisian complex, Scotland. Precambrian Research 105:227–245.
    OpenUrlCrossRefWeb of Science
    1. Whitehouse M. J.,
    2. Bridgwater D.,
    3. Park R. G.
    (1997) Detrital zircon ages from the Loch Maree Group, Lewisian Complex, NW Scotland: confirmation of a Palaeoproterozoic Laurentia-Fennoscandia connection. Terra Nova 9:260–263.
    OpenUrlCrossRefWeb of Science
  119. ↵
    1. Whitehouse M. J.,
    2. Fowler M. B.,
    3. Friend C. R .L.
    (1996) Conflicting mineral and whole-rock isochron ages from the Late-Archaean Lewisian Complex of northwestern Scotland: implications for geochronology in polymetamorphic high-grade terrains. Geochimica et Cosmochimica Acta 60:3085–3102.
    OpenUrlCrossRefWeb of Science
  120. ↵
    1. Wynn T. J.
    (1995) in Early Precambrian Processes, Deformation in the mid to lower continental crust: analogues from Proterozoic shear zones in NW Scotland, Geological Society, London, Special Publications, eds Coward M. P., Ries A. C. 95, pp 225–241.
    OpenUrlAbstract/FREE Full Text
    1. Zhu X. K.,
    2. Onions R. K.,
    3. Belshaw N. S.,
    4. Gibb A. J.
    (1997) Lewisian crustal history from in situ SIMS mineral chronometry and related metamorphic textures. Chemical Geology 136:205–218.
    OpenUrlCrossRefWeb of Science
PreviousNext
Back to top

In this volume

Geological Society, London, Special Publications: 335 (1)
Geological Society, London, Special Publications
Volume 335
2010
  • Table of Contents
  • Table of Contents (PDF)
  • About the Cover
  • Index by author
  • Back Matter (PDF)
  • Front Matter (PDF)
Alerts
Sign In to Email Alerts with your Email Address
Citation tools

The Lewisian Complex: insights into deep crustal evolution

J. Wheeler, R. G. Park, H. R. Rollinson and A. Beach
Geological Society, London, Special Publications, 335, 51-79, 1 January 2010, https://doi.org/10.1144/SP335.4
J. Wheeler
1Department of Earth and Ocean Sciences, Jane Herdman Building, Liverpool University, Liverpool L69 3GP, UK
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: [email protected]
R. G. Park
212 Provost Ferguson Drive, Tain, Ross-shire, IV19 1RE, UK
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
H. R. Rollinson
3Department of Geographical, Earth and Environmental Sciences, University of Derby, Kedleston Road, Derby DE22 1GB, UK
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
A. Beach
4Exploration Outcomes, 1 Huntly Gardens, Glasgow G12 9AS, UK
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Permissions
View PDF
Share

The Lewisian Complex: insights into deep crustal evolution

J. Wheeler, R. G. Park, H. R. Rollinson and A. Beach
Geological Society, London, Special Publications, 335, 51-79, 1 January 2010, https://doi.org/10.1144/SP335.4
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
Email to

Thank you for sharing this Geological Society, London, Special Publications article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
The Lewisian Complex: insights into deep crustal evolution
(Your Name) has forwarded a page to you from Geological Society, London, Special Publications
(Your Name) thought you would be interested in this article in Geological Society, London, Special Publications.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Print
Download PPT
Bookmark this article
  • Tweet Widget
  • Facebook Like
  • Google Plus One
  • Article
    • Abstract
    • The Lewisian Complex as understood in the Peach et al. 1907 memoir
    • The Lewisian Complex a century after the memoir
    • Summary of Lewisian evolution
    • Some unanswered questions concerning Lewisian evolution
    • Acknowledgments
    • References
  • Figures & Data
  • Info & Metrics
  • PDF

Related Articles

Similar Articles

Cited By...

  • Most read
  • Most cited
Loading
  • The history of the European oil and gas industry (1600s–2000s)
  • Himalayan earthquakes: a review of historical seismicity and early 21st century slip potential
  • Introduction to Himalayan tectonics: a modern synthesis
  • Seismic characterization of carbonate platforms and reservoirs: an introduction and review
  • An introduction to forensic soil science and forensic geology: a synthesis
More...

Special Publications

  • About the series
  • Books Editorial Committee
  • Submit a book proposal
  • Author information
  • Supplementary Publications
  • Subscribe
  • Pay per view
  • Copyright & Permissions
  • Activate Online Subscription
  • Feedback
  • Help

Lyell Collection

  • About the Lyell Collection
  • Lyell Collection homepage
  • Collections
  • Open Access Collection
  • Open Access Policy
  • Lyell Collection access help
  • Recommend to your Library
  • Lyell Collection Sponsors
  • MARC records
  • Digital preservation
  • Developing countries
  • Geofacets
  • Manage your account
  • Cookies

The Geological Society

  • About the Society
  • Join the Society
  • Benefits for Members
  • Online Bookshop
  • Publishing policies
  • Awards, Grants & Bursaries
  • Education & Careers
  • Events
  • Geoscientist Online
  • Library & Information Services
  • Policy & Media
  • Society blog
  • Contact the Society

Published by The Geological Society of London, registered charity number 210161

Print ISSN 
0305-8719
Online ISSN 
2041-4927

Copyright © 2022 Geological Society of London