15.02.2013 Views

A new mechanism for “dome and keel” geom - Department of ...

A new mechanism for “dome and keel” geom - Department of ...

A new mechanism for “dome and keel” geom - Department of ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Precambrian Research 212-213 (2012) 139–154<br />

Contents lists available at SciVerse ScienceDirect<br />

Precambrian Research<br />

journal homepage: www.elsevier.com/locate/precamres<br />

Regional shortening followed by channel flow induced collapse:<br />

A <strong>new</strong> <strong>mechanism</strong> <strong>for</strong> <strong>“dome</strong> <strong>and</strong> <strong>keel”</strong> <strong>geom</strong>etries in Neoarchaean<br />

granite-greenstone terrains<br />

Lyal B. Harris a,∗ , Laurent Godin b , Chris Yakymchuk b,1<br />

a Institut national de la recherche scientifique, Centre - Eau Terre Environnement, 490 de la Couronne, Québec, (Québec) G1K 9A9, Canada<br />

b <strong>Department</strong> <strong>of</strong> Geological Sciences <strong>and</strong> Geological Engineering, Queen’s University, Kingston, (Ontario) K7L 3N6, Canada<br />

article info<br />

Article history:<br />

Received 17 June 2011<br />

Received in revised <strong>for</strong>m 18 April 2012<br />

Accepted 27 April 2012<br />

Available online 9 May 2012<br />

Keywords:<br />

Channel flow<br />

Folding<br />

Centrifuge modelling<br />

“Dome <strong>and</strong> <strong>keel”</strong> <strong>geom</strong>etry<br />

Granite-greenstone terrains<br />

Archaean tectonics<br />

1. Introduction<br />

1.1. Research aims<br />

abstract<br />

Centrifuge analogue modelling is used to test a <strong>new</strong> hypothesis<br />

<strong>for</strong> the development <strong>of</strong> dome <strong>and</strong> keel structures in Archaean<br />

∗ Corresponding author. Tel.: +1 418 654 2568; fax: +1 418 654 2600.<br />

E-mail addresses: lyal harris@ete.inrs.ca (L.B. Harris),<br />

godin@geol.queensu.ca (L. Godin), cyak@umd.edu (C. Yakymchuk).<br />

1 Present address: <strong>Department</strong> <strong>of</strong> Geology, University <strong>of</strong> Maryl<strong>and</strong>, College Park,<br />

MD 20742, USA.<br />

0301-9268/$ – see front matter © 2012 Elsevier B.V. All rights reserved.<br />

http://dx.doi.org/10.1016/j.precamres.2012.04.022<br />

The lateral flow <strong>and</strong> extrusive exhumation <strong>of</strong> ductile migmatitic gneisses due to a horizontal gradient in<br />

lithostatic pressure, a process termed “channel flow” or “lateral protrusion”, has previously been proposed<br />

as an important tectonic process in large hot orogens. Centrifuge simulation <strong>of</strong> (i) layer-parallel shortening<br />

followed by (ii) collapse <strong>of</strong> a cover sequence during ductile flow <strong>of</strong> underlying layers in an analogous manner<br />

to channel flow suggests a <strong>new</strong> <strong>mechanism</strong> <strong>for</strong> the development <strong>of</strong> structures within Neoarchaean<br />

granite-greenstone terrains. The centrifuge model incorporates an upper package <strong>of</strong> silicone-modelling<br />

clay microlaminates that simulate a greenstone sequence that overlies slightly less dense ductile silicone<br />

putties, whose rheological properties simulate migmatitic felsic gneiss. A low viscosity, low-density layer<br />

along half <strong>of</strong> the infrastructure–superstructure interface simulates the presence <strong>of</strong> granitoid melt. During<br />

initial layer-parallel shortening upright folds <strong>for</strong>m in the upper “greenstone” package <strong>and</strong> the interface<br />

with underlying ductile layers is folded. In basal silicone layers, recumbent to overturned non-cylindrical<br />

<strong>and</strong> upright folds <strong>for</strong>m with increasing distance from the moving end wall. Removal <strong>of</strong> material parallel<br />

to fold hinges at the ‘<strong>for</strong>el<strong>and</strong>’ end <strong>of</strong> the model (i.e. furthest from the ram used to shorten models) in<br />

several stages simulates erosion <strong>and</strong> the difference in gravitational loading thus created induces ductile<br />

flow <strong>and</strong> lateral extrusion <strong>of</strong> basal silicone layers. Models aim to reproduce features comparable to<br />

those developed during extrusive channel flow <strong>and</strong> focused exhumation <strong>of</strong> basement migmatitic gneiss<br />

in nature. Early-<strong>for</strong>med recumbent to inclined folds are then accentuated during simulated channel flow,<br />

while <strong>new</strong> recumbent isoclinal folds in basal layers develop. Broad anti<strong>for</strong>ms <strong>and</strong> tight syn<strong>for</strong>ms similar<br />

to the <strong>“dome</strong> <strong>and</strong> <strong>keel”</strong> <strong>geom</strong>etry that typifies many Archaean granite-greenstone belts are produced as<br />

a late feature in the model. Channel flow <strong>and</strong> collapse <strong>of</strong> a fold-thickened crust is there<strong>for</strong>e proposed as a<br />

potential alternative <strong>mechanism</strong> <strong>for</strong> the <strong>for</strong>mation <strong>of</strong> structures in some Neoarchean granite-greenstone<br />

terrains. By analogy with other physical <strong>and</strong> numerical models, channel flow in the Neoarchaean may<br />

be enhanced by impingement <strong>of</strong> an upper mantle wedge into the base <strong>of</strong> the crust. Our results imply<br />

that both lithospheric shortening <strong>and</strong> non-diapiric gravitational instabilities may be responsible <strong>for</strong> the<br />

<strong>for</strong>mation <strong>of</strong> some Archaean <strong>“dome</strong> <strong>and</strong> <strong>keel”</strong> structures <strong>and</strong> may account <strong>for</strong> the juxtaposition <strong>of</strong> some<br />

granite-greenstone <strong>and</strong> high-grade gneiss terrains in the Archaean. Model results also show similarities<br />

to structures produced during lateral flow <strong>and</strong> withdrawal <strong>of</strong> salt in fold belts.<br />

© 2012 Elsevier B.V. All rights reserved.<br />

granite-greenstone belts. The progressive development <strong>of</strong> structures<br />

is studied in a model scaled to represent de<strong>for</strong>mation <strong>of</strong> a<br />

greenstone sequence <strong>of</strong> layered volcanic <strong>and</strong> sedimentary rocks<br />

upon basement granitoid gneisses. Models simulate two separate<br />

de<strong>for</strong>mation stages: (1) initial folding <strong>and</strong> crustal thickening<br />

during layer-parallel shortening, followed by (2) a separate period<br />

<strong>of</strong> “collapse” <strong>and</strong> gravity-driven re-equilibration during lateral<br />

ductile flow <strong>and</strong> <strong>of</strong> a ductile substrate analogous to extrusive<br />

channel flow (see Section 1.3 <strong>for</strong> definition). The model presented<br />

in this paper also illustrates <strong>mechanism</strong>s <strong>for</strong> the <strong>for</strong>mation <strong>of</strong><br />

overturned to recumbent folds in granite-greenstone terrains,<br />

their relative timing <strong>of</strong> <strong>for</strong>mation with respect to folds in the


140 L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154<br />

overlying greenstone sequence, <strong>and</strong> their spatial distribution.<br />

The likelihood <strong>for</strong> <strong>and</strong> factors promoting channel flow in the<br />

Neoarchaean <strong>and</strong> possible regional implications <strong>for</strong> juxtaposition<br />

<strong>of</strong> granite-greenstone <strong>and</strong> high-grade gneiss terrains are discussed.<br />

1.2. Previous models <strong>for</strong> dome <strong>and</strong> keel structures <strong>and</strong> recumbent<br />

folds in granite-greenstone belts<br />

The origin <strong>of</strong> tight syn<strong>for</strong>ms, “keels”, or “cusps” <strong>of</strong><br />

(meta-)sedimentary <strong>and</strong> (meta-)volcanic rocks <strong>of</strong> greenstone<br />

belts between open, rounded anti<strong>for</strong>ms, “arches”, or domes<br />

<strong>of</strong> tonalite–trondhjemite–granodiorite (TTG) gneiss that typify<br />

Archaean granitoid-greenstone belts (Condie, 1984; Windley,<br />

1995) is contentious. The following diverse <strong>mechanism</strong>s have been<br />

proposed:<br />

(i) Folding during one or more phases <strong>of</strong> regional shortening<br />

(e.g. Snowden <strong>and</strong> Bickle, 1976; Myers <strong>and</strong> Watkins, 1985;<br />

Blewett, 2002; Blewett et al., 2004, 2010). Weinberg et al. (2003)<br />

illustrate that granite-greenstone belt fold <strong>geom</strong>etry may be<br />

influenced by early granitoid intrusions that acted as rigid bodies<br />

during shortening.<br />

(ii) Domes <strong>for</strong>med due to Raleigh–Taylor (Wilcock, 1991), “gravitational”<br />

instabilities created by density inversion. Domal<br />

structures similar to those developed in centrifuge models<br />

<strong>of</strong> Ramberg (1967a,b, 1981a,b) <strong>and</strong> Dixon <strong>and</strong> Summers<br />

(1985) may occur where a denser greenstone sequence (ca.<br />

2.7–3.0 g/cm 3 depending on the proportion <strong>of</strong> sedimentary<br />

rocks <strong>and</strong> fresh or altered ultramafic <strong>and</strong> mafic rocks; de<br />

Bremond d’Ars et al., 1999) overlies less dense (ca. 2.7 g/cm 3 ; de<br />

Bremond d’Ars et al., 1999) gneiss ± granitoid. Domes <strong>for</strong>med<br />

due to gravitational instability created by such density differences<br />

are proposed <strong>for</strong> many granite-greenstone terrains<br />

(e.g. Gorman et al., 1978; Condie, 1984; Hickman, 1984; Robin<br />

<strong>and</strong> Bailey, 2009) where they are commonly referred to as<br />

“Raleigh–Taylor diapirs” (e.g. Bouhallier et al., 1995), or simply<br />

“diapirs” 2 (where <strong>for</strong>mation due to density inversion is<br />

implied). Shackleton (1995) contends that diapiric structures<br />

are essential elements <strong>of</strong> greenstone belt evolution. Granitoids<br />

preferentially intruding the core <strong>of</strong> domes may accentuate<br />

gravitational instabilities <strong>and</strong> hence their diapiric emplacement<br />

<strong>for</strong>ming polydiapirs (Weinberg <strong>and</strong> Schmeling, 1992).<br />

Diapirism may occur after thrust stacking <strong>and</strong> nappe tectonics<br />

(e.g. Dirks <strong>and</strong> Jelsma, 1998). Early extensional faults may be<br />

inverted during shortening <strong>and</strong> diapirism (Hippertt <strong>and</strong> Davis,<br />

2000). The interplay between regional shortening, folding, <strong>and</strong><br />

diapirism due to gravitational instabilities is documented by<br />

Park (1982), Bouhallier et al. (1995), Dalstra et al. (1998), Lin<br />

(2005), Parmenter et al. (2006), <strong>and</strong> Erickson (2010) <strong>and</strong> Lana<br />

et al. (2010b) describe development <strong>of</strong> the dome <strong>and</strong> keel structure<br />

<strong>of</strong> the Barberton granitoid-greenstone belt during a 30 m.y.<br />

period <strong>of</strong> crustal extension following crustal shortening. Nappelike<br />

structures, instead <strong>of</strong> simple, symmetrical domes, may also<br />

<strong>for</strong>m due to gravitational instabilities (Ramberg, 1967b), especially<br />

where the initial interface between basement gneisses<br />

<strong>and</strong> overlying greenstones is inclined (c.f. models <strong>of</strong> Talbot,<br />

1974).<br />

(iii) “Vertical tectonics” incorporating both “positive” <strong>and</strong> “negative”<br />

(i.e. descending) diapirs (Bédard et al., 2003; Bédard,<br />

2006), subsiding troughs or “sagduction” (Gorman et al., 1978;<br />

2 Note that the term “diapir” is, however, non-genetic <strong>and</strong> may be used <strong>for</strong><br />

piercement structures without gravitational instablity (c.f. Weinberg <strong>and</strong> Schmeling,<br />

1992), although its usage in Archaean granitoid-greenstone terrains is generally<br />

taken to imply a gravity-driven origin <strong>for</strong> domes.<br />

Goodwin <strong>and</strong> Smith, 1980; Chardon et al., 1996, 1998; François<br />

et al., 2012), or partial convective overturn (Van Kranendonk<br />

et al., 2004; Bodorkos <strong>and</strong> S<strong>and</strong>i<strong>for</strong>d, 2006). Vertical tectonic<br />

models <strong>of</strong> Bédard (2006) differ from simple diapir models in (ii)<br />

in that they include the complex interplay between de<strong>for</strong>mation<br />

<strong>and</strong> petrological phase (<strong>and</strong> hence density) changes, crustal<br />

melting, magmatic processes, <strong>and</strong> mantle delamination.<br />

(iv) Core complex <strong>for</strong>mation during regional extension (e.g.<br />

Williams <strong>and</strong> Whitaker, 1993; Kloppenburg et al., 2001; Zegers<br />

et al., 2001; Kisters et al., 2003; Lana et al., 2010a) that may<br />

follow an early period <strong>of</strong> folding <strong>and</strong> thrusting (e.g. Lobato<br />

et al., 2001). Late granitoids that intrude Archaean granitegreenstone<br />

terrains are interpreted by Kusky (1993) to result<br />

from decompression melting during collapse <strong>and</strong> core complex<br />

<strong>for</strong>mation.<br />

(v) Intrusion. In addition to TTG gneiss domes described above,<br />

late, intrusive granites may also <strong>for</strong>m ovoid, domal features<br />

(Williams <strong>and</strong> Whitaker, 1993; H<strong>of</strong>mann et al., 2003).<br />

The genesis <strong>of</strong> Palaeoproterozoic dome-<strong>and</strong>-keel structures is also<br />

conjectural, with thrust-related, orogenic collapse/core complex,<br />

diapir, <strong>and</strong> modified models combining all <strong>of</strong> them proposed<br />

(Marshak et al., 1997; Tinkham <strong>and</strong> Marshak, 2004). Whilst the<br />

above discussion has concentrated on rocks <strong>of</strong> low to medium<br />

grade, differential loading <strong>and</strong> “gravitational redistribution” is<br />

proposed by Gerya et al. (2000) to explain the <strong>for</strong>mation <strong>of</strong> domes<br />

<strong>of</strong> granulite facies gneiss in high-grade Precambrian terrains.<br />

The origin <strong>of</strong> recumbent folds in Archaean migmatitic gneiss<br />

domes is likewise contentious. Recumbent folds have been interpreted<br />

as <strong>for</strong>ming during vertical gravity-driven tectonics (Gorman<br />

et al., 1978) <strong>and</strong> are developed in Ramberg’s (1967a,b, 1981b) <strong>and</strong><br />

Talbot’s (1974) centrifuge models. Centrifuge models by Harris<br />

et al. (2002) <strong>and</strong> Harris <strong>and</strong> Koyi (2003) show that both recumbent<br />

folds <strong>and</strong> the upright folds that de<strong>for</strong>m them may develop during<br />

regional, layer parallel extension <strong>and</strong> may explain fold <strong>geom</strong>etries<br />

in Archaean terrains interpreted from deep crustal reflection<br />

seismic pr<strong>of</strong>iles (Blewett <strong>and</strong> Czarnota, 2007; Goscombe et al.,<br />

2009; Bédard et al., 2012). Alternatively, recumbent folds may have<br />

<strong>for</strong>med during one or more periods <strong>of</strong> regional thrusting <strong>and</strong> fold<br />

nappe emplacement prior to dome <strong>for</strong>mation, as proposed <strong>for</strong> the<br />

Pilbara Craton <strong>of</strong> Western Australia (Bickle et al., 1980; White et al.,<br />

1998; Van Kranendonk et al., 2004) or <strong>for</strong> Greenl<strong>and</strong> (Windley <strong>and</strong><br />

Garde, 2009).<br />

1.3. Channel flow, <strong>and</strong> similar tectonic processes<br />

Channel flow in an orogen is referred to here as the lateral<br />

flow <strong>of</strong> a weak, viscous crustal layer between relatively rigid yet<br />

de<strong>for</strong>mable bounding crustal slabs due to a horizontal gradient in<br />

lithostatic pressure created by differences in crustal thicknesses<br />

beneath the hinterl<strong>and</strong> compared to the <strong>for</strong>el<strong>and</strong>, <strong>and</strong>/or by erosion<br />

<strong>and</strong> focused denudation (as summarized by Godin et al., 2006<br />

<strong>and</strong> Grujic, 2006). In this process the mid-crust can be extruded<br />

toward the surface within a channel bounded by an upper normalsense<br />

boundary <strong>and</strong> a lower thrust-sense boundary. Channel flow<br />

<strong>and</strong> the similar process <strong>of</strong> “laminar flow” (Dewei, 2008) have been<br />

applied to explain first-order de<strong>for</strong>mation features in large or wide<br />

hot orogens such as the Himalaya (Beaumont et al., 2001, 2004,<br />

2006; Grujic et al., 2002; Jamieson et al., 2004; Burbank, 2005; Jones<br />

et al., 2006; Harris, 2007), the eastern Variscan belt (Schulmann<br />

et al., 2005, 2008; Dörr <strong>and</strong> Zulauf, 2010), the Ediacaran Petterman<br />

orogeny in Australia (Raimondo et al., 2009), the Southern British<br />

Colombia Cordillera (Brown <strong>and</strong> Gibson, 2006), the Appalachian<br />

Inner Piedmont (Hatcher <strong>and</strong> Merschat, 2006), <strong>and</strong> the Proterozoic<br />

Central Mozambique Belt in Tanzania/Southern Kenya (Fritz et al.,<br />

2009). V<strong>and</strong>erhaeghe (2009, Fig. 7b) illustrates contemporaneous


channel flow, granite diapirism, <strong>and</strong> magmatic intrusion within a<br />

Phanerozoic orogen. Additional examples <strong>and</strong> references are given<br />

in reviews by Godin et al. (2006) <strong>and</strong> Grujic (2006). A similar<br />

process to channel flow, involving flow <strong>of</strong> a ductile layer with constrained<br />

lateral boundaries beneath a brittle-ductile to brittle cover<br />

sequence, is termed protrusion tectonics (Leonov, 1994, 2008).<br />

Channel flow (Cagnard et al., 2006; Parmenter et al., 2006), gravity<br />

driven continental overflow tectonics (Bailey, 1999), <strong>and</strong> general<br />

deep crustal flow (Culshaw et al., 2006a) have also been proposed<br />

<strong>for</strong> the Archaean. Similarly, Chardon et al. (2011) propose that the<br />

structure <strong>of</strong> Archaean terrains is controlled by pervasive, threedimensional<br />

flow <strong>of</strong> the lower crust incorporating orogen-normal<br />

shortening, lateral constrictional stretching, <strong>and</strong> transtension. As<br />

thermal gradients were higher in the Archaean <strong>and</strong> hence the<br />

crust more ductile (Richter, 1985; Rey <strong>and</strong> Coltice, 2008; Flament<br />

et al., 2008; further enhanced through heating by mantle plumes;<br />

Choukroune et al., 1995), channel flow is likely to have then been<br />

an important tectonic process (Rey <strong>and</strong> Houseman, 2006). Bailey<br />

(1999) contends that in the Archaean a ductile layer would have<br />

been universally present beneath an upper, brittle crust, <strong>and</strong> that<br />

the ductile layer may flow <strong>and</strong> extrude laterally onto adjacent<br />

ocean basins due to lateral differences in crustal thickness. Lateral<br />

protrusion has been applied to explain mapped Archaean <strong>and</strong><br />

Palaeoproterozoic structures in the Baltic Shield by Kolodyazhny<br />

(2007) <strong>and</strong> Leonov (2008).<br />

1.4. Limitations <strong>of</strong> previous 1 g physical modelling <strong>of</strong> channel flow<br />

in simulating structures <strong>for</strong>med in granite-greenstone terrains<br />

Previous physical models <strong>of</strong> channel flow <strong>and</strong> analogous ductile<br />

flow processes in orogens (e.g. Miller, 1982; Chattopadhyay<br />

<strong>and</strong> M<strong>and</strong>al, 2002; Mukherjee <strong>and</strong> Koyi, 2010; Duretz et al., 2011;<br />

Mukherjee et al., 2012) de<strong>for</strong>med under normal gravity (1g) do<br />

not take into account the interaction between active folding <strong>of</strong> a<br />

cover sequence <strong>of</strong> mechanically anisotropic layers during flow <strong>of</strong><br />

the ductile substrate (e.g. Yakymchuk et al., 2012). Miller’s (1982)<br />

<strong>and</strong> Mukherjee <strong>and</strong> Koyi’s (2010) models simulate only structures<br />

developed in the zone <strong>of</strong> extrusive ductile flow where, apart from at<br />

the extrusive front, the boundaries are artificially fixed <strong>and</strong> planar.<br />

The models <strong>of</strong> Duretz et al. (2011) use a basal indentor to extrude<br />

the lower ductile layers instead <strong>of</strong> differences in gravitational loading<br />

<strong>and</strong>, as their upper crustal layers are simulated by s<strong>and</strong>, do<br />

not simulate active folding in a cover sequence. Whilst producing<br />

some <strong>of</strong> the features on granitoid greenstone terrains, physical<br />

1g models <strong>of</strong> orogen-parallel flow by Cagnard et al. (2006) during<br />

bulk shortening do not produce the observed differences in folds<br />

within granite-gneiss domes in comparison to an overlying greenstone<br />

sequence. A high-acceleration centrifuge is there<strong>for</strong>e used to<br />

simulate the <strong>for</strong>ce <strong>of</strong> gravity to achieve dynamic scaling <strong>of</strong> physical<br />

models incorporating modelling clays <strong>and</strong> silicone “bouncing putties”<br />

(Hubbert, 1937; Ramberg, 1967a,b, 1970, 1981a,b; Dixon <strong>and</strong><br />

Summers, 1985). The model presented herein aims to (1) simulate<br />

active folding <strong>of</strong> a cover sequence <strong>and</strong> ductile flow <strong>of</strong> basal layers<br />

induced by differential loading <strong>and</strong> (2) test a <strong>new</strong> hypothesis <strong>for</strong><br />

the <strong>for</strong>mation <strong>of</strong> dome <strong>and</strong> keel structures in Archaean terrains.<br />

2. Centrifuge modelling <strong>of</strong> channel flow<br />

2.1. Modelling procedure<br />

An upper crustal layered greenstone sequence <strong>of</strong> sedimentary<br />

<strong>and</strong> volcanic rocks (the “superstructure”, following the terminology<br />

<strong>of</strong> Wegmann (1935) <strong>and</strong> Culshaw et al., 2006b) is simulated<br />

using microlaminates <strong>of</strong> modelling clays, mixes <strong>of</strong> modelling clays<br />

<strong>and</strong> silicone, <strong>and</strong> silicone putty with an upper layer <strong>of</strong> low-density<br />

L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154 141<br />

modelling clay (Fig. 1). Silicone “bouncing putties” <strong>of</strong> different<br />

effective viscosities <strong>and</strong> densities, whose bulk density is slightly<br />

less than the superstructure (1.14 g/cm 3 ), are used to simulate<br />

basal felsic ± migmatitic gneisses comprising the lower ductile<br />

sequence or infrastructure. Polydimethylsiloxane (PDMS), a clear,<br />

low density (1.05 g/cm 3 ) <strong>and</strong> low viscosity polymer (Weijermars,<br />

1986; Boutelier et al., 2008) simulating granitoid melt is present<br />

along the superstructure-infrastructure contact over half the width<br />

<strong>of</strong> the model. The rheological properties <strong>and</strong> scaling <strong>of</strong> modelling<br />

materials used in the model presented herein, which are part<br />

<strong>of</strong> a larger program <strong>of</strong> centrifuge channel flow experiments, are<br />

addressed by Poulin (2006), Godin et al. (2011), Harris et al.<br />

(2012), <strong>and</strong> Yakymchuk et al. (2012) so are not repeated herein.<br />

Tomodensitometry (“CT scanning”) techniques <strong>and</strong> applications<br />

to modelling using silicone <strong>and</strong> modelling clay materials are<br />

described by Poulin (2006). The model presented (Fig. 1) is similar<br />

to models presented in Godin et al. (2011), Harris et al. (2008, 2009,<br />

2010), Harris et al. (2012), <strong>and</strong> Yakymchuk et al. (2012), however<br />

it differs in that the cover sequence simulating greenstones is<br />

made slightly denser than underlying layers simulating basement<br />

gneisses by incorporating a larger number <strong>of</strong> competent <strong>and</strong> denser<br />

modelling clay layers (simulating mafic volcanics in a greenstone<br />

sequence).<br />

The model (shown schematically in Fig. 1a) was first progressively<br />

shortened via a collapsing wedge whilst undergoing an<br />

acceleration <strong>of</strong> ca. 950g. Material was then cut from the end <strong>of</strong><br />

the model furthest from the shortening ram down into the infrastructure<br />

(Fig. 1b) in several stages to simulate focused erosion. The<br />

model is then allowed to isostatically equilibrate when returned<br />

to the centrifuge at ca. 950g but without imposing bulk shortening<br />

or extension. CT scans after each increment <strong>of</strong> de<strong>for</strong>mation<br />

enabled the progressive development <strong>of</strong> structures to be viewed<br />

along the same cross-section (Figs. 2 <strong>and</strong> 3), as orthogonal sections<br />

(Fig. 4), <strong>and</strong> reconstructed in 3D (Figs. 5 <strong>and</strong> 6; “fly through” animations<br />

after initial shortening <strong>and</strong> <strong>of</strong> the final model are provided<br />

as supplementary data files 1 <strong>and</strong> 2). The final model was chilled,<br />

sliced, <strong>and</strong> photographed (Fig. 7).<br />

2.2. Folding during layer-parallel shortening<br />

Line drawings based on CT scans <strong>for</strong> the same slice through<br />

the model at different increments <strong>of</strong> de<strong>for</strong>mation (Fig. 3a–c) <strong>and</strong><br />

3D reconstructions (Fig. 5) show that tight to open, chevron to<br />

rounded, upright folds develop in the upper sequence simulating<br />

greenstones during layer-parallel shortening. Parasitic folds are<br />

present in several anti<strong>for</strong>ms <strong>and</strong> syn<strong>for</strong>ms (Fig. 3a–c). Folds are<br />

cylindrical to slightly conical <strong>and</strong> fold hinges are slightly curved.<br />

The fold style in basal silicone layers varies with distance from the<br />

ram used to shorten the model. Adjacent to the ram, initially open<br />

to close overturned folds become isoclinal <strong>and</strong> recumbent <strong>and</strong> are<br />

then refolded by open upright folds with progressive shortening<br />

(Fig. 3a–c). Silicone layers in the center <strong>of</strong> the model are initially<br />

only slightly de<strong>for</strong>med, but small folds overturned in the direction<br />

<strong>of</strong> shortening develop at higher strains. Toward the end <strong>of</strong> the<br />

model furthest from the ram folds in basal silicone layers in the core<br />

<strong>of</strong> an anti<strong>for</strong>m in the upper layered sequence are tight <strong>and</strong> upright.<br />

When present, the PDMS infills the cores <strong>of</strong> anti<strong>for</strong>ms <strong>and</strong> creates<br />

a planar interface with the underlying silicone layers (Fig. 2i–k)<br />

where in the PDMS-absent half <strong>of</strong> the model, the basal silicone layers<br />

infill anti<strong>for</strong>m cores (Fig. 2b–d). Folds <strong>of</strong> larger wavelength are<br />

developed over the part <strong>of</strong> the model where PDMS décollement is<br />

present (Figs. 2 <strong>and</strong> 5). Air bubbles caught in the models (black spots<br />

in Figs. 2 <strong>and</strong> 5) are found in some anti<strong>for</strong>mal cores <strong>and</strong> may have<br />

assisted nucleation <strong>of</strong> some <strong>of</strong> these structures, or simply migrated<br />

toward fold hinges.


142 L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154<br />

Fig. 1. Centrifuge model. (a) Schematic diagram showing the model be<strong>for</strong>e shortening. Layers are assembled upon a curved base that matches the outer curvature <strong>of</strong>the<br />

centrifuge rotor. A wedge <strong>of</strong> modelling clay collapses during centrifugation, pushing a nylon plate to shorten the model parallel to initial layering, as illustrated in (b). (b)<br />

Schematic diagram <strong>of</strong> model after initial shortening. Material is incrementally cut from the end <strong>of</strong> the model furthest from the nylon plate to simulate erosion <strong>and</strong> the<br />

model is placed in the centrifuge to re-equilibrate. (c) Model composition. DMC = Demco ® modelling clay, DC 3179 = Dow Corning ® dilatant compound 3179 silicone putty,<br />

CMM = Crayola ® model magic modelling clay, a low density modelling material, DMC-60 = mix <strong>of</strong> 60% DMC <strong>and</strong> 40% CMM, Plasticine = Harbutts ® oil-based modelling clay,<br />

PDMS = polydimethylsiloxane, a low density <strong>and</strong> viscosity unfilled silicone, CATP = Crazy Aaron Enterprises ® “Thinking Putties” (silicone putties similar to Rhodorsil Gomme<br />

silicone putty, <strong>of</strong> lower viscosity in comparison to DC 3179). Physical <strong>and</strong> X-ray computed tomography (CT) scanning properties <strong>of</strong> materials are discussed by Poulin (2006),<br />

Godin et al. (2011), <strong>and</strong> Yakymchuk et al. (2012). Strontium europium-aluminate powder = “Glow in the dark” powder from Glow Inc. ® is added to enhance some layers on<br />

CT scans (Poulin, 2006).<br />

2.3. Collapse <strong>of</strong> the superstructure during channel flow<br />

When material is removed to simulate erosion (as shown<br />

schematically in Fig. 1b) lateral flow, extrusion, <strong>and</strong> isoclinal folding<br />

<strong>of</strong> basal silicone layers occur beneath the area where material is<br />

removed in a similar manner as proposed <strong>for</strong> extrusive channel flow<br />

(e.g. Godin et al., 2006). The crests <strong>of</strong> anti<strong>for</strong>ms in upper microlaminate<br />

layers simulating greenstones collapse (Figs. 2, 3d, <strong>and</strong> 4a–d),<br />

<strong>for</strong>ming broader flat- to open-crested, doubly plunging structures.<br />

Minor parasitic folds that were developed during initial layerparallel<br />

shortening are flattened <strong>and</strong> unfolded. The tight syn<strong>for</strong>ms<br />

remain relatively unchanged. The upright anti<strong>for</strong>m in the basal<br />

silicone layers closest to the ram is similarly flattened, while recumbent<br />

isoclinal folds within it are accentuated. Silicone layers in the<br />

central part <strong>of</strong> the models are folded by <strong>for</strong>el<strong>and</strong>-verging to overturned<br />

folds (Figs. 3d <strong>and</strong> 4). 3D views (Fig. 6) show that hinges <strong>of</strong><br />

tight syn<strong>for</strong>ms are curved. Orthogonal slices (Fig. 4) <strong>and</strong> details <strong>of</strong><br />

3D reconstructions (Fig. 6e–g) show that local large curvature <strong>of</strong><br />

(sometimes doubly plunging) fold hinges produces elongate domal<br />

structures cored by basal silicone layers, which are de<strong>for</strong>med by<br />

isoclinal recumbent folds that <strong>for</strong>m sheath-like structures.<br />

Oblique photographs <strong>of</strong> the final model are shown in Fig. 7. Thin<br />

layers <strong>of</strong> modelling clay <strong>and</strong> modelling clay-silicone mixes indicate<br />

that parasitic folds are best preserved in syn<strong>for</strong>ms in comparison<br />

to anti<strong>for</strong>ms <strong>and</strong> parasitic folds are not present in lowermost<br />

plasticine layers over the open anti<strong>for</strong>ms. PDMS (that simulates<br />

granitoid in nature) has migrated both laterally <strong>and</strong> longitudinally<br />

to infill crests <strong>of</strong> anti<strong>for</strong>ms (Fig. 7b <strong>and</strong> c). In part <strong>of</strong> the model<br />

(Fig. 7c) the microlaminate superstructural sequence is folded by<br />

overturned to recumbent folds on the flank <strong>of</strong> a dome such that it<br />

is overhung by both silicone <strong>and</strong> PDMS (= granitoid <strong>and</strong> gneiss in<br />

nature). It there<strong>for</strong>e can be envisaged that in nature, if erosion were<br />

to expose a granite-greenstone terrain at a level such as shown in<br />

Fig. 7b <strong>and</strong> c by a dashed green line, granitoid or migmatite domes<br />

flanked by, or with screens <strong>of</strong> complexly folded gneiss, between<br />

which a tightly folded greenstone sequence would outcrop.<br />

3. Discussion<br />

3.1. Folding <strong>and</strong> unfolding<br />

In our centrifuge model fold hinges in the lowermost superstructure<br />

<strong>and</strong> infrastructure that were only slightly curved after<br />

initial layer-parallel shortening are reoriented to become more<br />

curved during channel flow. In contrast, uppermost layers in<br />

the superstructure are little affected during the channel flow<br />

stage <strong>of</strong> the model resulting in marked variations <strong>of</strong> fold orientations<br />

with depth (e.g. Fig. 5). The broad, open anti<strong>for</strong>ms<br />

with few parasitic folds cored by PDMS <strong>and</strong> underlying silicones<br />

<strong>and</strong> tight syn<strong>for</strong>ms with parasitic folds in final stages <strong>of</strong> models<br />

(Figs. 2–4, 6, <strong>and</strong> 7) resemble cross-sections through many<br />

Archaean granite-greenstone terrains. Broad anti<strong>for</strong>ms observed<br />

on orthogonal sections (Fig. 4) indicate the presence <strong>of</strong> doubly<br />

plunging to domal structures. Our model implies that crests <strong>of</strong><br />

domes extended during the collapse stage, <strong>and</strong> that early parasitic<br />

folds may have been unfolded similar to the process <strong>of</strong> unfolding<br />

during “opening out” <strong>of</strong> early folds during superposed de<strong>for</strong>mation<br />

described by Sengupta et al. (2005). Unfolding (as reviewed by<br />

Harris et al., 2012) is also required or important in hinge migration<br />

during folding (Mercier et al., 2007), kink propagation (Price


L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154 143<br />

Fig. 2. CT scans showing slices at two positions in the model (a) <strong>and</strong> (h) during layer parallel shortening <strong>of</strong> 19, 31, <strong>and</strong> 39% respectively (left side <strong>of</strong> figure) <strong>and</strong> during<br />

collapse during lateral ductile flow <strong>of</strong> ductile basal layers toward the end <strong>of</strong> the model where material is removed to simulate focused erosion (right side <strong>of</strong> figure; indicated<br />

by horizontal line with double arrows). Note: (i) Differences in fold <strong>geom</strong>etry between upper layered sequence or superstructure <strong>and</strong> the basal silicone layers <strong>of</strong> the<br />

infrastructure, (ii) changes in the <strong>geom</strong>etry <strong>of</strong> structures in basal silicone layers from refolded recumbent folds near the advancing end wall (right side <strong>of</strong> slices in (b)–(d) <strong>and</strong><br />

(i)–(k)), <strong>and</strong> (iii) opening out <strong>and</strong> destruction <strong>of</strong> minor folds in anti<strong>for</strong>ms <strong>and</strong> preservation <strong>of</strong> tight syn<strong>for</strong>ms during flow <strong>of</strong> basal ductile layers in (e)–(g) <strong>and</strong> (l)–(n). Initial<br />

model height = 15 mm; final height = 22 mm.<br />

<strong>and</strong> Cosgrove, 1990), rotation <strong>of</strong> a folded layer into the extensional<br />

field (Ramsay et al., 1987), ductile de<strong>for</strong>mation in shear<br />

zones (Carreras et al., 2005), <strong>and</strong> during thrust duplex development<br />

(Boyer <strong>and</strong> Elliott, 1982). It is there<strong>for</strong>e likely, as discussed in detail<br />

by Harris et al. (2012), that local unfolding may also occur in nature<br />

in the context <strong>of</strong> layer-parallel extension during channel-flow<br />

induced collapse. This provides a <strong>new</strong> <strong>mechanism</strong> <strong>for</strong> the development<br />

<strong>of</strong> broad anti<strong>for</strong>ms <strong>and</strong> tight syn<strong>for</strong>ms in granite-greenstone<br />

terrains.<br />

The development <strong>of</strong> recumbent isoclinal folds close to the<br />

advancing ram, overturned folds in the center <strong>of</strong> models, <strong>and</strong><br />

upright folds furthest from the ram in basal silicone layers during<br />

upright folding <strong>of</strong> microlaminates was also produced in a 1g<br />

physical model <strong>of</strong> Bucher (1956; Figs. 2–4) where layers <strong>of</strong> different<br />

stitching waxes inter-layered with grease were shortened. In<br />

Bucher’s (1956) experiment, layers closest to the advancing wall<br />

were warmer, <strong>and</strong> hence more ductile, than layers further away<br />

from the advancing wall. As no lateral changes in viscosity were


144 L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154<br />

Fig. 3. Line drawing <strong>of</strong> changes in fold <strong>geom</strong>etry <strong>for</strong> a central slice through the model based on CT scans during layer-parallel shortening (a–c) <strong>of</strong> 19, 31, <strong>and</strong> 39% respectively<br />

<strong>and</strong> isostatic adjustments <strong>and</strong> ductile flow <strong>of</strong> basal silicone layers toward the end <strong>of</strong> the model where material was removed to simulate focused erosion in (d). The final<br />

height <strong>of</strong> the model in (d) is 22 mm whereas its initial height was 15 mm.<br />

Fig. 4. Orthogonal slices through the model after two stages <strong>of</strong> erosion-induced flow following layer-parallel shortening, illustrating closures in both sections across the<br />

central anti<strong>for</strong>m due to curvature <strong>of</strong> tight syn<strong>for</strong>ms. Left: sections (a) <strong>and</strong> (c) illustrate successive stages <strong>of</strong> slices parallel to the flow direction; right: sections (b) <strong>and</strong> (d)<br />

slices orthogonal to the flow direction in basal layers (locations indicated on a <strong>and</strong> c). Note that a non-cylindrical, sheath-like <strong>geom</strong>etry <strong>of</strong> folds in basal silicone layers is<br />

indicated from the eye-shapes <strong>and</strong> the opposed vergence <strong>of</strong> structures in sections perpendicular to the flow direction. Model height = 22 mm.


L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154 145<br />

Fig. 5. 3D reconstructions from CT scans showing progressive stages in shortening <strong>of</strong> the model with <strong>and</strong> without an upper layer (blue). Model width=8cm.(a)Unde<strong>for</strong>med<br />

model. (b–c) 19% shortening, (d–e) 31% shortening, (f–h) Two views after 39% shortening; the face viewed in (h) <strong>and</strong> (i) contains no PDMS, whereas the face in the other<br />

images is the side with PDMS. Note that folds show straight to slightly curved axes at the final stage <strong>of</strong> layer-parallel shortening. Folds are more open <strong>and</strong> have a larger<br />

dominant wavelength above the PDMS layer. (For interpretation <strong>of</strong> the references to colour in this figure legend, the reader is referred to the web version <strong>of</strong> the article.)<br />

present in our model, proximity to the ram is there<strong>for</strong>e more important<br />

in explaining the changes in fold style in basal ductile layers<br />

rather than slight lateral differences in viscosity. Folds produced in<br />

our model <strong>of</strong> channel flow also resemble those portrayed by Leonov<br />

(2008; Fig. 8c), which he interprets as having <strong>for</strong>med during extrusive<br />

ductile flow <strong>of</strong> underlying ductile crust (lateral protrusion).<br />

3.2. Comparison with diapir <strong>and</strong> vertical tectonic models<br />

In contrast to classical domes resulting from gravitational/<br />

Rayleigh–Taylor instabilities (Wilcock, 1991) in granite-greenstone<br />

terrains, as in the centrifuge models <strong>of</strong> Ramberg (1967a,b, 1981a,b;<br />

see discussion in Section 1.2 (ii)), the basal silicone layers in our<br />

model have flowed to infill the cores <strong>of</strong> anti<strong>for</strong>ms developed due to<br />

active folding <strong>and</strong> have subsequently undergone ductile flow parallel<br />

to the base <strong>of</strong> the model. No density driven upwelling into<br />

the upper sequence <strong>of</strong> microlaminates has taken place <strong>and</strong> there<br />

is no necking at depth <strong>of</strong> ductile basal layers as in diapir models.<br />

In our model, PDMS flowed toward anti<strong>for</strong>mal hinges due to its<br />

lesser density as granite melt would in both our proposed channel<br />

flow <strong>and</strong> diapir models in nature. Sinking <strong>of</strong> synclines <strong>and</strong> their<br />

refolding is similar to that proposed in some vertical tectonic <strong>and</strong><br />

incipient sagduction models (e.g. Bédard et al., 2003 <strong>and</strong> Chardon<br />

et al., 1996, respectively). In our model, vertical movements modify<br />

<strong>and</strong> accentuate the <strong>geom</strong>etry <strong>of</strong> folds <strong>for</strong>med during layer-parallel<br />

shortening. Vertical movements do not, however, initiate the <strong>for</strong>mation<br />

<strong>of</strong> folds as <strong>for</strong> diapir models.<br />

3.3. Implications <strong>for</strong> Archaean tectonics<br />

Structural styles <strong>and</strong> tectonic <strong>mechanism</strong>s may have been<br />

different in the Archaean in comparison to younger terrains<br />

(Marshak, 1999; Bédard et al., 2012) due to a higher geotherm,<br />

enhanced radiogenic heating, <strong>and</strong> lighter sub-crustal lithospheric<br />

mantle that favored gravitationally-driven tectonics <strong>and</strong> ductile<br />

flow (Rey <strong>and</strong> Houseman, 2006; Rey <strong>and</strong> Coltice, 2008). The<br />

necessity <strong>for</strong> an initial period <strong>of</strong> layer-parallel shortening <strong>of</strong> the<br />

model is supported by evidence <strong>for</strong> regional horizontal shortening<br />

in many Archaean terrains (see reviews by Cawood et al., 2006<br />

<strong>and</strong> Bédard et al., 2012). The observation in our model <strong>of</strong> collapse<br />

<strong>of</strong> the thickened superstructure during lateral flow <strong>of</strong> basal ductile<br />

layers is consistent with results <strong>of</strong> numerical simulations that<br />

suggest that thickened Archaean crust is likely to spread <strong>and</strong> flow<br />

under gravity <strong>and</strong> that thrusting typical <strong>of</strong> Phanerozoic orogens is<br />

inhibited (Rey <strong>and</strong> Houseman, 2006). These authors suggest that<br />

Archaean lithosphere may have de<strong>for</strong>med in a similar manner<br />

to that <strong>of</strong> modern <strong>and</strong> thermally mature “hot” orogens such as<br />

the Himalaya where crustal gravitational <strong>for</strong>ces play a significant


146 L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154<br />

Fig. 6. 3D reconstructions from CT scans showing modification <strong>of</strong> structures during two stages (a–b) <strong>and</strong> (c–i) <strong>of</strong> ductile flow <strong>of</strong> lower silicone layers with <strong>and</strong> without upper<br />

layer (blue). Model width=8cm.(e–i) illustrate enlargements <strong>of</strong> areas in the model illustrating curvature <strong>of</strong> fold axes in lowermost marker layers in the superstructure <strong>and</strong><br />

isoclinal recumbent folds in the infrastructure. In contrast, uppermost layers (blue) show little change during flow <strong>of</strong> basal silicone layers. (For interpretation <strong>of</strong> the references<br />

to colour in this figure legend, the reader is referred to the web version <strong>of</strong> the article.)<br />

role <strong>and</strong> channel flow is postulated to be an important process<br />

(see Section 1.2). Indeed, other centrifuge models incorporating<br />

simulated channel flow scaled to model de<strong>for</strong>mation in the<br />

Nepal Himalaya produce structures, such as hinterl<strong>and</strong>-verging<br />

folds, comparable to those observed in the field (Godin et al.,<br />

2011).<br />

Analysis <strong>of</strong> regional gravity data by Peshler et al. (2004) shows<br />

that granitoid-gneiss domes from a range <strong>of</strong> post-2.8 Ga terrains do<br />

not have the common mushroom-shaped <strong>for</strong>m <strong>and</strong> deep root zone<br />

predicted by centrifuge models <strong>of</strong> diapirs (e.g. Ramberg, 1967a,b,<br />

1981a,b; Talbot, 1974; Dixon, 1975), whereas the <strong>geom</strong>etry <strong>of</strong><br />

pre-2.8 Ga batholiths <strong>and</strong> greenstones are consistent with diapiric<br />

models. Greenstone keels (that may extend deeper than adjacent<br />

batholiths) in post-2.8 Ga terrains are interpreted by Peshler et al.<br />

(2004) to result from folding during horizontal shortening <strong>and</strong> not<br />

diapirism. Their pr<strong>of</strong>iles across pre-2.8 Ga terrains show <strong>geom</strong>etries<br />

similar to those <strong>for</strong> an equivalent erosion level in our model (Fig. 7b<br />

<strong>and</strong> c). An efficient, aggressive weathering system is proposed <strong>for</strong><br />

the Archaean (Lowe <strong>and</strong> Tice, 2004; Hessler <strong>and</strong> Lowe, 2006) due<br />

in part to the absence <strong>of</strong> plants, the likelihood <strong>for</strong> strong chemical<br />

erosion due to acid rain, <strong>and</strong> the presence <strong>of</strong> an anomalously buoyant<br />

sub-continental mantle lithosphere, which would enhance<br />

exhumation (Groves et al., 2006). For example, high erosion rates<br />

are indicated from regional studies <strong>of</strong> the Witswatersr<strong>and</strong> Basin<br />

(Groves et al., 2006; Schoene et al., 2008). However, until ca.<br />

2.8–2.5 Ga, continents were unable to sustain topography >2500 m<br />

(Rey <strong>and</strong> Coltice, 2008) <strong>and</strong> it has also been proposed that, prior<br />

to the Neoarchaean, much <strong>of</strong> the Earth may have been submerged<br />

(Flament et al., 2008), limiting or precluding extensive erosion.<br />

Channel flow models require a more rigid upper crustal greenstone<br />

sequence, which again is reached only in the Neoarchaean (Rey<br />

<strong>and</strong> Coltice, 2008), in part due to erosion <strong>and</strong> removal <strong>of</strong> heatproducing<br />

element-rich upper crust through extensive erosion that<br />

resulted in an increase in lithospheric strength with time (Schoene<br />

et al., 2008). These factors suggest that channel flow may be more<br />

applicable <strong>for</strong> Neoarchaean granitoid-greenstone terrains <strong>and</strong> that<br />

one or more <strong>of</strong> the alternative <strong>mechanism</strong>s described in Section 1,<br />

the most likely being diapirism <strong>and</strong>/or partial convective overturn,<br />

may be more significant <strong>for</strong> older terrains. In the early Earth, even<br />

the felsic substrate may not have been strong enough to enable the<br />

<strong>for</strong>mation <strong>of</strong> dome <strong>and</strong> keel structures (Hamilton, 2007a,b) <strong>and</strong><br />

vertical tectonic models, such as Bédard et al. (2003), may be more<br />

applicable.<br />

Classical domes, sub-circular in map view, were not fully developed<br />

in our model. In Archaean terrains, however, our modelling<br />

suggests that increased curvature <strong>of</strong> syn<strong>for</strong>mal fold hinges <strong>and</strong><br />

the <strong>for</strong>mation <strong>of</strong> domal structures during channel flow may be<br />

enhanced by:


L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154 147<br />

Fig. 7. Photographs <strong>of</strong> three slices through the model at the end <strong>of</strong> de<strong>for</strong>mation. Parasitic folds are best preserved in syn<strong>for</strong>mal areas in upper layered sequence (superstructure).<br />

If the model were sliced horizontally along the pale green dashed line representing a possible erosion level, then anti<strong>for</strong>ms <strong>and</strong> domes cored by clear PDMS with screens<br />

<strong>of</strong> yellow underlying silicone would be similar to granite <strong>and</strong> migmatitic gneisses in nature. Note that boudinage <strong>of</strong> denser, competent (blue) plasticine layers resulted from<br />

imperfections in model construction. Model height = 22 mm. (For interpretation <strong>of</strong> the references to colour in this figure legend, the reader is referred to the web version <strong>of</strong><br />

the article.)<br />

(i) The presence <strong>of</strong> or lateral changes in the thickness <strong>of</strong> a décollement<br />

horizon between greenstones <strong>and</strong> gneissic basement (c.f.<br />

models presented herein <strong>and</strong> by Yakymchuk et al., 2012).<br />

(ii) Lateral differences in dip <strong>of</strong> erosion fronts, or variable amounts<br />

<strong>of</strong> erosion that bring about ductile flow oblique to initial fold<br />

axes (as in similar models by Yakymchuk et al., 2012).<br />

(iii) Erosion fronts at a high angle to the trend <strong>of</strong> the orogen that may<br />

induce (or enhance, cf. Chardon et al., 2009, 2011) orogen parallel<br />

flow, potentially in addition to orogen normal or oblique<br />

flow, if erosion fronts <strong>of</strong> other orientation(s) were also present.<br />

(iv) Irregularities in the <strong>geom</strong>etry <strong>of</strong> a basal indentor, such as a<br />

wedge <strong>of</strong> sub-crustal lithospheric mantle that (as in models <strong>of</strong><br />

Duretz et al., 2011) may enhance channel flow.<br />

(v) Discontinuities in the flow pattern during multiple pulses <strong>of</strong><br />

channel flow (c.f. Hollister <strong>and</strong> Grujic, 2006).<br />

Our modelling results suggest that a greenstone sequence may portray<br />

fold styles that differ greatly from underlying gneisses (e.g.<br />

Figs. 2–4). These differences, although <strong>for</strong>med in a single event,<br />

could easily be misinterpreted to suggest different tectonic histories.<br />

Recumbent folds in basement gneisses or in the core <strong>of</strong><br />

gneiss domes may develop at the same time as upright folds in the<br />

greenstone sequence <strong>and</strong> may not necessarily be associated with<br />

regional thrusting. For example, the folds produced in the “channel”<br />

in our experiments resemble those depicted in Schulmann et al.<br />

(2008, especially their Fig. 10) during interpreted channel flow in<br />

the Moldanubian domain <strong>of</strong> the Bohemian Massif. Based on the<br />

localization <strong>of</strong> structures in our model, <strong>and</strong> that described above by<br />

Bucher (1956), early recumbent folds may not be expected across<br />

an entire granite-greenstone province but may better develop near<br />

the source <strong>of</strong> shortening, although during channel flow recumbent<br />

folds are produced in the entire channel. During channel flow,<br />

regional variations in the “ponding” <strong>of</strong> granitoid magma beneath<br />

a greenstone sequence (simulated by PDMS in our model) or in<br />

the orientation <strong>and</strong> extent <strong>of</strong> the erosion front with respect to initial<br />

fold axes in the superstructure may result in fold interference<br />

patterns previously interpreted as requiring superposed regional<br />

shortening events. Further experiments with variable orientations<br />

<strong>of</strong> the erosion front are planned to test this hypothesis. It is likely<br />

that structures previously thought to result from other <strong>mechanism</strong>s<br />

(e.g. superposed folds during two or more periods <strong>of</strong> bulk shortening,<br />

diapirism <strong>and</strong>/or partial convective overturn, extensional<br />

core-complexes, etc.; see Section 1.2) may be better explained by<br />

a combination <strong>of</strong> early folding during regional shortening followed<br />

by collapse during channel flow.<br />

Our model demonstrates that (<strong>of</strong>ten late) extensional shear<br />

zones documented in some Archaean granite-greenstone terrains<br />

(e.g. in Canada, Australia <strong>and</strong> Brazil: Kusky, 1993; Lobato et al.,


148 L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154<br />

Fig. 8. (a) Schematic cross-section illustrating structures in Meso-Neoarchaean crustal blocks <strong>of</strong> West Greenl<strong>and</strong> redrawn after Fig. 15e <strong>of</strong> Windley <strong>and</strong> Garde (2009). Simple<br />

“arches <strong>and</strong> cusps” in granite-greenstones in the upper crust <strong>and</strong> crustally derived granites are separated from complexly refolded folds in lower crustal, granulite facies<br />

gneisses beneath a décollement horizon. (No scale was included in the original figure.) (b) Postulated reconstructed initial configuration based on centrifuge model results.<br />

In this reconstruction, both upper <strong>and</strong> lower levels were initially de<strong>for</strong>med by folds <strong>of</strong> the same wavelength subsequently modified by ductile flow <strong>of</strong> lower crustal rocks<br />

along the décollement horizon interpreted by Windley <strong>and</strong> Garde (2009). (c) Schematic cross-section <strong>of</strong> the Karelian Massif modified after Leonov (2008) illustrating open<br />

anti<strong>for</strong>ms <strong>and</strong> pinched syn<strong>for</strong>ms developed during interpreted ductile flow (termed “lateral protrusion” by Leonov, 2008) <strong>of</strong> underlying gneiss.<br />

2001) may not require regional extension, but that (i) extension<br />

may be confined to the superstructure <strong>and</strong> ductile infrastructure as<br />

a result <strong>of</strong> collapse during channel flow, <strong>and</strong> (ii) the rigid basement<br />

beneath the ductile infrastructure may not necessarily be extended.<br />

The model may be particularly applicable as an alternative to<br />

diapirism in cases where there has been initial thrusting leading<br />

to crustal thickening <strong>and</strong> anomalously hot crust (e.g. Zimbabwe<br />

Craton; Dirks <strong>and</strong> Jelsma, 1998) that would continue to de<strong>for</strong>m in a<br />

ductile manner <strong>for</strong> a sustained period similar to large hot orogens<br />

in the Phanerozoic.<br />

Windley <strong>and</strong> Garde’s (2009, Fig. 15) schematic section<br />

across West Greenl<strong>and</strong> reproduced in Fig. 8a shows granulite<br />

facies orthogneiss <strong>and</strong> gabbro beneath a detachment horizon<br />

folded by upright structures <strong>of</strong> regular wavelength that<br />

refold isoclinal recumbent folds. Above the detachment horizon,<br />

tonalite–trondhjemite–granodiorite (TTG) orthogneisses <strong>and</strong><br />

volcanic rocks display open arches (rounded anti<strong>for</strong>ms) <strong>and</strong> cusps<br />

(tight syn<strong>for</strong>ms) typical <strong>of</strong> granite-greenstone terrains whereas<br />

beneath the inferred décollement in Windley <strong>and</strong> Garde’s (2009)<br />

schematic representation <strong>of</strong> structures, anti<strong>for</strong>ms <strong>and</strong> syn<strong>for</strong>ms<br />

are <strong>of</strong> similar, regular wavelength. Our centrifuge modelling suggests<br />

that structures beneath a décollement horizon may not be<br />

significantly modified during channel flow. Our modelling suggests<br />

that prior to displacement on the décollement horizon in Greenl<strong>and</strong><br />

across which fold <strong>geom</strong>etries change (Fig. 8a), sequences above<br />

<strong>and</strong> beneath the décollement horizon may have been de<strong>for</strong>med by<br />

folds <strong>of</strong> similar wavelength during the period <strong>of</strong> crustal shortening<br />

interpreted by Windley <strong>and</strong> Garde (2009). A suggested, prechannel<br />

flow reconstruction is shown schematically in Fig. 8b.<br />

Our centrifuge model suggests that folds above the décollement<br />

in Greenl<strong>and</strong> may have opened out during channel flow<br />

induced collapse during which ductile flow along the décollement<br />

occurred. It is likely that granitoids were preferentially emplaced<br />

into anti<strong>for</strong>mal closures similar to migration <strong>of</strong> PDMS in our<br />

model.<br />

3.4. Mechanisms <strong>for</strong> regional shortening <strong>and</strong> potential<br />

implications <strong>for</strong> enhanced channel flow in the Neoarchaean<br />

Despite considerable research in Archaean terrains since early<br />

discussions on the applicability <strong>of</strong> plate tectonic models to<br />

the Archaean in Kröner (1981) <strong>and</strong> the recognition <strong>of</strong> extensive<br />

imbrication <strong>and</strong> tectonic juxtaposition <strong>of</strong> terrains in some<br />

granitoid-greenstone <strong>and</strong> granulite-facies gneiss belts (Windley,


L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154 149<br />

Fig. 9. (a) Lithospheric cross-section redrawn from thermo-mechanical numerical model <strong>of</strong> Gray <strong>and</strong> Pysklywec (2010; Fig. 2B) showing structures developed during imposed<br />

shortening. Imbricate structures in the mantle lithosphere produced in models <strong>of</strong> Gray <strong>and</strong> Pysklywec (2010) would appear identical to those interpreted as fossil subduction<br />

zones on deep crustal seismic pr<strong>of</strong>iles across Archaean terrains. However, no subduction nor terrane accretion occurred in the numerical models. Upper crust (green on online<br />

version) = dolerite rheology, lower crust (pink) = felsic granulite rheology, lithospheric mantle = dry olivine rheology <strong>and</strong> lower mantle = wet olivine rheology (see Gray <strong>and</strong><br />

Pysklywec, 2010 <strong>for</strong> details). (b) <strong>and</strong> (c) Postulated evolution <strong>of</strong> an enlarged region indicated in (a) based on our centrifuge model <strong>of</strong> erosion-induced channel flow <strong>and</strong> the<br />

indentation-driven channel flow model <strong>of</strong> Duretz et al. (2011). In a similar manner to physical models <strong>of</strong> Duretz et al. (2011), indentation <strong>of</strong> sub-crustal lithospheric mantle<br />

may assist channel flow leading to exhumation <strong>of</strong> high-grade felsic gneisses <strong>and</strong> development <strong>of</strong> arch <strong>and</strong> cuspate folds in greenstones in the upper crust. Juxtaposition <strong>of</strong><br />

low-grade granite-greenstone <strong>and</strong> high-grade gneiss terrains in (c) would likely produce a similar expression on deep crustal reflection seismic pr<strong>of</strong>iles as those interpreted<br />

as juxtaposition <strong>of</strong> terranes above a fossil subduction zone. These figures suggest that different tectonic interpretations <strong>of</strong> seismic pr<strong>of</strong>iles may be made either supporting<br />

or contradicting Archaean plate tectonic models <strong>for</strong> terrane assembly. (For interpretation <strong>of</strong> the references to colour in this figure legend, the reader is referred to the web<br />

version <strong>of</strong> the article.)<br />

1998) there is still much debate as to the driving <strong>mechanism</strong>s <strong>for</strong><br />

Archaean tectonics <strong>and</strong> the relative importance <strong>of</strong> horizontal versus<br />

vertical motions (e.g. Lin, 2005; Condie et al., 2006; Bédard et al.,<br />

2012). Ramberg (1967b, chapter 17) proposes that mantle traction<br />

on the base <strong>of</strong> the lithosphere may result in areas <strong>of</strong> crustal shortening<br />

<strong>and</strong> extension. Similarly, cratonic mobilism in response to<br />

mantle flow acting on cratonic keels is proposed by Bédard et al.<br />

(2012) to explain the development <strong>of</strong> contractional <strong>and</strong> extensional<br />

structures in Archaean terrains. Bédard et al. (2012) suggest<br />

that there is little evidence <strong>for</strong> Archaean subduction zones <strong>and</strong><br />

hence that “alpine-style” collisional tectonics <strong>and</strong> orogenesis did<br />

not occur in the Archaean.<br />

Although shallowly-dipping reflectors in the upper mantle on<br />

deep crustal reflection seismic pr<strong>of</strong>iles <strong>of</strong> Archaean terrains have<br />

been taken by some authors as evidence <strong>for</strong> “fossilised” subduction<br />

zones <strong>and</strong> interpreted as implying the existence <strong>of</strong> plate tectonics<br />

in the Archaean (e.g. Ludden et al., 1993; Calvert et al., 1995;<br />

Cawood et al., 2006), alternative possibilities exist to explain these<br />

reflectors. For example, similar mantle reflectors in the North Sea<br />

are interpreted as extensional shear zones by Reston (1990). Alternatively,<br />

Robin <strong>and</strong> Bailey (2009) propose that low-angle reflectors<br />

could correspond to density boundaries. A <strong>geom</strong>etry resembling<br />

fossil subduction zones is indeed produced in centrifuge models<br />

<strong>of</strong> structures in orogenic belts by Ramberg (1967b, figs. 98–100)<br />

containing initial density boundaries. Ramberg’s initial models<br />

comprised (i) upper horizontal layers <strong>of</strong> different density <strong>and</strong><br />

viscosity, (ii) a middle layer package bounded laterally by <strong>and</strong><br />

overlying (iii) material <strong>of</strong> uni<strong>for</strong>m <strong>and</strong> greater density. De<strong>for</strong>mation<br />

during centrifugation resulting solely from “rearrangement <strong>of</strong><br />

such unstable mass distributions to a more stable state” (Ramberg,<br />

1967a,b), i.e. without bulk horizontal shortening, produced dipping<br />

shear zones in the uni<strong>for</strong>m basal layer(s). Complex fold nappe<br />

<strong>and</strong>/or domal, “diapiric” 3 (c.f. Dixon, 1975 <strong>and</strong> Ramberg, 1980,<br />

1981a,b) structures were developed in overlying layers depending<br />

on initial model configuration.<br />

Shear zones in the upper mantle <strong>of</strong> equivalent <strong>geom</strong>etry to<br />

those interpreted on seismic pr<strong>of</strong>iles as remnants <strong>of</strong> subduction<br />

zones are also produced in numerical models <strong>of</strong> bulk shortening<br />

<strong>of</strong> Neoarchaean continental lithosphere without subduction (Gray<br />

<strong>and</strong> Pysklywec, 2010). Gray <strong>and</strong> Pysklywec (2010) illustrate folding<br />

<strong>and</strong> thickening <strong>of</strong> layers where the upper crust (defined in their<br />

models by a dolerite rheology) overlies a ductile, felsic granulite<br />

lower crust, simultaneously with imbrication <strong>of</strong> a lithospheric mantle<br />

with a rheology <strong>of</strong> dry olivine (Fig. 9a). Erosion was not included<br />

in these numerical models. Similarly, given the <strong>geom</strong>etry produced<br />

in Gray <strong>and</strong> Pysklywec’s (2010) model shown in Fig. 9a, indentation<br />

<strong>of</strong> the upper mantle into the lower crust may occur in nature during<br />

further bulk horizontal shortening (Fig. 9b). If this were to be<br />

coupled with focused erosion in a natural situation comparable to<br />

the numerical model <strong>of</strong> Gray <strong>and</strong> Pysklywec (2010), then channel<br />

flow may occur (Fig. 9c). In a similar manner to the indentor used<br />

to induce channel flow in models by Duretz et al. (2011), impingement<br />

<strong>of</strong> the wedge <strong>of</strong> sub-crustal lithospheric mantle portrayed<br />

in the numerical model into the base <strong>of</strong> the crust may enhance<br />

channel flow <strong>and</strong> exhumation <strong>of</strong> mid-crustal material (Fig. 9c). Our<br />

centrifuge model suggests that the fold <strong>geom</strong>etry <strong>of</strong> greenstones<br />

may be modified during this enhanced channel flow to produce<br />

open anti<strong>for</strong>ms <strong>and</strong> tight syn<strong>for</strong>ms (Fig. 9c).<br />

3 Note that the term “diapir” was not used by Ramberg (1967a,b) to describe the<br />

structures in his centrifuge models, although they have subsequently been called<br />

diapirs.


150 L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154<br />

Fig. 10. Cross sections through the Penobsquis salt structure <strong>of</strong> the Moncton Basin, New Brunswick, Canada, based on depth-converted seismic reflection pr<strong>of</strong>iles <strong>and</strong> borehole<br />

logs; simplified after Wilson et al. (2006). (a) <strong>and</strong> (b) Folds developed above an evaporite layer that has acted as a ductile décollement (similar to the PDMS layer in half <strong>of</strong><br />

our model). (c) Collapsed <strong>and</strong> vertically flattened anticline, interpreted by Wilson et al. (2006) as being <strong>for</strong>med due to modification <strong>of</strong> an initially taller salt structure (as in a<br />

<strong>and</strong> b) during gravity spreading <strong>of</strong> the underlying salt.<br />

A channel flow model, such as portrayed in Fig. 9c, may also provide<br />

an alternative interpretation <strong>for</strong> the juxtaposition <strong>of</strong> de<strong>for</strong>med<br />

Archaean granite-greenstone terrains <strong>and</strong> high-grade migmatite<br />

<strong>and</strong> granulite facies gneiss terrains <strong>of</strong> the same or similar metamorphic<br />

age attributed to collisional tectonics (Shackleton, 1995)<br />

or plume tectonics (Sharkov <strong>and</strong> Bogatikov, 2010). Instead <strong>of</strong><br />

marking a fossil subduction zone beneath the contact between<br />

granite-greenstone <strong>and</strong> high-grade gneiss terrains, shallowlydipping<br />

reflectors on seismic pr<strong>of</strong>iles may highlight upper mantle<br />

shear zones resembling those produced in numerical models by<br />

Gray <strong>and</strong> Pysklywec (2010).<br />

No unique process may explain the dome <strong>and</strong> keel <strong>geom</strong>etry <strong>of</strong><br />

granite–greenstone terrains. Bodorkos <strong>and</strong> S<strong>and</strong>i<strong>for</strong>d (2006) suggest<br />

that different <strong>mechanism</strong>s may apply depending on the age <strong>of</strong><br />

the crust (reflecting temporal changes in heat-producing elements<br />

<strong>and</strong> crustal rheology with time) <strong>and</strong> the thicknesses <strong>of</strong> greenstone<br />

sequences. Bodorkos <strong>and</strong> S<strong>and</strong>i<strong>for</strong>d (2006) use this to explain<br />

differences in <strong>geom</strong>etries between elongate domes in the Neoarchaean<br />

Eastern Goldfields Province <strong>and</strong> more circular domes (in<br />

map view) in the Mesoarchaean East Pilbara in Western Australia.<br />

Age differences in de<strong>for</strong>mation style <strong>and</strong> <strong>mechanism</strong>s are not, however,<br />

conclusive. Choukroune et al. (1997) note that whilst the<br />

Superior Province in Canada is dominated by elongate belts the<br />

younger Dharwar Craton in India is characterized by diapiric dome<br />

<strong>and</strong> basin features, which they attribute to a major thermal event<br />

such as the impingement <strong>of</strong> mantle plume leading to reheating <strong>of</strong><br />

the lower <strong>and</strong> middle crust enhancing diapirism, a process common<br />

in the late Archaean (Windley, 1998).<br />

Lateral protrusion (a <strong>mechanism</strong> similar to channel flow; see<br />

Section 1.3) is proposed by Leonov (2008) to explain folding<br />

<strong>of</strong> Palaeoproterozoic rocks producing tight syn<strong>for</strong>ms <strong>and</strong> broad<br />

anti<strong>for</strong>ms in the Karelian Massif <strong>of</strong> Baltica (Fig. 8c). This suggests<br />

that our modelling may also be applicable to some Palaeoproterozoic<br />

<strong>and</strong> possibly younger terrains.<br />

3.5. Implications <strong>for</strong> mineral exploration<br />

The sequence <strong>of</strong> development <strong>of</strong> structures proposed from<br />

our centrifuge modelling has implications <strong>for</strong> regional structural<br />

interpretation, localization <strong>of</strong> dilatational sites through “stress<br />

mapping” (e.g. Groves et al., 2000; Mair et al., 2001), <strong>and</strong> fluid<br />

flow modelling (e.g. Drummond et al., 2004) applied to mineral<br />

exploration in Archaean terrains. Gold mineralization in greenstone<br />

belts commonly post-dates the peak <strong>of</strong> regional metamorphism<br />

(Groves et al., 2000). For example, in the Barberton greenstone<br />

belt, a classic dome <strong>and</strong> keel province (Lana et al., 2010b), mineralization<br />

occurred during a long period <strong>of</strong> extensional doming<br />

<strong>and</strong> buoyant rise <strong>of</strong> the basement during orogen-parallel extension<br />

<strong>and</strong> infolding <strong>of</strong> the greenstone sequence between gneiss domes<br />

at least 200 m.y. after peak metamorphism <strong>and</strong> crustal thickening<br />

(Dziggel et al., 2010; Lana et al., 2010a,b). Dziggel et al. (2010)<br />

suggest hydrothermal fluid flow related to mineralization toward


sites <strong>of</strong> reduced mean stress occurred during late-stage doming<br />

<strong>and</strong> extension. Stress mapping to predict sites <strong>of</strong> reduced stress<br />

requires knowledge <strong>of</strong> the far-field stresses during mineralization<br />

(Groves et al., 2000). It is there<strong>for</strong>e important in the <strong>for</strong>mulation<br />

<strong>of</strong> gold exploration strategies to either determine from field criteria<br />

whether <strong>for</strong>mation <strong>of</strong> regional domes occurred due to diapirism<br />

during regional subhorizontal shortening (e.g. Dalstra et al., 1998),<br />

during regional extension, or in a stress field that may locally be<br />

highly oblique to that responsible <strong>for</strong> initial folds, due to local differences<br />

in crustal loading without regional shortening or extension,<br />

as in our model. Field criteria <strong>for</strong> the recognition that unfolding has<br />

occurred are described by Harris et al. (2012).<br />

3.6. Comparison with salt tectonics<br />

Model results also show similarities to structures developed in a<br />

sedimentary sequence in fold <strong>and</strong> thrust belts above a salt layer (as<br />

reviewed by Hudec <strong>and</strong> Jackson, 2007). Upright folds with rounded,<br />

salt cored anticlines similar to that produced in the contractional<br />

stage <strong>of</strong> our models are developed during shortening <strong>of</strong> a sedimentary<br />

sequence (that contains internal décollement horizons) upon<br />

a salt layer are observed in nature <strong>and</strong> produced in numerical <strong>and</strong><br />

s<strong>and</strong>box models (e.g. Colletta et al., 1995; Costa <strong>and</strong> Vendeville,<br />

2002; Rowan et al., 2004; Yamato et al., 2011). Without interlayered<br />

ductile horizons within the cover sequence, or where salt layers are<br />

thin (Rowan et al., 2004), de<strong>for</strong>mation is dominated by thrusting<br />

<strong>and</strong> the development <strong>of</strong> box or kink folds (Cotton <strong>and</strong> Koyi, 2000;<br />

Costa <strong>and</strong> Vendeville, 2002; Sans, 2003; Yamato et al., 2011).<br />

The presence or absence <strong>of</strong> a basal PDMS layer in the model presented<br />

above <strong>and</strong> models in Yakymchuk et al. (2012) is a dominant<br />

factor in developing curved fold axes in overlying microlaminate<br />

layers. Curvature <strong>of</strong> fold axes (<strong>and</strong> associated thrusts) is also controlled<br />

by the <strong>geom</strong>etry <strong>of</strong> a ductile décollement representing salt<br />

(Cotton <strong>and</strong> Koyi, 2000; Luján et al., 2003). In the Santos Basin,<br />

Brazil, Fiduk <strong>and</strong> Rowan (2012) document complex dome-<strong>and</strong>basin<br />

fold <strong>geom</strong>etries developed in the sedimentary cover sequence<br />

during (possibly differential or convergent) flow <strong>of</strong> underlying layered<br />

evaporites in which recumbent <strong>and</strong> sheath folds were <strong>for</strong>med.<br />

These structures <strong>for</strong>med during flow <strong>of</strong> the evaporates <strong>and</strong> resemble<br />

those in the centrifuge model we present above. The thick, basal<br />

evaporite sequence in the Santos Basin would correspond to our<br />

whole basal silicone package (<strong>and</strong> not just the thin PDMS layer<br />

present in part <strong>of</strong> our model).<br />

Similarities in the processes <strong>of</strong> channel flow <strong>and</strong> salt ‘escape’<br />

or ‘extrusion’ in fold <strong>and</strong> thrust belts are suggested by McQuarrie<br />

(2004). Lateral flow <strong>of</strong> salt during salt withdrawal or salt nappe<br />

extrusion in the distal part <strong>of</strong> a sedimentary basin results in a broad<br />

zone <strong>of</strong> extension <strong>of</strong> a proximal sedimentary sequence (e.g. Rowan<br />

et al., 2004). In sedimentary basins <strong>and</strong> their simulation in s<strong>and</strong>box<br />

(e.g. Adam <strong>and</strong> Krézsek, 2012) <strong>and</strong> numerical (e.g. Ings et al.,<br />

2004) models, the layers above the salt generally de<strong>for</strong>m in a brittle<br />

manner <strong>and</strong> normal faults develop. Folds are developed or modified<br />

due to rollover on faults <strong>and</strong> not through active folding, as in<br />

our models. Salt (or its model analogue) upwelling occurs in normal<br />

fault-controlled syn<strong>for</strong>ms, a process we do not see in the centrifuge<br />

model described herein nor in models <strong>of</strong> unfolding during collapse<br />

<strong>of</strong> the superstructure during channel flow in Harris et al. (2012).An<br />

example that does, however, resemble our model is the Penobsquis<br />

salt structure in the Moncton Basin <strong>of</strong> eastern Canada described<br />

by Wilson et al. (2006) as the cover sequence above an evaporite<br />

layer de<strong>for</strong>med in a ductile manner. The Penobsquis salt structure<br />

comprises elongate, asymmetric domes (Wilson et al., 2006). Cross<br />

sections based on depth-converted seismic reflection pr<strong>of</strong>iles <strong>and</strong><br />

borehole data in Fig. 10 are interpreted by Wilson et al. (2006) as<br />

illustrating (i) <strong>for</strong>mation <strong>of</strong> anticlines during crustal shortening, followed<br />

by (ii) collapse <strong>and</strong> vertical flattening <strong>of</strong> an initially taller<br />

L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154 151<br />

salt structure during gravity spreading <strong>of</strong> the underlying salt, to<br />

produce the structure illustrated in Fig. 10c. Flattening <strong>and</strong> opening<br />

out <strong>of</strong> anti<strong>for</strong>ms <strong>and</strong> <strong>for</strong>mation <strong>of</strong> local tight, overturned folds<br />

in Fig. 10c is mechanically comparable to the process we envisage<br />

<strong>for</strong> the <strong>for</strong>mation <strong>of</strong> some Neoarchaean dome <strong>and</strong> keel structures<br />

based on our centrifuge model.<br />

4. Conclusions<br />

Centrifuge model-generated broad, open elongate anti<strong>for</strong>ms<br />

<strong>and</strong> domes with few parasitic folds <strong>and</strong> pinched syn<strong>for</strong>ms with<br />

abundant parasitic folds simulate structures within a typical<br />

Archaean greenstone sequence <strong>of</strong> sedimentary <strong>and</strong> volcanic rocks<br />

overlying migmatitic basement gneiss. A combination <strong>of</strong> regional<br />

shortening followed by collapse during channel flow parallel to<br />

the initial axis <strong>of</strong> shortening produces structures that resemble<br />

the dome <strong>and</strong> keel <strong>geom</strong>etry typical in granite-greenstone terrains.<br />

Refolded recumbent folds developed during shortening <strong>and</strong><br />

accentuated during channel flow in the in the basal ductile layers<br />

also resemble folds in Archaean granitic gneisses. It is suggested<br />

that similar processes may have taken place in the Archaean, especially<br />

in the Neoarchaean, when erosion, increased rigidity <strong>of</strong> the<br />

upper crust, horizontal tectonics, <strong>and</strong> channel flow are possible<br />

<strong>and</strong> may produce early folding <strong>and</strong> crustal thickening. We propose<br />

that the typical <strong>for</strong>m <strong>of</strong> some granite–greenstone belts may<br />

be a late feature that is characteristic <strong>of</strong> the Archaean as rocks<br />

remain ductile after initial folding due to greater heat flow facilitating<br />

extrusive channel flow <strong>and</strong> collapse <strong>of</strong> a denser overlying <strong>and</strong><br />

tectonically thickened greenstone sequence. Results <strong>of</strong> our study<br />

<strong>and</strong> the ensuing hypotheses do not constitute unique solutions<br />

nor preclude the likelihood <strong>for</strong> vertical tectonics <strong>and</strong> diapirism in<br />

some terrains, especially prior to the Neoarchaean. Whilst there<br />

are inherent limitations, such as the ability to change rheology<br />

with time <strong>and</strong> the limited size <strong>of</strong> models, centrifuge modelling<br />

combined with CT scanning is shown to be an excellent means<br />

<strong>of</strong> simulating tectonic processes <strong>for</strong> testing alternate hypotheses,<br />

providing “4D” visualization <strong>of</strong> the <strong>for</strong>mation <strong>of</strong> structures.<br />

Model results have applications in mineral exploration as they suggest<br />

changes in tectonic regime during <strong>for</strong>mation <strong>of</strong> structures in<br />

greenstone belts <strong>and</strong> possibly <strong>for</strong> juxtaposition <strong>of</strong> Archaean terrains<br />

<strong>of</strong> different metamorphic grade. Our model may also help<br />

explain the <strong>for</strong>mation <strong>of</strong> structures in some fold <strong>and</strong> thrust belts<br />

where sedimentary layers above an evaporite horizon have undergone<br />

dominantly ductile de<strong>for</strong>mation. Centrifuge modelling also<br />

suggests factors to include in future numerical models. Further<br />

modelling is required to develop the hypotheses presented herein,<br />

<strong>and</strong> to test the effects <strong>of</strong> different erosion <strong>geom</strong>etries <strong>and</strong> to study<br />

differences in the <strong>geom</strong>etry <strong>of</strong> structures developed during orogenparallel<br />

(e.g. Cagnard et al., 2006) <strong>and</strong> lateral constrictional (e.g.<br />

Chardon et al., 2011) ductile flow or protrusion tectonics (Leonov,<br />

1994, 2008), density differences between sequences, granitoid<br />

intrusion, <strong>and</strong> changes in rheology with time.<br />

Acknowledgements<br />

Acknowledgment is made to the Donors <strong>of</strong> the American Chemical<br />

Society Petroleum Research Fund <strong>for</strong> centrifuge modelling <strong>and</strong><br />

CT scanning support to L. Harris <strong>and</strong> to NSERC <strong>for</strong> Discovery grants<br />

to L. Harris <strong>and</strong> L. Godin. Modelling was undertaken by C. Yakymchuk<br />

whilst recipient <strong>of</strong> an NSERC Summer Research Scholarship.<br />

CT scanning was per<strong>for</strong>med by L.-F. d’Aigle in the Quebec Multidisciplinary<br />

Scanning Laboratory. J. Poulin <strong>and</strong> E. Konstantinovskaya<br />

assisted in characterization <strong>of</strong> model materials <strong>and</strong> development <strong>of</strong><br />

CT scanning techniques. 3D reconstructions were undertaken using<br />

Osirix Open Source s<strong>of</strong>tware. The laboratory <strong>for</strong> physical, numerical


152 L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154<br />

<strong>and</strong> geophysical simulations was funded by CFI <strong>and</strong> MELS-Q grants<br />

to L. Harris with contributions from INRS-ETE, BEG (U. Texas at<br />

Austin), Sun Microsystems, Seismic Microtechnology, <strong>and</strong> Norsar.<br />

S. Marshak (who suggested comparisons with salt tectonics) <strong>and</strong><br />

S. Mukherjee are thanked <strong>for</strong> their comments <strong>and</strong> reviews.<br />

Appendix A. Supplementary data<br />

Supplementary data associated with this article can be found,<br />

in the online version, at http://dx.doi.org/10.1016/j.precamres.<br />

2012.04.022.<br />

References<br />

Adam, J., Krézsek, C., 2012. Basin-scale salt tectonic processes <strong>of</strong> the Laurentian<br />

Basin, Eastern Canada: insights from integrated regional 2D seismic interpretation<br />

<strong>and</strong> 4D physical experiments. In: Alsop, G.I., Archer, S.G., Hartley, A.J.,<br />

Grant, N.T., Hodgkinson, R. (Eds.), Salt Tectonics, Sediments <strong>and</strong> Prospectivity,<br />

Special Publications 363. Geological Society, London, pp. 331–360.<br />

Bailey, R.C., 1999. Gravity-driven continental overflow <strong>and</strong> Archaean tectonics.<br />

Nature 39, 413–415.<br />

Beaumont, C., Jamieson, R.A., Nguyen, M.H., Lee, B., 2001. Himalayan tectonics<br />

explained by extrusion <strong>of</strong> a low-viscosity crustal channel coupled to focused<br />

surface denudation. Nature 414, 738–742.<br />

Beaumont, C., Jamieson, R.A., Nguyen, M.H., Medvedev, S., 2004. Crustal channel<br />

flows. 1. Numerical models with applications to the tectonics <strong>of</strong> the<br />

Himalayan-Tibetan orogen. Journal <strong>of</strong> Geophysical Research 109, B06406,<br />

http://dx.doi.org/10.1029/2003JB002809.<br />

Beaumont, C., Nguyen, M.H., Jamieson, R.A., Ellis, S., 2006. Crustal flow modes in large<br />

hot orogens. In: Law, R.D., Searle, M.P., Godin, L. (Eds.), Channel Flow, Ductile<br />

Extrusion <strong>and</strong> Exhumation in Continental Collision Zones, Special Publication<br />

268. Geological Society, London, pp. 91–145.<br />

Bédard, J.H., 2006. A catalytic delamination-driven model <strong>for</strong> coupled genesis <strong>of</strong><br />

Archaean crust <strong>and</strong> sub-continental lithospheric mantle. Geochimica et Cosmochimica<br />

Acta 70, 1188–1214.<br />

Bédard, J.H., Brouillette, P., Madore, L., Berclaz, A., 2003. Archaean cratonization <strong>and</strong><br />

de<strong>for</strong>mation in the northern Superior Province, Canada: an evaluation <strong>of</strong> plate<br />

tectonic versus vertical tectonic models. Precambrian Research 127, 61–87.<br />

Bédard, J.H., Harris, L.B., Thurston, P., 2012. The hunting <strong>of</strong> the snArc. Precambrian<br />

Research, http://dx.doi.org/10.1016/j.precamres.2012.04.001.<br />

Bickle, M.J., Bettenay, L.F., Boulter, C.A., Groves, D.I., Morant, P., 1980. Horizontal<br />

tectonic interaction <strong>of</strong> an Archean gneiss belt <strong>and</strong> greenstones, Pilbara block,<br />

Western Australia. Geology 8, 525–529.<br />

Blewett, R.S., 2002. Archaean Tectonic Processes: a case <strong>for</strong> horizontal shortening in<br />

the North Pilbara granite-greenstone terrane, Western Australia. Precambrian<br />

Research 113, 87–120.<br />

Blewett, R.S., Czarnota, K., 2007. Tectonostratigraphic architecture <strong>and</strong> uplift history<br />

<strong>of</strong> the Eastern Yilgarn Craton. Module 3: Terrane Structure, Project Y1-P763,<br />

Geoscience Australia, Record 2007/15.<br />

Blewett, R.S., Czarnota, K., Henson, P.A., 2010. Structural-event framework <strong>for</strong> the<br />

eastern Yilgarn Craton, Western Australia, <strong>and</strong> its implications <strong>for</strong> orogenic gold.<br />

Precambrian Research 183, 203–229.<br />

Blewett, R.S., Shevchenko, S., Bell, B., 2004. The North Pole Dome: a non-diapiric<br />

dome in the Archaean Pilbara Craton, Western Australia. Precambrian Research<br />

133, 105–120.<br />

Bodorkos, S., S<strong>and</strong>i<strong>for</strong>d, M., 2006. Thermal <strong>and</strong> mechanical controls on the evolution<br />

<strong>of</strong> Archean crustal de<strong>for</strong>mation: examples from Western Australia, in: Benn,<br />

K., Mareschal, J.-C., Condie, K. (Eds.), Archean Geodynamic <strong>and</strong> Environments,<br />

American Geophysical Union Geophysical Monograph Series 164, pp. 131–147.<br />

Bouhallier, H., Chardon, D., Choukroune, P., 1995. Strain patterns in Archaean<br />

dome-<strong>and</strong>-basin structures: the Dharwar craton (Kamataka, South India). Earth<br />

<strong>and</strong> Planetary Science Letters 135, 57–75.<br />

Boutelier, D., Schrank, C., Cruden, A., 2008. Power-law viscous materials <strong>for</strong> analogue<br />

experiments: <strong>new</strong> data on rheology <strong>of</strong> highly-filled silicone polymers. Journal<br />

<strong>of</strong> Structural Geology 30, 341–353.<br />

Boyer, S.E., Elliott, D., 1982. Thrust systems. American Association <strong>of</strong> Petroleum<br />

Geologists Bulletin 66, 1196–1230.<br />

Brown, R.L., Gibson, H.D., 2006. An argument <strong>for</strong> channel flow in the southern Canadian<br />

Cordillera <strong>and</strong> comparison with Himalayan tectonics. Geological Society<br />

Special Publication 268, 543–559.<br />

Bucher, W.H., 1956. Role <strong>of</strong> gravity in orogenesis. Geological Society <strong>of</strong> America<br />

Bulletin 67, 1295–1318.<br />

Burbank, D.W., 2005. Earth science: cracking the Himalaya. Nature 434, 963–964.<br />

Cagnard, F., Durrieu, N., Gapais, D., Brun, J.-P., Ehlers, C., 2006. Crustal thickening <strong>and</strong><br />

lateral flow during compression <strong>of</strong> hot lithospheres, with particular reference<br />

to Precambrian times. Terra Nova 18, 72–78.<br />

Calvert, A.J., Sawyer, E.W., Davis, W.J., Ludden, J.N., 1995. Archaean subduction<br />

inferred from seismic images <strong>of</strong> a mantle suture in the Superior Province. Nature<br />

375, 670–674.<br />

Carreras, J., Druguet, E., Griera, A., 2005. Shear zone-related folds. Journal <strong>of</strong> Structural<br />

Geology 27, 1229–1251.<br />

Cawood, P.A., Kröner, A., Pisarevsky, S., 2006. Precambrian plate tectonics: criteria<br />

<strong>and</strong> evidence. GSA Today 16, 4–11.<br />

Chardon, D., Choukroune, P., Jayan<strong>and</strong>a, M., 1996. Strain patterns, décollement <strong>and</strong><br />

incipient sagducted greenstone terrains in the Archaean Dharwar craton (south<br />

India). Journal <strong>of</strong> Structural Geology 18, 991–1004.<br />

Chardon, D., Choukroune, P., Jayan<strong>and</strong>a, M., 1998. Sinking <strong>of</strong> the Dharwar Basin<br />

(South India): implications <strong>for</strong> Archaean tectonics. Precambrian Research 91,<br />

15–39.<br />

Chardon, D., Gapais, D., Cagnard, F., 2009. Flow <strong>of</strong> ultra-hot orogens: a view from the<br />

Precambrian, clues <strong>for</strong> the Phanerozoic. Tectonophysics 477, 105–118.<br />

Chardon, D., Jayan<strong>and</strong>a, M., Peucat, J.-J., 2011. Lateral constrictional flow <strong>of</strong><br />

hot orogenic crust: insights from the Neoarchean <strong>of</strong> south India, geological<br />

<strong>and</strong> geophysical implications <strong>for</strong> orogenic plateaux. Geochemistry Geophysics<br />

Geosystems 12, 24, http://dx.doi.org/10.1029/2010GC003398.<br />

Chattopadhyay, A., M<strong>and</strong>al, N., 2002. Progressive change in strain patterns <strong>and</strong> fold<br />

styles in a de<strong>for</strong>ming ductile orogenic wedge: an experimental study. Journal <strong>of</strong><br />

Geodynamics 33, 353–376.<br />

Choukroune, P., Bouhallier, H., Arndt, N.T., 1995. S<strong>of</strong>t Lithosphere During Periods <strong>of</strong><br />

Archaean Crustal Growth or Crustal Reworking, Special Publications 95. Geological<br />

Society, London, pp. 67–86.<br />

Choukroune, P., Ludden, J.N., Chardon, D., Calvert, A.J., Bouhallier, H., 1997. Archaean<br />

crustal growth <strong>and</strong> tectonic processes: a comparison <strong>of</strong> the Superior Province,<br />

Canada <strong>and</strong> the Dharwar Craton, India. In: Burg, J.-P., Ford, M. (Eds.), Orogeny<br />

Through Time, Special Publications 121. Geological Society, London, pp. 63–98.<br />

Colletta, B., Vially, R., Chermette, J.C., 1995. Evolution <strong>of</strong> salt-related structures in<br />

compressional settings. In: Jackson, M.P., Roberts, D.G., Snelson, S. (Eds.), Salt<br />

Tectonics: A Global Perspective, AAPG Special Volume 65. , pp. 41–60.<br />

Condie, K.C., 1984. Archaean Crustal Evolution. Developments in Precambrian Geology,<br />

vol. 3. Elsevier, Amsterdam.<br />

Condie, K.C., Kröner, A., Stern, R.J., 2006. When did plate tectonics begin? GSA Today<br />

10, 40–41.<br />

Costa, E., Vendeville, B.C., 2002. Experimental insights on the <strong>geom</strong>etry <strong>and</strong> kinematics<br />

<strong>of</strong> fold-<strong>and</strong>-thrust belts above a weak, viscous evaporite décollement.<br />

Journal <strong>of</strong> Structural Geology 24, 1729–1739.<br />

Cotton, J., Koyi, H.A., 2000. Modelling <strong>of</strong> thrust fronts above ductile <strong>and</strong> frictional<br />

décollements; examples from the Salt Range <strong>and</strong> Potwar Plateau, Pakistan. Geological<br />

Society <strong>of</strong> America Bulletin 112, 351–363.<br />

Culshaw, N., Purves, M., Reynolds, P., Stott, G., 2006a. Post-collisional upper crustal<br />

faulting <strong>and</strong> deep crustal flow in the eastern Wabigoon subprovince <strong>of</strong> the Superior<br />

Province, Ontario: evidence from structural <strong>and</strong> 40 Ar/ 39 Ar data from the<br />

Humboldt Bay High Strain Zone. Precambrian Research 145, 272–288.<br />

Culshaw, N.G., Beaumont, C., Jamieson, R.A., 2006b. The orogenic superstructureinfrastructure<br />

concept: revisited, quantified, <strong>and</strong> revived. Geology 34,<br />

733–736.<br />

Dalstra, H.J., Bloem, E.J.M., Ridley, J.R., Groves, D.I., 1998. Diapirism synchronous<br />

with regional de<strong>for</strong>mation <strong>and</strong> gold mineralisation, a <strong>new</strong> concept <strong>for</strong> granitoid<br />

emplacement in the Southern Cross Province, Western Australia. Geologie en<br />

Mijnbouw 76, 321–338.<br />

de Bremond d’Ars, J., Lécuyer, C., Reynard, B., 1999. Hydrothermalism <strong>and</strong> diapirism<br />

in the Archean: gravitational instability constraints. Tectonophysics 304, 29–39.<br />

Dewei, L.I., 2008. Continental lower-crustal flow: channel flow <strong>and</strong> laminar flow.<br />

Earth Science Frontiers 15, 130–139.<br />

Dirks, P.H.G.M., Jelsma, H.A., 1998. Horizontal accretion <strong>and</strong> stabilization <strong>of</strong> the<br />

Archean Zimbabwe Craton. Geology 26, 11–14.<br />

Dixon, J.M., 1975. Finite strain progressive de<strong>for</strong>mation in models <strong>of</strong> diapiric structures.<br />

Tectonophysics 28, 89–124.<br />

Dixon, J.M., Summers, J.M., 1985. Recent developments in centrifuge modelling <strong>of</strong><br />

tectonic processes: equipment, model construction techniques <strong>and</strong> rheology <strong>of</strong><br />

model materials. Journal <strong>of</strong> Structural Geology 7, 83–102.<br />

Dörr, W., Zulauf, G., 2010. Elevator tectonics <strong>and</strong> orogenic collapse <strong>of</strong> a Tibetanstyle<br />

plateau in the European Variscides: the role <strong>of</strong> the Bohemian shear zone.<br />

International Journal <strong>of</strong> Earth Sciences 99, 299–325.<br />

Duretz, T., Kaus, B.J.P., Schulmann, K., Gapais, D., Kermarrec, J.-J., 2011. Indentation<br />

as an extrusion <strong>mechanism</strong> <strong>of</strong> lower crustal rocks: insight from analogue <strong>and</strong><br />

numerical modelling, application to the Eastern Bohemian Massif. Lithos 124,<br />

158–168.<br />

Drummond, B.J., Hobbs, B.E., Goleby, B.R., 2004. The role <strong>of</strong> crustal fluids in the tectonic<br />

evolution <strong>of</strong> the Eastern Goldfields Province <strong>of</strong> the Archaean Yilgarn Craton,<br />

Western Australia. Earth, Planets <strong>and</strong> Space 56, 1163–1169.<br />

Dziggel, A., Poujol, M., Otto, A., Kisters, A.F.M., Triel<strong>of</strong>, M., Schwarz, W.H., Meyer, F.M.,<br />

2010. New U–Pb <strong>and</strong> 40 Ar/ 39 Ar ages from the northern margin <strong>of</strong> the Barberton<br />

greenstone belt, South Africa: implications <strong>for</strong> the <strong>for</strong>mation <strong>of</strong> Mesoarchaean<br />

gold deposits. Precambrian Research 179, 206–220.<br />

Erickson, E.J., 2010. Structural <strong>and</strong> kinematic analysis <strong>of</strong> the Shagawa Lake shear<br />

zone, Superior Province, northern Minnesota: implications <strong>for</strong> the role <strong>of</strong> vertical<br />

versus horizontal tectonics in the Archean. Canadian Journal <strong>of</strong> Earth Sciences<br />

47, 1463–1479.<br />

Fiduk, J.C., Rowan, M.G., 2012. Analysis <strong>of</strong> Folding <strong>and</strong> De<strong>for</strong>mation Within Layered<br />

Evaporites in Blocks BM-S-8 & -9, Santos Basin, Brazil, Special Publications 363.<br />

Geological Society, London, pp. 471–487.<br />

Flament, N., Coltice, N., Rey, P.F., 2008. A case <strong>for</strong> late-Archaean continental emergence<br />

from thermal evolution models <strong>and</strong> hypsometry. Earth <strong>and</strong> Planetary<br />

Science Letters 275, 326–336.<br />

Fritz, H., Tenczer, V., Hauzenberger, C., Wallbrecher, E., Muhongo, S., 2009. Hot granulite<br />

nappes—tectonic styles <strong>and</strong> thermal evolution <strong>of</strong> the Proterozoic granulite<br />

belts in East Africa. Tectonophysics 477, 160–173.


François, C., Philippot, P., Rey, P., 2012. Formation <strong>and</strong> exhumation <strong>mechanism</strong>s<br />

<strong>of</strong> high-grade rocks: sagduction <strong>and</strong> subduction processes during the Archean.<br />

Geophysical Research Abstracts 14, EGU2012-5136.<br />

Gerya, T.V., Perchuk, L.L., van Reenen, D.D., Smit, C.A., 2000. Two-dimensional<br />

numerical modeling <strong>of</strong> pressure–temperature–time paths <strong>for</strong> the exhumation<br />

<strong>of</strong> some granulite facies terrains in the Precambrian. Journal <strong>of</strong> Geodynamics 30,<br />

17–35.<br />

Godin, L., Grujic, D., Law, R.D., Searle, M.P., 2006. Channel flow, ductile extrusion<br />

<strong>and</strong> exhumation in continental collision zones: an introduction. In: Law, R.D.,<br />

Searle, M.P., Godin, L. (Eds.), Channel Flow, Ductile Extrusion <strong>and</strong> Exhumation in<br />

Continental Collision Zones, Special Publication 268. Geological Society, London,<br />

pp. 1–23.<br />

Godin, L., Yakymchuk, C., Harris, L.B., 2011. Himalayan hinterl<strong>and</strong>-verging<br />

superstructure folds related to <strong>for</strong>el<strong>and</strong>-directed infrastructure ductile flow:<br />

insights from centrifuge analogue modelling. Journal <strong>of</strong> Structural Geology 33,<br />

239–342.<br />

Goodwin, A.M., Smith, I.E.M., 1980. Chemical discontinuities in Archaean metavolcanic<br />

terrains <strong>and</strong> the development <strong>of</strong> Archaean crust. Precambrian Research<br />

10, 301–311.<br />

Gorman, B.E., Pearce, T.H., Birkette, T.C., 1978. On the structure <strong>of</strong> Archean greenstone<br />

belts. Precambrian Research 6, 23–41.<br />

Goscombe, B., Blewett, R.S., Czarnota, K., Groe<strong>new</strong>ald, P.B., Maas, R., 2009. Metamorphic<br />

Evolution <strong>and</strong> Integrated Terrane Analysis <strong>of</strong> the Eastern Yilgarn Craton:<br />

Rationale, Methods, Outcomes <strong>and</strong> Interpretation. Geoscience Australia, Record<br />

2009/23, 270 pp.<br />

Gray, R., Pysklywec, R.N., 2010. Geodynamic models <strong>of</strong> Archean continental collision<br />

<strong>and</strong> the <strong>for</strong>mation <strong>of</strong> mantle lithosphere keels. Geophysical Research Letters 37,<br />

L19301, http://dx.doi.org/10.1029/2010gl043965.<br />

Groves D.I., Engl<strong>and</strong>, G., Rasmussen, B., Krapez, B., 2006. A comparison<br />

between the Witwatersr<strong>and</strong> <strong>and</strong> orogenic gold systems: a test<br />

<strong>of</strong> the hydrothermal Witwatersr<strong>and</strong> model. Powerpoint presentation,<br />

ftp://ftp.uwa.edu.au/pub/segs/2 Wits.ppt (accessed January 2011).<br />

Groves, D.I., Goldfarb, R.J., Knox-Robinson, C.M., Ojala, J., Gardoll, S.J., Yun, G.Y., Holyl<strong>and</strong>,<br />

P., 2000. Late-kinematic timing <strong>of</strong> orogenic gold deposits <strong>and</strong> significance<br />

<strong>for</strong> computer-based exploration techniques with emphasis on the Yilgarn Block,<br />

Western Australia. Ore Geology Reviews 96, 101–112.<br />

Grujic, D., 2006. Channel flow <strong>and</strong> continental collision tectonics: an overview. In:<br />

Law, R.D., Searle, M.P., Godin, L. (Eds.), Channel Flow, Ductile Extrusion <strong>and</strong><br />

Exhumation in Continental Collision Zones, Special Publication 268. Geological<br />

Society, London, pp. 25–37.<br />

Grujic, D., Hollister, L.S., Parrish, R.R., 2002. Himalayan metamorphic sequence as an<br />

orogenic channel: insight from Bhutan. Earth <strong>and</strong> Planetary Science Letters 198,<br />

177–191.<br />

Hamilton, W.B., 2007a. Driving <strong>mechanism</strong> <strong>and</strong> 3-D circulation <strong>of</strong> plate tectonics. In:<br />

Sears, J.W., Harms, T.A., Evenchick, C.A. (Eds.), Whence the Mountains? Inquiries<br />

into the Evolution <strong>of</strong> Orogenic Systems: a Volume in Honour <strong>of</strong> Raymond A. Price,<br />

Geological Society <strong>of</strong> America Special Paper 433. , pp. 1–25.<br />

Hamilton, W.B., 2007b. Earth’s first two billion years—the era <strong>of</strong> internally mobile<br />

crust. In: Hatcher Jr., R.D., Carlson, M.P., McBride, J.H., Martínez Catalán, J.R.<br />

(Eds.), 4-D Framework <strong>of</strong> Continental Crust, vol. 200. Geological Society <strong>of</strong> America<br />

Memoir, pp. 233–296.<br />

Harris, L.B., Koyi, H.A., 2003. Centrifuge modelling <strong>of</strong> folding in high-grade rocks<br />

during rifting. Journal <strong>of</strong> Structural Geology 25, 291–305.<br />

Harris, L.B., Koyi, H.A., Fossen, H., 2002. Mechanisms <strong>for</strong> folding <strong>of</strong> highgrade<br />

rocks in extensional tectonic settings. Earth Science Reviews 59,<br />

163–210.<br />

Harris, L.B., Yakymchuk, C., Godin, L., 2008. Implications <strong>of</strong> centrifuge modelling<br />

<strong>of</strong> folding <strong>and</strong> channel flow <strong>for</strong> regional structural interpretation <strong>and</strong><br />

mineral exploration in fold belts <strong>and</strong> granite-greenstone terrains. Québec<br />

Exploration, 2008. http://www.quebecexploration.qc.ca/2008/english/exhibit-<br />

179.asp (accessed January 2011).<br />

Harris, L.B., Yakymchuk, C., Godin, L., 2009. Implications <strong>of</strong> centrifuge modelling <strong>of</strong><br />

folding <strong>and</strong> channel flow <strong>for</strong> regional structural interpretation in fold belts <strong>and</strong><br />

granite-greenstone terrains. In: Williams, P.J., et al. (Eds.), 10th Biennial SGA<br />

Meeting <strong>of</strong> The Society <strong>for</strong> Geology Applied to Mineral Deposits: Smart Science<br />

<strong>for</strong> Exploration <strong>and</strong> Mining 2. Economic Geology Research Unit, James Cook<br />

University, Townsville, Australia, pp. 821–823.<br />

Harris, L.B., Yakymchuk, C., Godin, L., 2010. Centrifuge modelling <strong>of</strong> folding<br />

<strong>and</strong> channel flow in orogenic belts <strong>and</strong> Archaean granite-greenstone<br />

terrains. Extended abstracts, GeoMod 2010, 27–29 September, Lisbon,<br />

Portugal. http://<strong>geom</strong>od2010.fc.ul.pt/abstracts/Harris%20et%20al.pdf (accessed<br />

December 2010).<br />

Harris, L.B., Yakymchuk, C., Godin, L., 2012. Implications <strong>of</strong> centrifuge simulations <strong>of</strong><br />

channel flow <strong>for</strong> opening out or destruction <strong>of</strong> folds. Tectonophysics 526–529,<br />

67–87.<br />

Harris, N., 2007. Channel flow <strong>and</strong> the Himalayan-Tibetan orogen: a critical review.<br />

Journal <strong>of</strong> the Geological Society, London 164, 511–523.<br />

Hatcher, R.D., Merschat, A.J., 2006. The Appalachian Inner Piedmont: an exhumed<br />

strike-parallel, tectonically <strong>for</strong>ced orogenic channel. In: Law, R.D., Searle, M.P.,<br />

Godin, L. (Eds.), Channel Flow, Ductile Extrusion <strong>and</strong> Exhumation in Continental<br />

Collision Zones, Special Publication 268. Geological Society, London, pp.<br />

517–541.<br />

Hessler, A.M., Lowe, D.R., 2006. Weathering <strong>and</strong> sediment generation in the Archean:<br />

an integrated study <strong>of</strong> the evolution <strong>of</strong> siliciclastic sedimentary rocks <strong>of</strong> the<br />

3.2 Ga Moodies Group, Barberton Greenstone Belt, South Africa. Precambrian<br />

Research 151, 185–210.<br />

L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154 153<br />

Hickman, A.H., 1984. Archaean diapirism in the Pilbara Block, Western Australia. In:<br />

Kröner, A., Greiling, R. (Eds.), Precambrian Tectonics Illustrated. E. Schweizerbart’sche<br />

Verlagsbuchh<strong>and</strong>lung, Stuttgart, pp. 113–127.<br />

Hippertt, J., Davis, B., 2000. Dome emplacement <strong>and</strong> <strong>for</strong>mation <strong>of</strong> kilometre-scale<br />

synclines in a granite-greenstone terrain (Quadrilátero Ferrífero, southeastern<br />

Brazil). Precambrian Research 102, 99–121.<br />

H<strong>of</strong>mann, A., Jagoutz, O., Kröner, A., Dirks, P.H.G.M., Jelsma, H.A., 2003. The Chirwa<br />

dome: granite emplacement during late Archaean thrusting along the northeastern<br />

margin <strong>of</strong> the Zimbabwe craton. South African Journal <strong>of</strong> Geology 105,<br />

285–300.<br />

Hollister, L.S., Grujic, D., 2006. Pulsed channel flow in Bhutan. In: Law, R.D., Searle,<br />

M.P., Godin, L. (Eds.), Channel Flow, Ductile Extrusion <strong>and</strong> Exhumation in Continental<br />

Collision Zones, Special Publications 268. Geological Society, London, pp.<br />

415–423.<br />

Hubbert, M.K., 1937. Theory <strong>of</strong> scale models as applied to the study <strong>of</strong> geologic<br />

structures. Bulletin <strong>of</strong> the Geological Society <strong>of</strong> America 48, 1459–1519.<br />

Hudec, M.R., Jackson, M.P.A., 2007. Terra infirma: underst<strong>and</strong>ing salt tectonics.<br />

Earth-Science Reviews 82, 1–28.<br />

Ings, S., Beaumont, C., Gemmer, L., 2004. Numerical modeling <strong>of</strong> salt tectonics<br />

on passive continental margins: preliminary assessment <strong>of</strong> the effects <strong>of</strong><br />

sediment loading, buoyancy, margin tilt, <strong>and</strong> isostasy, in: Post, P., Olson, D.,<br />

Lyons, K., Palmes, S., Harrison, P., Rosen, N. (Eds.), Salt Sediment Interactions<br />

<strong>and</strong> Hydrocarbon Prospectivity: Concepts, Applications, <strong>and</strong> Case Studies<br />

<strong>for</strong> the 21st Century. Proceedings <strong>of</strong> the 24th Annual GCSSEPM Foundation<br />

Bob F. Perkins Research Conference 2004, 37 pp. CD Rom – ISSN: 1544-2462<br />

<strong>and</strong> http://myweb.dal.ca/sings/webpage files/GCSSEPM04/GCSSEPM04.html<br />

(accessed April 2012).<br />

Jamieson, R.A., Beaumont, C., Medvedev, S., Nguyen, M.H., 2004. Crustal channel<br />

flows. 2. Numerical models with implications <strong>for</strong> metamorphism in<br />

the Himalayan-Tibetan orogen. Journal <strong>of</strong> Geophysical Research 109, B064,<br />

http://dx.doi.org/10.1029/2003JB002811.<br />

Jones, R.R., Holdsworth, R.E., H<strong>and</strong>, M., Goscombe, B., 2006. Ductile extrusion in<br />

continental collision zones: ambiguities in the definition <strong>of</strong> channel flow <strong>and</strong><br />

its identification in ancient orogens. In: Law, R.D., Searle, M.P., Godin, L. (Eds.),<br />

Channel Flow, Ductile Extrusion <strong>and</strong> Exhumation in Continental Collision Zones,<br />

Special Publication 268. Geological Society, London, pp. 201–219.<br />

Kisters, A.F.M., Stevens, G., Dziggel, A., Armstrong, R.A., 2003. Extensional<br />

detachment faulting <strong>and</strong> core-complex <strong>for</strong>mation in the southern Barberton<br />

granite-greenstone terrain, South Africa: evidence <strong>for</strong> a 3.2 Ga orogenic collapse.<br />

Precambrian Research 127, 355–378.<br />

Kloppenburg, A., White, S.H., Zegers, T.E., 2001. Structural evolution <strong>of</strong> the Warrawoona<br />

Greenstone Belt <strong>and</strong> adjoining granitoid complexes, Pilbara Craton,<br />

Australia: implications <strong>for</strong> Archaean tectonic processes. Precambrian Research<br />

112, 107–147.<br />

Kolodyazhny, S.Y., 2007. Structural-kinematic evolution <strong>of</strong> the Central Belomorian-<br />

Lapl<strong>and</strong> Belt in the Paleoproterozoic. Geotectonics 41, 210–230.<br />

Kröner, A., 1981. Precambrian plate tectonics. In: Kroner, A. (Ed.), Precambrian Plate<br />

Tectonics. Elsevier, pp. 57–90.<br />

Kusky, T.M., 1993. Collapse <strong>of</strong> Archean orogens <strong>and</strong> the generation <strong>of</strong> late- to<br />

postkinematic granitoids. Geology 21, 925–928.<br />

Lana, C., Kisters, A.F.M., Stevens, G., 2010a. Exhumation <strong>of</strong> Mesoarchaean TTG<br />

gneisses from the middle crust: insights from the Steynsdorp core complex, Barberton<br />

granitoid greenstone terrain, South Africa. Geological Society <strong>of</strong> America<br />

Bulletin 122, 183–197.<br />

Lana, C., Tohver, E., Cawood, P., 2010b. Quantifying rates <strong>of</strong> dome-<strong>and</strong>-keel <strong>for</strong>mation<br />

in the Barberton granitoid-greenstone belt, South Africa. Precambrian Research<br />

177, 199–211.<br />

Leonov, M.G., 1994. Interior mobility <strong>of</strong> the basement <strong>and</strong> tectonogenesis <strong>of</strong> activated<br />

plat<strong>for</strong>ms. Geotectonics 27, 369–383.<br />

Leonov, M.G., 2008. Lateral protrusions in the structure <strong>of</strong> the Earth’s lithosphere.<br />

Geotectonics 42, 327–356.<br />

Lin, S., 2005. Synchronous vertical <strong>and</strong> horizontal tectonism in the Neoarchean:<br />

kinematic evidence from a synclinal keel in the northwestern Superior Craton,<br />

Manitoba. Precambrian Research 139, 181–194.<br />

Lobato, L.M., Ribeiro-Rodrigues, L.C., Zucchetti, M., Noce, C.M., Baltazar, O.F., da Silva,<br />

C.L., Pinto, C.P., 2001. Brazil’s premier gold province. Part I. The tectonic, magmatic,<br />

<strong>and</strong> structural setting <strong>of</strong> the Archean Rio das Velhas greenstone belt,<br />

Quadrilatero Ferrifero. Mineralium Deposita 36, 228–249.<br />

Lowe, D.R., Tice, M.M., 2004. Geologic evidence <strong>for</strong> Archean atmospheric <strong>and</strong> climatic<br />

evolution: fluctuating levels <strong>of</strong> CO2, CH4, <strong>and</strong> O2 with an overriding tectonic<br />

control. Geology 32, 493–496.<br />

Ludden, J., Hubert, C., Barnes, A., Milkereit, B., Sawyer, E., 1993. A three dimensional<br />

perspective on the evolution <strong>of</strong> Archaean crust: LITHOPROBE seismic reflection<br />

images in the southwestern Superior province. Lithos 30, 357–372.<br />

Luján, M., Storti, F., Balanyá, J.-C., Crespo-Blanc, A., Rossetti, F., 2003. Role <strong>of</strong> décollement<br />

material with different rheological properties in the structure <strong>of</strong> the<br />

Aljibe thrust imbricate (Flysch Trough, Gibraltar Arc): an analogue modelling<br />

approach. Journal <strong>of</strong> Structural Geology 25, 867–881.<br />

Mair, J.L., Ojala, V.J., Salier, B.P., Groves, D.I., 2001. Application <strong>of</strong> stress mapping in<br />

cross-section to underst<strong>and</strong>ing ore <strong>geom</strong>etry, predicting ore zones <strong>and</strong> development<br />

<strong>of</strong> drilling strategies. Australian Journal <strong>of</strong> Earth Sciences 47, 895–912.<br />

Marshak, S., 1999. De<strong>for</strong>mation style way back when: thoughts on the contrasts<br />

between Archean/Paleoproterozoic orogens <strong>and</strong> modern ones. Journal <strong>of</strong> Structural<br />

Geology, 1175–1182 (20th anniversary issue 21).<br />

Marshak, S., Tinkham, D., Alkmim, F., Brueckner, H., Bornhorst, T., 1997. Dome<strong>and</strong>-keel<br />

provinces <strong>for</strong>med during Paleoproterozoic orogenic collapse-core


154 L.B. Harris et al. / Precambrian Research 212-213 (2012) 139–154<br />

complexes, diapirs, or neither? Examples from the Quadrilátero Ferrífero <strong>and</strong><br />

the Penokean orogen. Geology 25, 415–418.<br />

McQuarrie, N., 2004. Crustal scale <strong>geom</strong>etry <strong>of</strong> the Zagros fold—thrust belt, Iran.<br />

Journal <strong>of</strong> Structural Geology 26, 519–535.<br />

Mercier, E., Rafi, S., Ahmadi, R., 2007. Folds kinematics in “fold-<strong>and</strong>-thrust belts” the<br />

“hinge migration” question, a review. In: Thrust Belts <strong>and</strong> Forel<strong>and</strong> Basins, From<br />

Fold Kinematics to Hydrocarbon Systems, Chapter 7, Frontiers in Earth Sciences.<br />

Springer, Berlin/Heidelberg, pp. 135–147.<br />

Miller, V., 1982. Lateral <strong>and</strong> sub-lateral rock flow <strong>and</strong> its role in the <strong>for</strong>mation <strong>of</strong><br />

structural patterns (ПOCЛOЙНOE И CУБПOCЛOЙНOE TEЧEНИE ПOРOД<br />

И EƔO РOЛЬВCTРУКTУРOOБРAЗOВAНИИ). Geotectonics 6, 88–96 (in<br />

Russian).<br />

Mukherjee, S., Koyi, H.A., 2010. Higher Himalayan Shear Zone, Sutlej section: structural<br />

geology <strong>and</strong> extrusion <strong>mechanism</strong> by various combinations <strong>of</strong> simple<br />

shear, pure shear <strong>and</strong> channel flow in shifting modes. International Journal <strong>of</strong><br />

Earth Sciences 99, 1267–1303.<br />

Mukherjee, S., Koyi, H.A., Talbot, C.J., 2012. Implications <strong>of</strong> channel flow analogue<br />

models <strong>for</strong> extrusion <strong>of</strong> the Higher Himalayan Shear Zone with special reference<br />

to the out-<strong>of</strong>-sequence thrusting. International Journal <strong>of</strong> Earth Sciences 101,<br />

253–272.<br />

Myers, J.S., Watkins, K.P., 1985. Origin <strong>of</strong> granite-greenstone patterns, Yilgarn Block,<br />

Western Australia. Geology 13, 778–780.<br />

Park, R.G., 1982. Archaean tectonics. Geologische Rundschau 71, 22–37.<br />

Parmenter, A.C., Lin, S., Corkery, M.T., 2006. Structural evolution <strong>of</strong> the Cross Lake<br />

greenstone belt in the northwestern Superior Province, Manitoba: implications<br />

<strong>for</strong> relationship between vertical <strong>and</strong> horizontal tectonism. Canadian Journal <strong>of</strong><br />

Earth Sciences 43, 767–787.<br />

Peschler, A.P., Benn, K., Roest, W.R., 2004. Insights on Archean continental geodynamics<br />

from gravity modelling <strong>of</strong> granite-greenstone terranes. Journal <strong>of</strong><br />

Geodynamics 38, 185–207.<br />

Poulin, J., 2006. De la médecine à la géologie—visualisation des modèles physiques<br />

par tomodensitométrie. MSc Thesis, INRS-ETE, Québec. http://ete.inrs.ca/<br />

pub/theses/T000478.pdf <strong>and</strong> http://ete.inrs.ca/pub/theses/T000478.zip<br />

(accessed September 2010).<br />

Price, N.J., Cosgrove, J.W., 1990. Analysis <strong>of</strong> Geological Structures. Cambridge University<br />

Press, 502 pp.<br />

Raimondo, T., Collins, A.S., H<strong>and</strong>, M., Walker-Hallam, A., Smithies, R.H., Evins,<br />

P.M., Howard, H.M., 2009. Ediacaran intracontinental channel flow. Geology 37,<br />

291–294.<br />

Ramberg, H., 1967a. Model experimentation <strong>of</strong> the effect <strong>of</strong> gravity on tectonic<br />

processes. Geophysical Journal <strong>of</strong> the Royal Astronomical Society 14,<br />

307–329.<br />

Ramberg, H., 1967b. Gravity, De<strong>for</strong>mation <strong>and</strong> the Earth’s Crust. Academic Press,<br />

London, 214 pp.<br />

Ramberg, H., 1970. Folding <strong>of</strong> laterally compressed multilayers in the field <strong>of</strong> gravity.<br />

Physics <strong>of</strong> the Earth <strong>and</strong> Planetary Interiors, 203–232.<br />

Ramberg, H., 1980. Diapirism <strong>and</strong> gravity collapse in the Sc<strong>and</strong>inavian Caledonides.<br />

Journal <strong>of</strong> the Geological Society 137, 261–270.<br />

Ramberg, H., 1981a. Gravity, De<strong>for</strong>mation <strong>and</strong> the Earth’s Crust, 2nd ed. Academic<br />

Press, London/New York/Toronto/Sydney/San Francisco, 452 pp.<br />

Ramberg, H., 1981b. The Role <strong>of</strong> Gravity in Orogenic Belts, Special Publications 9.<br />

Geological Society, London, pp. 125–140.<br />

Ramsay, J.G., Huber, M.I., Lisle, R.J., 1987. The Techniques <strong>of</strong> Modern Structural Geology,<br />

Volume 2: Folds <strong>and</strong> Fractures. Academic Press, 392 pp.<br />

Reston, T.J., 1990. Mantle shear zones <strong>and</strong> the evolution <strong>of</strong> the northern North Sea<br />

basin. Geology 18, 272–275.<br />

Rey, P.F., Coltice, N., 2008. Neoarchean lithospheric strengthening <strong>and</strong> the coupling<br />

<strong>of</strong> Earth’s geochemical reservoirs. Geology 36, 635–638.<br />

Rey, P.F., Houseman, G., 2006. Lithospheric scale gravitational flow: the impact<br />

<strong>of</strong> body <strong>for</strong>ces on orogenic processes from Archaean to Phanerozoic. In:<br />

Buiter, S.J.H., Schreurs, G. (Eds.), Analogue <strong>and</strong> Numerical Modelling <strong>of</strong><br />

Crustal-Scale Processes, Special Publications 253. Geological Society, London,<br />

pp. 153–167.<br />

Richter, F.M., 1985. Models <strong>for</strong> the Archean thermal regime. Earth <strong>and</strong> Planetary<br />

Science Letters 73, 350–360.<br />

Robin, C.M.I., Bailey, R.C., 2009. Simultaneous generation <strong>of</strong> Archean crust <strong>and</strong> subcratonic<br />

roots by vertical tectonics. Geology 37, 523–526.<br />

Rowan, M.G., Peel, F.J., Vendeville, B.C., 2004. Gravity-driven foldbelts on passive<br />

margins. In: McClay, K.R. (Ed.), Thrust Tectonics <strong>and</strong> Hydrocarbon Systems, vol.<br />

82. AAPG Memoir, pp. 157–182.<br />

Sans, M., 2003. From thrust tectonics to diapirism. The role <strong>of</strong> evaporites in the kinematic<br />

evolution <strong>of</strong> the eastern South Pyrenean front. Geologica Acta 1, 239–259.<br />

Schoene, B., de Wit, M.J., Bowring, S.A., 2008. Mesoarchean assembly <strong>and</strong> stabilization<br />

<strong>of</strong> the eastern Kaapvaal craton: a structural-thermochronological<br />

perspective. Tectonics 27, TC5010, http://dx.doi.org/10.1029/2008TC002267.<br />

Schulmann, K., Kröner, A., Hegner, E., Wendt, I., Konopásek, J., Lexa, O., ˇStípská, P.,<br />

2005. Chronological constraints on the pre-orogenic history, burial <strong>and</strong> exhumation<br />

<strong>of</strong> deep-seated rocks along the eastern margin <strong>of</strong> the Variscan Orogen,<br />

Bohemian Massif, Czech Republic. American Journal <strong>of</strong> Science 305, 407–448.<br />

Schulmann, K., Lexa, O., ˇStípská, P., Racek, M., Tajčanová, L., Konopásek, J., Edel, J.-B.,<br />

Peschler, A., Lehmann, J., 2008. Vertical extrusion <strong>and</strong> horizontal channel flow <strong>of</strong><br />

orogenic lower crust: key exhumation <strong>mechanism</strong>s in large hot orogens? Journal<br />

<strong>of</strong> Metamorphic Geology 26, 273–297.<br />

Sengupta, S., Ghosh, S.K., Deb, S.K., Khan, D., 2005. Opening <strong>and</strong> closing <strong>of</strong> folds in<br />

superposed de<strong>for</strong>mations. Journal <strong>of</strong> Structural Geology 27, 1282–1299.<br />

Shackleton, R.M., 1995. Tectonic evolution <strong>of</strong> greenstone belts. In: Coward, M.E.,<br />

Ries, A.C. (Eds.), Early Precambrian Processes, Special Publications 95. Geological<br />

Society, London, pp. 53–65.<br />

Sharkov, E.V., Bogatikov, O.A., 2010. Tectonomagmatic evolution <strong>of</strong> the Earth <strong>and</strong><br />

Moon. Geotectonics 44, 85–101.<br />

Snowden, P.A., Bickle, M.J., 1976. The Chinamora Batholith: diapiric intrusion or<br />

interference fold? Journal <strong>of</strong> the Geological Society 132, 131–137.<br />

Talbot, C.J., 1974. Fold nappes as asymmetric mantled gneiss domes <strong>and</strong> ensialic<br />

orogeny. Tectonophysics 24, 259–276.<br />

Tinkham, D.K., Marshak, S., 2004. Precambrian Dome-<strong>and</strong>-Keel Structure in the<br />

Penokean Orogenic Belt <strong>of</strong> Northern Michigan, USA. GSA Special Papers 380,<br />

pp. 321–338.<br />

V<strong>and</strong>erhaeghe, O., 2009. Migmatites, granites <strong>and</strong> orogeny: flow modes <strong>of</strong> partiallymolten<br />

rocks <strong>and</strong> magmas associated with melt/solid segregation in orogenic<br />

belts. Tectonophysics 477, 119–134.<br />

Van Kranendonk, M.J., Collins, W.J., Hickman, A., Pawley, M.J., 2004. Critical tests<br />

<strong>of</strong> vertical vs. horizontal tectonic models <strong>for</strong> the Archaean East Pilbara Granite-<br />

Greenstone Terrane, Pilbara Craton, Western Australia. Precambrian Research<br />

131, 173–211.<br />

Wegmann, C.E., 1935. Zur Deutung der Migmatite. Geologische Rundschau 26,<br />

305–350.<br />

Weijermars, R., 1986. Flow behaviour <strong>and</strong> physical chemistry <strong>of</strong> bouncing putties<br />

<strong>and</strong> related polymers in view <strong>of</strong> tectonic laboratory applications. Tectonophysics<br />

124, 325–358.<br />

Weinberg, R.F., Moresi, L., van der Borgh, P., 2003. Timing <strong>of</strong> de<strong>for</strong>mation in<br />

the Norseman-Wiluna Belt, Yilgarn Craton, Western Australia. Precambrian<br />

Research 120, 219–239.<br />

Weinberg, R.F., Schmeling, H., 1992. Polydiapirs: multiwavelength gravity structures.<br />

Journal <strong>of</strong> Structural Geology 14, 425–436.<br />

White, S.H., Zegers, T.E., van Haaften, W.M., Kloppenburg, A., Wijbrans, J., 1998. Tectonic<br />

evolution <strong>of</strong> the Eastern Pilbara, Australia (extended abstract). Geologie<br />

en Mijnbouw 76, 343–347.<br />

Wilcock, W.S.D., 1991. The Rayleigh–Taylor instability <strong>of</strong> an embedded layer <strong>of</strong> lowviscosity<br />

fluid. Journal <strong>of</strong> Geophysical Research 96, 12193–12200.<br />

Wilson, P., White, J.C., Roulston, B.V., 2006. Structural geology <strong>of</strong> the Penobsquis<br />

salt structure: late Bashkirian inversion tectonics in the Moncton Basin, New<br />

Brunswick, eastern Canada. Canadian Journal <strong>of</strong> Earth Sciences 43, 405–419.<br />

Williams, P.R., Whitaker, A.J., 1993. Gneiss domes <strong>and</strong> extensional de<strong>for</strong>mation in<br />

the highly mineralised Archaean Eastern Goldfields Province, Western Australia.<br />

Ore Geology Reviews 8, 141–162.<br />

Windley, B.F., 1995. The Evolving Continents, 3rd ed. John Wiley, Chester, 526 pp.<br />

Windley, B.F., 1998. Tectonic models <strong>for</strong> the geological evolution <strong>of</strong> crust, cratons<br />

<strong>and</strong> continents in the Archaean. Revista Brasileira de Geociencias 28, 183–188.<br />

Windley, B.F., Garde, A.A., 2009. Arc-generated blocks with crustal sections in the<br />

North Atlantic craton <strong>of</strong> West Greenl<strong>and</strong>: crustal growth in the Archean with<br />

modern analogues. Earth-Science Reviews 93, 1–30.<br />

Yakymchuk, C., Harris, L.B., Godin, L., 2012. Centrifuge modelling <strong>of</strong> de<strong>for</strong>mation <strong>of</strong><br />

a multi-layered sequence over a ductile substrate. 1. Style <strong>and</strong> 4D <strong>geom</strong>etry <strong>of</strong><br />

active cover folds during layer-parallel shortening. International Journal <strong>of</strong> Earth<br />

Sciences 101, 463–482.<br />

Yamato, P., Kaus, B.J.P., Mouthereau, M., Castelltort, S., 2011. Dynamic constraints<br />

on the crustal-scale rheology <strong>of</strong> the Zagros fold belt, Iran. Geology 39, 815–818.<br />

Zegers, T.E., Nelson, D.R., Wijbrans, J.R., White, S.H., 2001. SHRIMP U–Pb zircon dating<br />

<strong>of</strong> Archean core complex <strong>for</strong>mation <strong>and</strong> pancratonic strike-slip de<strong>for</strong>mation in<br />

the East Pilbara granite-greenstone terrain. Tectonics 20, 883–908.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!