08.08.2013 Views

Van Kranendonketal_2010.pdf - STARPLAN

Van Kranendonketal_2010.pdf - STARPLAN

Van Kranendonketal_2010.pdf - STARPLAN

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.

Age, lithology and structural evolution of the c. 3.53 Ga Theespruit Formation in the<br />

Tjakastad area, southwestern Barberton Greenstone Belt, South Africa, with<br />

implications for Archaean tectonics<br />

M.J. <strong>Van</strong> Kranendonk a, ⁎, A. Kröner b , E. Hegner c , J. Connelly d<br />

a Geological Survey of Western Australia, 100 Plain St., East Perth, WA 6004, Australia<br />

b Institut für Geowissenschaften, Universität Mainz, 55099 Mainz, Germany<br />

c Department für Geo-und Umweltwissenschaften, Universität München, Theresienstrasse 41, D-80333 München, Germany<br />

d Department of Geological Sciences, University of Texas at Austin, Austin, Texas, 78712-1101, USA<br />

article info<br />

Article history:<br />

Accepted 4 November 2008<br />

Keywords:<br />

Archean<br />

Barberton greenstone belt<br />

Theespruit Formation<br />

Tectonics<br />

1. Introduction<br />

abstract<br />

A major debate in the Earth sciences concerns the style of early<br />

Archaean tectonics, whether dominated by vertical movements due to<br />

vigorous mantle convection, plume-derived magmatism, and crustal<br />

diapirism, or horizontal plate interactions akin to the modern tectonic<br />

paradigm (e.g. de Wit, 1998 versus Hamilton, 1998). Although both<br />

tectonic processes have been separately identified as affecting<br />

different Archaean cratons, or different parts of such cratons (e.g.,<br />

Myers and Kröner, 1994; <strong>Van</strong> Kranendonk et al., 2007), it is the relative<br />

contribution of these two processes that is still widely debated. This is<br />

particularly true regarding the formation of Palaeoarchaean granitoidgreenstone<br />

terrains (GGTs) that have a characteristic dome-and-basin<br />

map pattern, the likes of which do not occur in modern geological<br />

environments (Anhaeusser et al., 1969).<br />

This debate is particularly vigorous regarding the formation of<br />

Archaean crust in the Barberton Mountain Land part of the Kaapvaal<br />

Craton, southern Africa (Fig. 1). On the one hand, Viljoen and Viljoen<br />

⁎ Corresponding author.<br />

E-mail address: martin.vankranendonk@doir.wa.gov.au (M.J. <strong>Van</strong> Kranendonk).<br />

Chemical Geology 261 (2009) 115–139<br />

Contents lists available at ScienceDirect<br />

Chemical Geology<br />

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

0009-2541/$ – see front matter. Crown Copyright © 2008 Published by Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.chemgeo.2008.11.006<br />

A field and petrographic re-assessment of rocks from the Theespruit Formation near Tjakastad, southwestern<br />

Barberton Greenstone Belt, is combined with new zircon U–Pb ages and whole-rock Sm–Nd isotopic results<br />

to reveal several important inconsistencies with a previous thrust-accretion model for this area. It was found<br />

that faults previously interpreted as ‘thrusts’ are extensional, ‘thrust slices of basement orthogneiss’ are little<br />

deformed quartz-feldspar porphyries or sheared felsic volcaniclastic sedimentary rocks, and ‘sedimentary<br />

diamictites’ are felsic agglomerates. These observations, combined with consistent facing directions of<br />

bedding, the recognition of distinct stratigraphic variations across strike, and U–Pb zircon dates from 13 felsic<br />

metavolcanic samples, all suggest that the Theespruit Formation is a moderately strained, coherent<br />

stratigraphic succession deposited at c. 3530 Ma and affected by extensional shear deformation at c. 3230 Ma.<br />

A 3453±6 Ma population of zircons analysed previously from an amphibolite-facies Theespruit Formation<br />

felsic schist is interpreted as being of metamorphic, rather than detrital, origin, and arising from heat<br />

associated with felsic plutonic rocks that stitch the Theespruit Formation to the overlying Komati Formation<br />

at ca. 3450 Ma, 220 Ma earlier than the proposed thrusting event.<br />

Crown Copyright © 2008 Published by Elsevier B.V. All rights reserved.<br />

(1969) first interpreted the evolution of this area in terms of a littledeformed<br />

volcano-sedimentary (greenstone) succession, the Barberton<br />

Greenstone Belt, intruded by granitoid magmas and then tilted<br />

and folded by the vertical rise of tonalitic diapirs (Anhaeusser, 1984).<br />

However, the diapiric model was challenged by de Wit and colleagues<br />

who identified nappes, reverse faults, high strain zones, and local<br />

repetitions of stratigraphy in the greenstones (de Wit, 1982, 1983,<br />

1991; de Wit et al., 1983, 1987a). These authors deduced that the<br />

Barberton Mountain Land part of the Kaapvaal Craton formed through<br />

horizontal thrust-accretionary plate tectonic processes in which<br />

granitoid diapirism played no part (de Wit et al., 1992).<br />

Whereas nappes and reverse faults may form in a variety of<br />

environments, including diapirism (e.g. Ramberg, 1967; Dixon and<br />

Summers, 1983), the single critical piece of geological evidence<br />

supporting the thrust-accretion hypothesis is the interpreted largescale<br />

occurrence of older-over-younger rocks in the Theespruit area,<br />

located in the southwestern part of the Barberton Greenstone Belt<br />

(Fig. 1) (de Wit et al., 1983, 1992; Armstrong et al., 1990). In this area,<br />

ca. 3.49–3.48 Ga (Lopez-Martinez et al., 1992; Dann, 2000) komatiites,<br />

basalts and a layer of felsic tuff of the Komati Formation occur<br />

structurally above interlayered pillow basalt, sedimentary ‘diamictite’,<br />

and ‘orthogneiss’ of the Theespruit Formation, which has been


116 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Fig. 1. Simplified geological map of southwestern part of the Barberton Mountain Land, Kaapvaal Craton, showing the location of Figs. 3 and 4 (rectangle), and of Fig. 7a, b, and c<br />

(dots). Inset shows location of the Barberton Greenstone Belt in the Kaapvaal Craton, within southern Africa. Age data for granitic rocks cited in text.<br />

interpreted to be ≤3458 Ma – and therefore younger than the Komati<br />

Formation – based on a population of zircons from a felsic ‘diamictite’<br />

within the Theespruit Formation (Armstrong et al., 1990). The<br />

kilometre-wide Komati schist zone (KSZ) that separates the two<br />

formations in the southwestern part of the Barberton Greenstone<br />

Belt was interpreted as a plate-boundary thrust (de Wit et al., 1992)<br />

(Fig. 1). In this model, the Komati Formation was interpreted to<br />

represent structurally thickened oceanic crust (“Jamestown Ophiolite<br />

Complex”; de Wit et al., 1987a), obducted onto a “tectono-stratigraphic<br />

assemblage” of Theespruit Formation rocks across the KSZ at<br />

ca. 3230 Ma (de Wit et al., 1983, 1987a, 1992; Kamo and Davis, 1994).<br />

Tectonic thickening was interpreted to have given rise to partial<br />

melting of the lower parts of the overthickened crust and the<br />

generation of granitoid rocks which were emplaced into active thrusts<br />

(de Wit et al., 1987b). Evidence of local high-pressure metamorphism<br />

from the granitoid domains flanking the belt has also been used to<br />

support models of continental collision (e.g., Dziggel et al., 2001, 2002;<br />

Diener et al., 2006; Stevens and Moyen, 2007).<br />

However, several papers have cast doubt on the thrust-accretion<br />

hypothesis for this part of the Barberton Greenstone Belt. First,<br />

detailed volcanological studies have shown that a section with<br />

proposed sheeted dykes within the Komati Formation, which de Wit<br />

et al. (1987a) suggested was part of an ophiolite, is rather a series of<br />

thick flows with primary spinifex textures and sills or massive flows<br />

(Cloete and King, 1991; Cloete, 1999; Dann, 2000). In addition, more<br />

detailed volcanological studies indicate that the Komati and Hooggenoeg<br />

Formations, the latter of which contains felsic volcanic rocks,<br />

represent largely intact, autochthonous volcanic stratigraphy, more<br />

than 6 km thick (e.g., Cloete, 1999; Dann, 2000; Dann and Grove,<br />

2007), and is thus very different from modern ophiolite sections. A<br />

gabbro interpreted as part of the ophiolite stratigraphy was shown to<br />

be some 100 Ma younger than the Komati volcanics (Kamo and Davis,<br />

1994). These data preclude an ophiolite interpretation for the lower<br />

Onverwacht Group.<br />

Second, Byerly et al. (1996) showed that the greenstones above the<br />

KSZ are a contiguous, progressively upward-younging succession<br />

(cf. Viljoen and Viljoen,1969) which formed over more than 200 Ma of<br />

Earth history. This implies that much of the succession is not a<br />

tectono-stratigraphic assemblage as implied in the thrust-accretion<br />

hypothesis (de Wit et al., 1987a) — at best, an accretion component is<br />

reduced to a single thrust (the KSZ).<br />

Third, Kisters and Anhaeusser (1995) showed that granitoid rocks<br />

flanking the greenstones show several features that are consistent with<br />

emplacement of granitoids as diapirs, thus indicating a more complicated<br />

story than proposed by the thrust-accretion hypothesis alone.<br />

The above, when combined with more detailed observations<br />

presented below, are incompatible with the proposed thrust-accretion<br />

hypothesis, as it has previously been described. The cumulative weight of<br />

these arguments, together with our new data, requires a re-assessment<br />

of the critical Tjakastad area, the results of which are presented here.<br />

The new results indicate that the published thrust-accretion model<br />

is unlikely to have been the primary tectonic mechanism responsible<br />

for development of the Palaeoarchaean rocks in the Tjakastad area.<br />

Instead, the structural features of this area are more consistent with<br />

formation through partial convective overturn of a dense succession of<br />

continually upward-younging, dominantly volcanic, supracrustal<br />

rocks and a structurally lower, more buoyant layer of intraplated<br />

granitoid sheets, following previously published diapiric models. The<br />

conclusions have important implications for early Archaean tectonic<br />

models, as the Barberton area has been widely cited as a key example<br />

in support of modern plate-tectonic processes operating in the ancient<br />

past (cf. Windley, 1995).


2. Regional geology and previous geochronology<br />

The Barberton Mountain Land part of the Kaapvaal Craton is a<br />

Palaeoarchaean (3.55–3.20 Ga) granite–greenstone terrain that forms<br />

part of the eastern Kaapvaal Craton (Brandl et al., 2006). The volcanosedimentary<br />

strata, known as the Barberton Supergroup, are preserved<br />

in the tightly folded, ENE trending Barberton Greenstone Belt (BGB),<br />

which is surrounded by granitoid rocks of a variety of ages (Lowe and<br />

Byerly,1999). The most recent overview of the BGB is by Lowe and Byerly<br />

(2007), who also provide an extensive literature survey.<br />

2.1. Barberton Supergroup<br />

The Barberton Supergroup comprises a basal Onverwacht Group of<br />

komatiitic, basaltic and felsic volcanic rocks, and conformably to<br />

unconformably overlying argillaceous and arenaceous metasedimen-<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

tary and minor felsic volcanic rocks of the Fig Tree and Moodies<br />

Groups, respectively (Fig. 2) (Viljoen and Viljoen, 1969; Brandl et al.,<br />

2006). The Onverwacht Group has been broadly subdivided into a<br />

lower ultramafic unit and an upper mafic to felsic unit, separated by a<br />

prominent unit of chert known as the Middle Marker (Viljoen and<br />

Viljoen, 1969). The lower and upper units are each composed of three<br />

formations. The structurally lowermost is the undated Sandspruit<br />

Formation composed of strongly deformed and metamorphosed<br />

mafic–ultramafic schists adjacent to, and infolded with, TTG rocks in<br />

the southwestern part of the BGB. Apparently conformably overlying<br />

the Sandspruit Formation is the heterolithic Theespruit Formation<br />

which regionally contains komatiites and basalts and characteristic<br />

units of metamorphosed felsic lava and tuff and which, in the<br />

Tjakastad area, has been interpreted to also include slivers of<br />

orthogneiss and diamictite (de Wit et al., 1983). The Theespruit<br />

rocks occur in two areas separated by younger granitoid intrusives,<br />

Fig. 2. Generalized stratigraphic section of Barberton Supergroup, showing available age constraints (see text for references). SF = Sandspruit Formation. Modified from Lowe and<br />

Byerly (2007) and Byerly et al. (1996); age data cited in text.<br />

117


118 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

namely around and east of the village of Tjakastad and partly in the<br />

Songimvelo Game Reserve (in this paper referred to as the Tjakastad<br />

area), and farther east in the area around Steynsdorp, close to the<br />

border between South Africa and Swaziland; here referred to as the<br />

Steynsdorp area (Fig. 1).<br />

Single zircon 207 Pb/ 206 Pb evaporation ages of 3548±3 to 3544±<br />

3 Ma for four separate samples of felsic tuff from the Theespruit<br />

Formation in the Steynsdorp area and flanking the ca. 3510 Ma<br />

Steynsdorp Pluton provide a well-constrained date of deposition for<br />

this unit (Kröner et al.,1996). Zircon xenocrysts in the Steynsdorp Pluton<br />

vary in age from 3553 to 3531 Ma, providing further evidence on the age<br />

of the underlying Theespruit and Sandspruit Formations (op cit.). In the<br />

Tjakastad area, along strike to the west, a sample from what was<br />

interpreted as a thin sliver of basement orthogneiss (but see below) in<br />

the Theespruit Formation yielded similar results of 3538±6 Ma and<br />

3538+4/−2 Ma, respectively, in two separate zircon dating efforts<br />

(Armstrong et al., 1990; Kamo and Davis, 1994). A felsic volcaniclastic<br />

rock from immediately above the so-called basement gneiss sliver, but<br />

still within the Theespruit Formation, yielded zircons with identical ages<br />

of 3538±10 Ma (Armstrong et al., 1990). However, this sample also<br />

contained a second population of zircons with an age of 3453±6 Ma,<br />

and it is this result that has been widely cited as a maximum age for the<br />

deposition of the Theespruit Formation (see below: Armstrong et al.,<br />

1990). An arkosic rock from the Sandspruit Formation to the west of our<br />

study area contained concordant zircons at 3521±6 Ma, with older<br />

grains up to ca. 3540 Ma (Dziggel et al., 2001, 2002).<br />

The Sandspruit and Theespruit Formations are separated from the<br />

overlying, lower-grade, Komati Formation by a broad zone of ductile<br />

shear, 600–800 m wide, and known as the Komati schist zone (KSZ). The<br />

overlying Komati Formation is composed of komatiites and interlayered<br />

basalts dated at between ca. 3490 Ma (Lopez-Martinez et al., 1992)and<br />

3481 Ma (Dann, 2000). The Middle Marker at the top of the lower<br />

ultramafic unit is a persistent horizon of silicified tuffaceous rocks dated<br />

at ca. 3472 Ma (Armstrong et al., 1990), the same age as quartz-feldspar<br />

porphyry dykes and sills intruding the Komati Formation (Kamo and<br />

Davis, 1994).<br />

Conformably overlying the Middle Marker is the Hooggenoeg<br />

Formation of tholeiitic basalt, komatiitic basalt, komatiite, and felsic<br />

volcanic flows and sills, dated at 3470 Ma (Byerly et al., 2002). Intruding<br />

the top of the Hooggenoeg Formation is a 3457–3438 Ma subvolcanic sill<br />

and related felsic lavas and volcaniclastic rocks (de Wit et al., 1987b;<br />

Kröner and Todt, 1988; Armstrong et al.,1990; Kröner et al., 1991; Byerly<br />

et al., 1996; de Vries et al., 2006). These rocks are locally eroded and<br />

conformably overlain by metamorphosed flow and volcaniclastic rocks<br />

of the Kromberg and Mendon Formations, dated at ≤3416 and<br />

N3298 Ma (Byerly et al., 1996; Lowe and Byerly, 2007).<br />

The Fig Tree Group conformably overlies the Onverwacht Group and<br />

is largely composed of clastic and pelitic metasedimentary rocks, but<br />

includes felsic volcanic and volcaniclastic rocks dated at between 3258<br />

and 3226 Ma, the same age as the Kaap Valley Tonalite and related<br />

intrusions to the north of the BGB (Kröner et al., 1991; Kamo and Davis,<br />

1994; Byerly et al., 1996; Lowe and Byerly, 2007). The Fig Tree Group is<br />

conformably to unconformably overlain by conglomeratic and arkosic<br />

units of the Moodies Group which, near the top of the section, contain<br />

several intra-formational unconformities indicative of syntectonic<br />

deposition in a contractional regime (Lamb, 1987). Deposition of the<br />

Moodies Group commenced about 3230 Ma and ended some time<br />

before 3110 Ma (Huebeck and Lowe, 1994). Synchronous folding and<br />

tilting of the entire Barberton Supergroup to near vertical occurred<br />

before 3216+2/−1 Ma, the age of the discordant and undeformed<br />

Dalmein Pluton (Fig. 1; Kamo and Davis, 1994).<br />

2.2. Granitoid rocks<br />

Tonalitic, trondhjemitic and granodioritic (TTG) rocks intrude and<br />

envelope the BGB and may be divided into five age groups. The oldest<br />

of these is the trondhjemitic Steynsdorp Pluton, dated at between<br />

3510 and 3502 Ma (Kröner et al., 1996). The second, more voluminous<br />

group of plutons is represented by linked, foliated to gneissic TTG<br />

rocks of the Stolzburg, Theespruit and Doornhoek plutons along the<br />

southwestern flank of the BGB. These plutons are not isolated<br />

intrusions but represent three domal culminations, or lobes, of a<br />

compositionally and geochronologically homogeneous granitoid layer<br />

separated by infolded, synformal septae of strongly deformed greenstones<br />

of the Sandspruit and Theespruit Formations (Fig. 1). These<br />

rocks display a characteristic dome-and-keel geometry in which<br />

radially-plunging lineations in the greenstone keels trend away from<br />

granitoid domes and increase in intensity and plunge towards the<br />

cores of the tightly folded greenstone keels, where the rocks are<br />

deformed into L-tectonite schists (Kisters and Anhaeusser, 1995).<br />

Anhaeusser (1984) and Kisters and Anhaeusser (1995) showed that<br />

the domal plutonic lobes were not formed by cross-folding but display<br />

characteristic features of diapiric emplacement. These plutonic rocks<br />

have been dated between 3469 and 3437 Ma, the same age as quartzfeldspar<br />

porphyry dykes intruding the Komati Formation and felsic<br />

volcanism in the Hooggenoeg Formation (Fig. 2: Kröner and Todt, 1988;<br />

Armstrong et al., 1990; Kamo and Davis, 1994; Kröner et al., 1996).<br />

Titanite from a 3476±10 Ma quartz-feldspar porphyry dyke cutting the<br />

Komati Formation in the Tjakastad area was dated at 3458±2 Ma and<br />

interpreted to represent the minimum age of intrusion of the dyke<br />

(Kamo and Davis, 1994). A second zircon generation from the ca.<br />

3445 Ma Stolzburg Pluton, considered by Kamo and Davis (1994) to be of<br />

metamorphic origin, is ca. 3237 Ma. Metamorphic titanite from the ca.<br />

3445 Ma Doornhoek and Stolzburg plutons was dated at ca. 3215 Ma and<br />

ca. 3201 Ma, respectively (Kamo and Davis, 1994).<br />

The third group of plutons is represented by the foliated Kaap<br />

Valley Tonalite and related granitoids along the northern flank of the<br />

BGB, dated at 3229±5 Ma (Tegtmeyer and Kröner, 1987), the same<br />

age as felsic volcanism in the Fig Tree Group (Kröner et al., 1991). The<br />

undeformed Dalmein Pluton was emplaced at 3216+2/−1 Ma<br />

(Kamo and Davis, 1994). Finally, a number of potassic granitoids, all<br />

∼3105 Ma in age, surround the BGB, among them the 3105±3 Ma<br />

Mpuluzi Batholith in the south (Kamo and Davis, 1994). These coarsegrained,<br />

porphyritic granites are largely undeformed.<br />

2.3. Metamorphism<br />

Viljoen and Viljoen (1969) identified an amphibolite-facies<br />

metamorphic aureole in greenstones around granitoids that surround<br />

the BGB. de Wit (1983) noted that greenstones of the Theespruit<br />

Formation adjacent to the Theespruit Pluton in the Tjakastad area<br />

were transformed into garnet amphibolites, and aluminous felsic<br />

schists contained lineated kyanite porphyroblasts.<br />

Cloete (1993, 1999) found that pressure estimates from fluid<br />

inclusion geobarometry in the Komati and Hooggenoeg Formations<br />

increased with stratigraphic depth to a maximum of 3.9 kbar and that<br />

the obtained values correspond to the lithological thickness of the<br />

overlying Barberton Supergroup, thus indicating that these rocks had<br />

only been subject to burial metamorphism. Temperature estimates<br />

using the plagioclase-amphibole thermometer vary from 320–420 °C<br />

in the Hooggenoeg Formation, to 490–530 °C in the Komati Formation.<br />

Rocks adjacent to the Theespruit Pluton were affected by a dynamic<br />

metamorphic overprint in addition to the static metamorphism, as<br />

evidenced by retrograde temperatures of 280 °C. Prograde P–T paths<br />

for Komati and Hooggenoeg Formation rocks have a slightly anticlockwise<br />

shape, from which Cloete (1993, 1999) concluded that<br />

metamorphism of the BGB was not analogous to that in Phanerozoic<br />

thrust-accretion belts.<br />

<strong>Van</strong> Vuren and Cloete (1995) studied the metamorphism of rocks<br />

in and around the Theespruit Pluton. These authors found that<br />

amphibolite-facies metamorphic conditions within the pluton were<br />

3–5 kbar and 470 °C. Marginal greenstones recrystallised under


conditions of 4.8–7 kbar, 500 °C. The highest metamorphic grade was<br />

found in the narrow septum of strongly deformed greenstones<br />

between the Theespruit and Stolzburg Plutons (Fig. 1), where<br />

garnet-clinopyroxene mafic granulites yielded P–T estimates of<br />

9 kbar and 700 °C.<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Fig. 3. Geology of the Tjakastad area of the Barberton Greenstone Belt, according to de Wit et al. (1983), showing inferred thrust faults (f–f), inferred thrust packages (double arrows 1<br />

through 6), and “diatexites” within the Theespruit Formation (modified slightly after Fig. 4 of de Wit et al., 1983). Localities A, D, E of inferred basement orthogneiss are discussed in<br />

the text. Zircon age data at locality D are from Armstrong et al. (1990) and show the results of two dated zircon populations from a sample of the ‘basement orthogneiss’ (but see text<br />

for discussion).<br />

Dziggel et al. (2002) recorded high-P, low-T metamorphic conditions<br />

of 8–11 kbar and 650–700 °C in remnants of the Theespruit<br />

and Sandspruit Formations within the western part of the Stolzburg<br />

Pluton. Two phases of titanite were dated from rocks of this area, at<br />

3418 ± 8 Ma and 3236±5 Ma, indicating two periods of metamorphism<br />

Fig. 4. Geological map of the Tjakastad area, based on data presented in this paper, showing the localities of dated felsic volcaniclastic samples and samples analysed for Sm–Nd, as discussed<br />

in text. Blue line with rectangles denotes transition from amphibolite- to greenschist-facies. Numbers inside boxes refer to other figures in this paper. DP = Doornhoek Pluton.<br />

119


120 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139


(and deformation) (Dziggel et al., 2001, 2002). Similar observations<br />

were made by Diener et al. (2006) for the area just west of Tjakastad,<br />

with prograde conditions of 5.5–6.3 kbar and 490 °C. Higher grade<br />

conditions in rocks closer to the granitoids gave peak P–T estimates of 7–<br />

7.7 kbar and 550 °C, with retrograde assemblages of 3.8 kbar and 543 °C,<br />

indicating nearly isothermal decompression. The estimates of peak<br />

metamorphic conditions indicate low geothermal gradients of 20 °C/km<br />

for the western part of the Barberton Mountain Land (Dziggel et al.,<br />

2002; Diener et al., 2006). A syn-tectonic tonalite dated at 3231±5 Ma<br />

from the Schapenburg schist belt in the southwestern part of the<br />

Barberton Mountain Land gives a maximum age of amphibolite-facies<br />

metamorphism in this area, as does a 3219 Ma date on metamorphic<br />

zircon from an amphibolite in the western Tjakastad schist belt, only 10–<br />

20 Ma after sediment deposition (Stevens et al., 2002; Kisters et al.,<br />

2003; Diener et al., 2006).<br />

3. Details of the proposed thrust-accretion model for the<br />

Tjakastad area<br />

As evident from the above overview, there is only one exception to a<br />

generally conformable interpretation for the Barberton Supergroup. This<br />

occurs in the Tjakastad area, where purported evidence of large-scale<br />

tectonic inversion of older-over-younger rocks has been documented. de<br />

Wit et al. (1983) suggested that the Theespruit Formation in the<br />

Tjakastad area was a tectono-stratigraphic assemblage of tectonically<br />

intercalated mafic volcanic rocks, metasedimentary diamictite and<br />

associated quartz-sericite schist, quartz-feldspar porphyries, and slivers<br />

of banded granitoid basement gneiss. In this model, the Theespruit<br />

Formation was “…most readily explained as a set of imbricate slices…<br />

that represents an inclined (oblique?) section through an imbrication<br />

zone beneath a major thrust … represented by the Komati schist zone…”<br />

(de Wit et al., 1983, p. 27: word in italics added).<br />

The most critical field evidence in support of a tectono-stratigraphic<br />

interpretation for the Theespruit Formation in the Tjakastad<br />

area was the recognition of tectonic slivers and cobbles of inferred<br />

basement orthogneiss (de Wit et al., 1983). Orthogneiss slivers were<br />

recognised in three distinct settings: (1) as a 500 m long wedge within<br />

quartz-sericite schist, diamictite, and quartz-plagioclase porphyries<br />

within the KSZ (locality E on Fig. 3); (2) as tectonic slivers within the<br />

Theespruit Formation, in one location being unconformably overlain<br />

by sedimentary diamictite (locality D on Fig. 3); (3) as cobbles in<br />

metasedimentary diamictite (locality A on Fig. 3). In setting 1, de Wit<br />

et al. (1983) suggested that the orthogneiss had experienced an<br />

earlier deformation history as indicated by isoclinal folds of<br />

metamorphic layering, the axial planes of which were oriented at a<br />

high angle to the regional foliation (Fig. 6b and c of de Wit et al.,<br />

1983). Similarly, cobbles of orthogneiss within diamictite in setting 3<br />

were interpreted to exhibit a pre-existing gneissosity oriented at a<br />

high angle to both bedding and the regional upright cleavage in the<br />

matrix diamictite.<br />

That thrusting had occurred in the Tjakastad area was interpreted<br />

from the penetratively deformed nature of the rocks, the presence of<br />

faults that offset the stratigraphy and were interpreted to indicate a<br />

reverse sense of displacement, the presence of tectonic contacts as<br />

indicated by changes in the orientation and intensity of mineral<br />

elongation lineations across unit contacts, the occurrence of a southfacing<br />

bedding top direction within the KSZ, and along-strike differences<br />

in the thickness and components of the “tectono-stratigraphy”.<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Supporting a tectono-stratigraphic interpretation for the Theespruit<br />

Formation were zircon ages of 3538+4/−2 Ma and 3538 ±<br />

6 Ma (Armstrong et al., 1990; Kamo and Davis, 1994) for the<br />

“basement gneiss sliver” at locality D. At the time these results<br />

were obtained, they were the oldest recorded from the BGB and<br />

apparently confirmed a basement interpretation for these rocks relative<br />

to the overlying Komati Formation and younger conformable stratigraphy.<br />

Furthermore, euhedral to subhedral zircons with well-developed<br />

facets from the unconformably overlying ‘diamictite’ (which<br />

Armstrong et al., 1990 interpreted as a volcaniclastic rock), about 2 m<br />

above the gneiss wedge, yielded two distinct age populations: an older<br />

group at 3531±10 Ma, indistinguishable from the “gneiss sliver”,anda<br />

younger group at 3453±6 Ma. Although “…core and rim structures are<br />

common…” (Armstrong et al., 1990, p. 94) in zircons from these<br />

greenschist- to amphibolite-facies schists, the younger ages were<br />

interpreted to be derived from detrital grains and thus to indicate a<br />

maximum age of deposition for the Theespruit Formation. These data<br />

apparently confirmed both a younger age of the ‘diamictite’ relative to<br />

the gneiss sliver, and that the KSZ represents a major thrust across which<br />

older rocks of the Komati Formation (3490 Ma) were tectonically<br />

emplaced over the younger rocks of the Theespruit Formation<br />

(b3453 Ma).<br />

3.1. Controversial or unexplained elements of the previously proposed<br />

model<br />

Several features of both the geological interpretation by de Wit<br />

et al. (1983; Fig. 3) and the geochronology (Armstrong et al., 1990)do<br />

not appear easily reconcilable with a thrust-accretion origin for the<br />

rocks in this region:<br />

1) Why do the identified faults within the Theespruit Formation (f–f<br />

on Fig. 3) extend the lithological layering, rather than duplicate it<br />

as required for a thrust-accretion model?<br />

2) What is the sense of shear across the KSZ? If, as proposed, the KSZ<br />

represents a crustal-scale thrust, then it should contain evidence<br />

of reverse (Komati-over-Theespruit Formation) displacement.<br />

3) What is the origin of the ‘diamictites’ and how do they differ from<br />

volcaniclastic rocks or igneous intrusive breccias described<br />

elsewhere in the map area, or to felsic volcanic schist and metaagglomerate<br />

along strike to the east and west as described by<br />

Kisters and Anhaeusser (1995) and Kröner et al. (1996)?<br />

4) Why does a large unit of quartz-plagioclase porphyry with<br />

intrusion breccia textures pass laterally along strike into<br />

‘diamictite’ (just east of locality E in Fig. 3)?<br />

5) Why is there a discrepancy in the interpreted origin of felsic<br />

schists of the Theespruit Formation between the Tjakastad area<br />

(sedimentary diamictite and orthogneiss: de Wit et al., 1983) and<br />

those areas along strike to the west?<br />

6) What is the relationship between the proposed orthogneiss<br />

wedge (locality E in Fig. 3) and adjacent ‘diamictite’ and quartzplagioclase<br />

porphyries that surround, or structurally overlie it;<br />

does the wedge represent the limb of an inverted nappe, or is it<br />

located within a tectonic mélange?<br />

7) What role, if any, did the ca. 3450 Ma granitoid rocks play in the<br />

deformation of the Tjakastad area? Is it the same as for ca.<br />

3230 Ma granitoid rocks higher up in the stratigraphy which were<br />

emplaced during regional thrusting (cf. de Wit et al., 1987b), or do<br />

Fig. 5. Outcrop and thin section features of Theespruit Formation rocks: a) Well-preserved pillow basalts near the base of the formation. View is to the west-northwest, and drainage<br />

cavities (arrowed) and pillow tails (i.e. at hammer head) indicate way-up to north-northeast; b) Well-rounded cobble of fine-grained siliceous metasediment in foliated and lineated<br />

felsic agglomerate with stretched lapilli (not visible in photo): clast shows coarse layering defined by white and light grey colour banding; c) Cross-sectional view of crude layering<br />

parallel to bedding defined by thin, irregular sills of darker grey, weakly plagioclase feldspar-porphyritic dacite in fine-grained felsic volcaniclastic siltstone; d) Cross-sectional view of<br />

massive, coarse felsic volcaniclastic rock (agglomerate) with irregular, ragged-edged fiamme of feldspar-phyric, devitrified rhyolite glass (arrow) and well rounded clasts of chert; e)<br />

View, looking down on a horizontal surface, of alkali trachytic breccia: note angular, unstrained nature of clasts in this view; f) View, looking horizontally, on alkali trachytic breccia,<br />

showing strongly downdip lineated nature of fragments; g) Cross-polarised thin section view of alkali trachyte fragment in breccia, showing Carlsbad-twinned subhedral K-feldspar<br />

phenocryst in fine matrix of flow-aligned albite. Pen knife in b)–f) is 10 cm. All locations are shown in Fig. 4.<br />

121


122 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139


they display evidence of diapiric emplacement as previously<br />

proposed (Anhaeusser, 1984; Kisters and Anhaeusser, 1995)?<br />

8) What is the significance of high-grade metamorphic minerals<br />

(kyanite in pelitic schists and garnet in amphibolite) within the<br />

Theespruit Formation? Is the variation in metamorphic grade across<br />

the Tjakastad area consistent with, and exclusive to, a thrustaccretion<br />

model of formation, or can it be explained in other ways?<br />

9) What is the significance that the ‘basement gneiss sliver’ in the<br />

Tjakastad area is identical in age to felsic volcanic schists of the<br />

Theespruit Formation along strike to the southeast?<br />

10) What is the significance of the younger population of zircons in the<br />

dated, amphibolite- to greenschist-facies ‘diamictite’ (ca. 3453 Ma)<br />

which display core–rim overgrowth relationships and are identical<br />

in age to titanite in a quartz-feldspar porphyry sill cutting the<br />

Komati Formation (ca. 3458 Ma) and to widespread TTG plutonism<br />

(3470–3437 Ma) and felsic volcanism (3452–3438 Ma)?<br />

4. Results of field mapping and petrography<br />

Field mapping was undertaken in order to investigate the apparent<br />

inconsistencies in the thrust-accretion model for the Tjakastad area. The<br />

results are presented in Fig. 4, and below, under subheadings which<br />

individually address each of the apparent inconsistencies. The map<br />

differs in several key ways from that published by de Wit et al. (1983),<br />

particularly by way of different interpretations for the origin of several<br />

rock units in the Theespruit Formation, a stratigraphic interpretation for<br />

the Theespruit Formation, the nature of contact relationships between<br />

units, the position of the northeastern part of the Komati Fault, and the<br />

kinematics of faulting/shearing across the Tjakastad area.<br />

4.1. Lithology<br />

4.1.1. General stratigraphy<br />

The map area spans the transition from amphibolite-facies schists<br />

of the Theespruit Formation, which lie adjacent to the trondhjemitic<br />

Theespruit Pluton, to lower greenschist-facies komatiites of the<br />

Komati Formation across the high-strain KSZ (Fig. 4). The best and<br />

most continuous exposures are in a small streambed in the hills east of<br />

Tjakastad. The Theespruit Formation consists of lower pillowed and<br />

ocellar metabasalts (Fig. 5a), three compositionally and texturally<br />

distinct horizons of felsic schists and metachert, and interlayered<br />

mafic–ultramafic schists of indeterminate extrusive or intrusive<br />

origin. Way-up indicators of bedding are defined by graded bedding<br />

in felsic volcaniclastic layers and pillow structures in basalts. These<br />

consistently face to the northeast, away from the Theespruit Pluton<br />

(Fig. 4), except on the limbs of small-scale tight folds of bedding.<br />

Lithologic units of the Theespruit Formation continue into the KSZ, but<br />

then change across the Komati Fault into the base of the Komati<br />

Formation, which includes the type area of bladed olivine spinifextextured<br />

komatiites, massive ultramafic rocks, talc-serpentine schist,<br />

and intrusive quartz-feldspar porphyry.<br />

4.1.2. Felsic volcanic rocks<br />

Three compositionally and texturally distinct horizons of felsic<br />

rock occur in the Theespruit Formation (Fig. 4). The structurally and<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

stratigraphically lowest felsic horizon is ≤65 m thick. It consists of<br />

irregularly layered felsic volcanic schist, grey-and-white layered, finegrained,<br />

cherty metasediment (siltstone/mudstone), kyanite-bearing<br />

quartz-sericite schist derived from aluminous greywacke, subordinate<br />

felsic tuff, and some massive felsic volcanic rocks. This horizon<br />

contains the dated unconformity locality that is discussed in detail in<br />

Section 4.1.4.2. The kyanite–andalusite schists contain faint relict<br />

bedding and lineated kyanite porphyroblasts that locally retain fresh<br />

cores of andalusite (see Section 4.3). The cherty metasediment is<br />

characterised by irregular, wispy layering and common chaotic folds<br />

that were possibly developed during, or soon after sedimentation, as<br />

they contain no axial planar foliation and have highly non-cylindrical<br />

fold axes, in contrast to folds of clearly tectonic origin elsewhere.<br />

Locally, cherty metasediment occurs as mis-oriented blocks in<br />

massive, fine-grained quartz-sericite rock (Fig. 5b). Nearby, the<br />

layered metasediment has an irregular contact with massive, weakly<br />

feldspar-phyric quartz-sericite rock that cuts bedding within the<br />

metasediment (Fig. 5c). These textures are interpreted as indicative of<br />

coeval sedimentation of felsic volcaniclastic detritus and felsic<br />

magmatism, the latter formed as either a felsic volcanic flow or a sill<br />

emplaced directly below the sediment–water interface.<br />

The middle horizon of felsic schist, which de Wit et al. (1983)<br />

interpreted to contain extensive sedimentary diamictite with local clasts<br />

of basement granitoid orthogneiss, is similar in thickness to the lower<br />

horizon, but is texturally heterogeneous along strike. The best exposures<br />

of the coarse clastic rocks are in the small streambed mentioned above,<br />

where they comprise almost the entire thickness of the horizon. This<br />

unit contains subrounded to subangular, pebble to boulder size clasts of<br />

abundant grey to black chert, less abundant white and grey banded<br />

chert, clastic metasediment, and fine-grained felsic volcanic rock, and<br />

rare mafic volcanicrockandgraniteinafine grained quartz-sericite<br />

matrix (Fig. 5d). We interpret these rocks to be proximal felsic volcanic<br />

pyroclastic agglomerates and not sedimentary diamictite, because:<br />

1. Well-preserved parts of the unit contain common, irregular,<br />

ragged-edged fiamme of feldspar-phyric, devitrified rhyolite glass<br />

(arrow in Fig. 5d), whose delicate shapes could not have survived<br />

transport by water together with the rounded boulders of chert, as<br />

they would have almost instantly been disaggregated.<br />

2. The unit locally contains elliptical bombs of vesicular to pumiceous<br />

felsic volcanic rock in an otherwise massive, fine-grained felsic<br />

matrix that lacks bedding.<br />

3. The coarsely fragmental rocks lack bedding, are unsorted, and<br />

laterally pinch and swell along strike, grading into felsic lavas and/<br />

or fine-grained volcaniclastic sediment.<br />

4. In thin section, the matrix is seen to be generally fine-grained with<br />

no evidence of original detrital grains, and it contains elliptical to<br />

irregular-shaped patches of slightly coarser quartz–biotite–plagioclase<br />

with serrate, gradational contacts with the matrix, that are<br />

interpreted to represent fiamme.<br />

The stratigraphically highest felsic horizon locally displays metreto<br />

centimetre-scale bedding and is characterised by layers of quartzphyric<br />

felsic porphyry with subordinate felsic schist, fine-grained<br />

quartz-biotite metasedimentary rock of possible volcaniclastic origin,<br />

and minor felsic volcanic breccia.<br />

Fig. 6. a) Representative outcrop example of the quartz-feldspar porphyry intrusion at ‘basement orthogneiss’ locality of de Wit et al. (1983) (locality E in Fig. 3 of this paper),<br />

showing medium-grained, equigranular texture and crude igneous layering; b) Cross-polarised thin section view of a), showing unstrained, euhedral plagioclase feldspar<br />

phenocrysts in weakly foliated matrix of fine-grained quartz and feldspar; c) boulder of unstrained quartz-feldspar porphyry – previously interpreted as basement orthogneiss (de<br />

Wit et al., 1983) – with well defined igneous flow banding that outlines isoclinal folds; d) Closeup view of hinge region of isoclinal igneous flow fold in c), showing unstrained,<br />

bipryramidal quartz phenocrysts (arrows); e) Cross-polarised thin section view of rock shown in parts c) and d), showing unstrained, euhedral bipyramidal quartz phenocrysts in an<br />

undeformed matrix of fine-grained quartz and feldspar: unstrained, euhedral plagioclase phenocrysts are also visible in this thin section, but are not shown; f) Outcrop view, looking<br />

south-southeast along strike, of the ‘unconformity locality’ of de Wit et al. (1983) (locality D in Fig. 3 of this paper), showing foliated quartz-sericite schist (right side, under penknife)<br />

and breccia-textured unit (left side, oblique view onto bedding/foliation plane); g) Plane polarised light thin section view of homogeneous quartz-sericite schist from the<br />

‘unconformity locality’, showing tourmaline crystals and chloritoid porphyroblasts; h) View looking west-southwest onto steeply dipping bedding/foliation surface of felsic volcanic<br />

breccia unit shown on the left side of (f), showing angular nature of felsic clasts in this volcaniclastic breccia unit and well-developed stretching lineation at this outcrop. Pen knife in<br />

a), c) and f) is 10 cm long. All locations are shown in Fig. 4.<br />

123


124 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Fig. 7. a) Large, unstrained amphibolite-facies pillow basalt in a 100–200 m xenolith from core of Theespruit Pluton (see Fig. 1 for location): pillows face to the southeast; b) Vertical<br />

cross-sectional view of strongly lineated, amphibolite-facies pillows within narrow greenstone septum between lobes of the Theespruit Pluton; c) Sheeted, heterogeneous margin of<br />

Theespruit Pluton, which contrasts with the unstrained, homogeneous core of the pluton; d) Rodded L-tectonite felsic volcanic schists from the KSZ, used as building materials;<br />

e) Plane polarised light view of a vertically-oriented thin section, looking southeast, showing SW-side-up displacement in sheared metabasalt from the KSZ (NE = northeast, SW =<br />

southwest); f) Plane polarised thin section view of aluminous felsic schist of Theespruit Formation, showing kyanite (Ky = dark grey) overgrowing andalusite (And = light grey, with<br />

well develop cleavage) along extension cracks, indicative of an increase in pressure during metamorphism and the development of linear fabrics in these rocks. Locations are shown<br />

in Figs. 1 and 4.


Two thin units of strongly sheared, texturally variable felsic schist<br />

in the western part of the KSZ are similar to the lower felsic unit in the<br />

Theespruit Formation. The topmost unit contains fine-grained tuff,<br />

grey chert, and fine-grained quartzite. In the southern-central part of<br />

the KSZ, a ∼100 m thick unit of felsic schist that de Wit et al. (1983)<br />

interpreted as sedimentary diamictite varies, from base to top, from a<br />

massive rock with plagioclase-phyric texture, a middle unit of felsic<br />

volcanic breccia with elliptical, ragged-shaped inclusions of carbonate-altered<br />

felsic material, and ≤50 m of grey and white layered<br />

chert.<br />

In the eastern part of the KSZ in the study area, de Wit et al. (1983)<br />

described an intrusion breccia within the core of a quartz-plagioclase<br />

porphyry intrusion that they interpreted to pass along strike<br />

westwards into sedimentary diamictite (see Fig. 3). However, it was<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Fig. 8. a) Structural map of the Tjakastad area, showing stretching lineations, kinematic indicators, and areas of high strain within the Komati schist zone (KSZ). Locations of transects<br />

in b) are also indicated; b) Transects across the Theespruit Formation and KSZ, showing variations in orientation of stretching lineations (see text for discussion).<br />

found that the rocks in this area comprise two distinct units (see<br />

Fig. 4): 1) an intrusive unit of quartz-feldspar porphyry to the north<br />

(described below); and 2) a fine-grained, highly siliceous felsic<br />

volcanic rock with closely packed, monomict, angular breccia<br />

fragments (Fig. 5e). Detailed outcrop mapping showed that parts of<br />

this unit display crude layering on a metre scale, with lenses or layers<br />

of massive rock alternating with layers of homogeneous, monomict,<br />

angular breccia, and other layers of less well sorted breccia with<br />

irregular-shaped blocks of massive rock up to a metre in size. Overall,<br />

bedding becomes more clearly defined at a finer-scale to the north,<br />

suggestive of graded bedding and a top to the north facing<br />

direction. Despite excellent preservation of igneous textures on<br />

horizontal outcrops, the rocks have been deformed into verticallyplunging<br />

L-tectonites (Fig. 5f). In thin section, the massive unit at this<br />

125


126 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

locality is characterised by euhedral K-feldspar phenocrysts with<br />

Carlsbad twinning in a matrix of fine, flow-aligned albite with<br />

trachytic texture (Fig. 5g). Quartz phenocrysts are lacking in this<br />

lithology, and thus the rock is classified as an alkali trachyte breccia.<br />

The textural uniqueness of each of the felsic horizons in the<br />

Theespruit Formation and the KSZ indicate that they do not represent<br />

a tectonically-duplicated series of thrust slices as suggested by de Wit<br />

et al. (1983). Rather, the above data indicate distinct lithological<br />

variations across strike, and the presence of consistent bedding top<br />

indicators to the north in both pillowed mafic volcanics and felsic<br />

volcanic/volcaniclastic horizons suggests that the rocks of the<br />

Theespruit Formation in the Tjakastad area, through into the KSZ,<br />

represent a strained, but otherwise intact volcano-sedimentary<br />

succession. This hypothesis is discussed more fully in Section 4.2.<br />

4.1.3. Granitic dyke in the Theespruit Formation<br />

A7–8 m wide sheared granite dyke occurs at one locality in the small<br />

stream bed to east of Tjakastad, about 100 m north of the Theespruit<br />

Formation/Theespruit Pluton contact. The foliation in this granite is<br />

parallel to that in the adjacent Theespruit schists, and the dyke could not<br />

be followed along strike due to poor exposure. The dyke is similar in<br />

appearance to granitic to aplitic dykes and veins emplaced into the<br />

Theespruit and Stolzburg plutons in the Tjakastad region. The dyke was<br />

sampled for zircon dating, and the results are reported below.<br />

4.1.4. “Granitoid gneiss”<br />

The detailed map of the Theespruit area by de Wit et al. (1983)<br />

shows the precise locations where exposures of granitoid gneiss were<br />

thought to occur in three distinct tectonic settings. These included a<br />

large wedge of orthogneiss in the northeastern part of the KSZ, tectonic<br />

slivers in the lower part of the Theespruit Formation, and cobbles<br />

within sedimentary “diamictite”. Re-examination of these easily<br />

recognisable sites and detailed petrographic studies of collected<br />

samples indicate that granitoid rocks were misidentified at these<br />

localities and that no such rocks occur in the Theespruit area. Instead,<br />

these rocks are re-interpreted as either quartz-plagioclase porphyries<br />

or volcaniclastic metasedimentary rocks, for reasons detailed below.<br />

4.1.4.1. Northeastern wedge of granitoid gneiss. The wedge of<br />

‘granitoid gneiss’ in the northeastern part of the map area (locality E in<br />

Fig. 3) consists of a homogeneous, medium-grained rock of trondhjemitic<br />

composition with characteristic euhedral, bipyramidal quartz and<br />

plagioclase phenocrysts (≤3 mm) scattered throughout a finer-grained<br />

quartzo-feldspathic matrix (Fig. 6a). In thin section, igneous textures are<br />

well preserved, consisting of subhedral to euhedral quartz and<br />

plagioclase phenocrysts in a fine-grained quartzo-feldspathic matrix<br />

that is overprinted by radiating sheaths of riebeckite needles (Fig. 6b).<br />

In contrast to the map of de Wit et al. (1983), this unit occurs<br />

outside, to the north of the KSZ, the northern margin of which is clearly<br />

marked by a discrete, fault-filled, 2 m wide quartz vein (Fig. 4). The<br />

rocks contain none of the strong LNS tectonite fabric characteristic of<br />

the KSZ (see Section 4.2 below); they are almost completely<br />

unstrained except for a very local, weak foliation developed within<br />

10 m of the KSZ. These rocks contain no leucosome veins, migmatitic<br />

textures, or gneissic layering with palaeosome–leucosome relations<br />

to indicate that they were derived through metamorphic segregation<br />

processes related to an earlier thermo-tectonic event, as previously<br />

interpreted (de Wit et al., 1983). The compositional layering which<br />

does occur in this unit is only locally developed in the southern part of<br />

the unit, and in only one outcrop does it outline isoclinal folds (Fig.<br />

6c; same as Fig. 6b and c of de Wit et al., 1983). This instantly<br />

recognisable outcrop with the isoclinally folded layering contains<br />

undeformed, euhedral bipyramidal quartz and plagioclase phenocrysts<br />

throughout, even across the hinge regions and limbs of the<br />

folds (Fig. 6d, e). The presence of these igneous phenocrysts indicate<br />

that the folded compositional layering represents igneous flow<br />

banding rather than a metamorphic segregation, and that the rock<br />

is thus a quartz-plagioclase porphyry intrusion, similar to the type<br />

dated by Kamo and Davis (1994) at ca. 3470 Ma; it is not a basement<br />

orthogneiss.<br />

Finally, the “evidence” cited by de Wit et al. (1983) that the<br />

isoclinally folded ‘granitoid gneiss’ at this locality had experienced an<br />

earlier phase of deformation on the basis that the folds were oriented<br />

at a high angle to the regional foliation, cannot be substantiated since<br />

the rock containing the folds is a misoriented boulder that has partly<br />

slid down the side of a hill.<br />

4.1.4.2. ‘Granitoid gneiss’ at the ‘unconformity’ locality. The ‘unconformity’<br />

locality described by de Wit et al. (1983: locality D in Fig. 3)is<br />

underlain by highly strained schists over a strike length of 10 m and<br />

30 cm across. The interpreted unconformity in this outcrop consists of<br />

a lower unit (‘orthogneiss sliver’ of de Wit et al., 1983), 28 cm wide, of<br />

millimetre- to centimetre-scale interbedded quartz-sericite, quartzsericite-tourmaline-chloritoid,<br />

quartz-chlorite, sericite-quartz-chlorite-actinolite<br />

schist (Fig. 6f, g). The overlying schist unit (‘diamictite’ of<br />

de Wit et al., 1983) is 2 cm wide and displays a distinctive breccia of<br />

strongly lineated felsic fragments (L=62°→107°) on foliation surfaces<br />

(Fig. 6h). The fragments are composed of sericite (95%)-chloritetitanite<br />

up to 120×40×2 mm in size (X:Y:Z axes) within a matrix of<br />

slightly darker quartz-sericite-biotite-chlorite schist. These rocks are<br />

overlain by quartz-sericite schist, finely layered grey and white<br />

Table 1<br />

Location and field relations of samples for zircon dating and Sm–Nd isotopic analysis.<br />

Sample Location Rock type Field relationship<br />

BA56 S26°00′15.3″<br />

E30°50′24.3″<br />

Felsic tuff Bottom of 50 m wide layer<br />

BA57 S26°00′04.1″<br />

E30°50′15.4″<br />

Felsic tuff Top of 50 m wide layer<br />

BA62 S26°00′03.8″<br />

E30°50′00.2″<br />

Felsic tuff Close to Theespr. tonalite contact<br />

BA65 S26°00′04.3″ Granite Sheared dyke cutting Theespruit<br />

E30°49′59.8″<br />

Fm.<br />

BA66 S26°00′00.7″<br />

E30°50′06.9″<br />

Amphibolite Interlayered with felsic tuff<br />

BA69 S25°59′54.6″<br />

E30°50′05.3″<br />

Amphibolite Interlayered with felsic tuff<br />

BA70 S25°59′54.7″<br />

E30°50′04.9″<br />

Felsic tuff 30 m wide layer<br />

BA71 S25°59′54.3″<br />

E30°50′05.6″<br />

Felsic tuff 6 m wide layer<br />

BA73 S25°59′41.0″<br />

E30°50′12.2″<br />

Felsic tuff 20 m wide layer, very fresh<br />

BA74 S25°59′40.7″<br />

E30°50′12.6″<br />

Basaltic komat. 10 m wide pillowed layer<br />

BA75 S26°00′33.8″<br />

E30°51′11.7″<br />

Gabbro Sill in felsic tuff<br />

BA79 S26°01′34.5″<br />

E30°53′36.8″<br />

Felsic tuff 60 m wide layer<br />

BA84 S26°00′32.3″<br />

E30°53′31.5″<br />

Felsic tuff Interlayered with chert<br />

BA85 S26°00′32.2″<br />

E30°53′31.5″<br />

Felsic tuff Layer 10 mN of BA 85<br />

BA 104 S25°59′30.5″<br />

E30°50′08.2″<br />

fsp-porphyry 25 thick layer in felsic tuff<br />

BA 105 S25°59′35.8″<br />

E30°49′59.4″<br />

fsp-porphyry ca. 25 m wide layer, fresh<br />

BA 107 S26°00′52.2″<br />

E30°51′07.5″<br />

Felsic tuff Closure of isoclinal fold<br />

BA 108 S26°03′00.9″<br />

E30°53′01.8″<br />

Felsic agglom. ca. 50 m thick layer, fresh<br />

BA 109 S26°02′21.8″<br />

E30°53′34.1″<br />

Felsic tuff ca. 80 m wide layer, bio-rich<br />

BA 110 S26°02′04.7″<br />

E30°53′34.1″<br />

Felsic tuff Same layer as BA 109, bio-free<br />

160932 S25°59′41.9″ Felsic volcanic Thick, coarse felsic volcaniclastic<br />

E30°49′52.7″<br />

conglomerate<br />

All samples listed in the table are plotted on Fig. 4.


Table 2<br />

Isotopic data from evaporation of single zircons from rocks of the Theespruit Formation and one intrusive granite, Tjakastad–Songimvelo area, Barberton Greenstone Belt.<br />

Sample number Zircon colour and<br />

morphology<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Grain # Mass scans 1<br />

Evaporation<br />

temp. in °C<br />

Mean 207 /Pb/ 206 Pb ratio 2<br />

and 2-σm error<br />

207 Pb/ 206 Pb age<br />

and 2-σm error<br />

BA 56 Clear to yellowish, 1 109 1599 0.310606±64 3523.9±0.3<br />

Long-prismatic, 2 150 1601 0.310751±67 3524.6±0.3<br />

Near-idiomorphic 3 83 1598 0.310486±139 3523.3±0.7<br />

4 102 1599 0.310533±93 3523.6±0.5<br />

Mean of grains 1–4 1–4 444 0.310593 ±44 3524.0 ±0.2<br />

BA 57 Yellow-brown, 1 128 1603 0.310214±148 3522.0±0.7<br />

Long-prismatic, 2 110 1602 0.310412±156 3523.0±0.8<br />

Idiomorphic 3 82 1599 0.310337±148 3522.6±0.7<br />

Mean of grains 1–3 1–3 320 0.310336±80 3522.6±0.4<br />

BA 62 Clear to light 1 98 1598 0.310602±110 3523.9±0.5<br />

Yellow-brown, 2 111 1598 0.310678±118 3524.3±0.6<br />

Long-prismatic, 3 103 1600 0.310546 ±122 3523.6±0.6<br />

and stubby 4 52 1601 0.310623±164 3524.0±0.8<br />

Mean of grains 1–4 1–4 364 0.310612±62 3524.0 ±0.3<br />

BA 65 Clear to yellow- 1 64 1602 0.288251±126 3408.2 ±0.7<br />

Brown, long- 2 64 1600 0.288265±81 3408.3 ±0.4<br />

Prismatic, idiom. 3 65 1599 0.288197±65 3407.9±0.4<br />

Mean of grains 1–3 1–3 193 0.288238±54 3408.1±0.3<br />

BA 70 Clear, thin, long- 1 66 1595 0.310381 ±86 3522.8±0.4<br />

Prismatic, 2 88 1599 0.310363 ±64 3522.7±0.3<br />

Slightly rounded 3 66 1598 0.310962 ±73 3525.7±0.4<br />

Mean of grains 1–3 1–3 220 0.310548±56 3523.6±0.3<br />

BA 71 Dark red-brown, 1 75 1608 0.311461±51 3528.2±0.3<br />

Long-prismatic, 2 58 1604 0.311338±69 3527.6±0.3<br />

Idiomorphic to 3 51 1602 0.311498±91 3528.4±0.5<br />

Slightly rounded 4 68 1600 0.311572±65 3528.7±0.3<br />

Mean of grains 1–4 1–4 252 0.311470±35 3528.2±0.2<br />

BA 73 Clear to yellowish, 1 52 1599 0.310174±92 3521.8±0.5<br />

Long-prismatic, 2 66 1601 0.310412±69 3523.0±0.3<br />

Slightly rounded 3 102 1599 0.310164 ±88 3521.7±0.4<br />

Mean of grains 1–3 1–3 220 0.310241±53 3522.1±0.3<br />

BA 84 Clear, thin, long- 1 95 1595 0.312116±87 3531.4±0.4<br />

Prismatic, 2 44 1599 0.312437±87 3533.0±0.4<br />

Slightly rounded 3 82 1598 0.312364 ±148 3532.6±0.7<br />

Mean of grains 1–3 1–3 221 0.312273±71 3532.2±0.4<br />

BA 85 Clear, 1 74 1608 0.312382±125 3532.7±0.6<br />

Long-prismatic, 2 115 1614 0.312687±102 3534.2±0.5<br />

Well rounded 3 96 1618 0.312393±110 3532.8±0.5<br />

Terminations 4 95 1612 0.312448 ±136 3533.1±0.5<br />

5 97 1620 0.312384±59 3532.7±0.3<br />

6 104 1614 0.312467±70 3533.1±0.3<br />

Mean of grains 1–6 1–6 581 0.312468±43 3533.2 ±0.2<br />

BA 105 Dark grey, round, 1 128 1597 0.311002 ±37 3525.9±0.2<br />

Detrital 2 82 1594 0.310763 ±46 3524.7±0.2<br />

Grains 3 92 1595 0.310867±58 3525.2±0.3<br />

Mean of grains 1–3 1–3 302 0.310896 ±43 3525.4±0.2<br />

BA 107 Grey-brown, 1 127 1600 0.312022±48 3531.0±0.2<br />

Long-prismatic, 2 87 1599 0.311711±64 3529.4±0.3<br />

Well rounded 3 54 1601 0.311789±85 3529.8±0.4<br />

Terminations 4 114 1600 0.312024±52 3531.0±0.3<br />

Mean of grains 1–4 1–4 382 0.311919±37 3530.4±0.2<br />

BA 108 Clear, 1 110 1608 0.311212±45 3526.9±0.2<br />

Long-prismatic, 2 93 1614 0.311249±46 3527.1±0.2<br />

Well rounded 3 142 1618 0.311146±35 3526.6±0.2<br />

Terminations 4 204 1612 0.311154±27 3526.7±0.1<br />

5 107 1620 0.311346±50 3527.6±0.2<br />

Mean of grains 1–5 1–5 656 0.311207±18 ⁎3526.9±0.2<br />

BA 109 Clear, 1 104 1608 0.311834±74 3530.0±0.4<br />

Long-prismatic, 2 74 1614 0.311789±99 3529.8±0.5<br />

Well rounded 3 114 1618 0.311893±62 3530.3±0.3<br />

Terminations 4 64 1612 0.311882±99 3530.3±0.5<br />

Mean of grains 1–4 1–4356 0.311852±40 3530.1±0.2<br />

(continued on next page)<br />

127


128 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Table 2 (continued)<br />

Sample number Zircon colour and<br />

morphology<br />

tuffaceous sandstone/siltstone, and wispy to irregularly layered felsic<br />

schist. Layering in the lowermost felsic unit (the so-called ‘granitoid<br />

gneiss sliver’ of de Wit et al., 1983) is parallel to the contact with the<br />

overlying brecciated unit, such that there is no evidence of an angular<br />

relationship to indicate the presence of an unconformity. Nor is there<br />

any textural evidence in the lower schist unit that metamorphic<br />

segregation (e.g., leucosome–palaeosome) processes, or introduced<br />

granitoid melt, were involved in the formation of the compositional<br />

layering. Rather, the high quartz content, the presence of tourmaline,<br />

and textural features of the layering are consistent with an interpretation<br />

as a sheared felsic, aluminous, metasediment.<br />

The origin of the overlying, brecciated unit is more obscure.<br />

However, the monomict nature of the ‘clasts’ and the petrographic<br />

evidence that the ‘clasts’ are similar in composition to the matrix<br />

suggests that this rock is a felsic agglomerate, as with other rocks in<br />

the Theespruit Formation. Such an origin is supported by the fact that<br />

this unit passes along strike to the northwest, as well as across strike,<br />

into the felsic volcanic rocks described above.<br />

4.1.4.3. ‘Basement orthogneiss’ cobbles in ‘diamictite’. The third<br />

setting in which ‘basement orthogneisses’ were identified by de Wit<br />

et al. (1983), was as boulders within ‘sedimentary diamictite’.<br />

Although the boulder depicted in Fig. 6a of de Wit et al. (1983)<br />

could not be relocated, boulders of identical appearance and texture,<br />

with a similar misoriented compositional layering with respect to the<br />

regional foliation, were observed at several places (e.g. Fig. 5b). All<br />

such boulders consist of fine-grained, siliceous, weakly and irregularly<br />

colour-layered rock, identical to the grey-white cherty metasediment<br />

(silicified tuff?) in the lowermost felsic unit. The irregular layering in<br />

the boulders is defined only by a change in colour, not in grain size or<br />

composition, as both grey and white layers are composed of N95%<br />

fine-grained silica. Locally, the darker layers are formed by remobilisation<br />

of grey-black chert into fractures within a white-cream quartz<br />

siltstone/chert. Again, no evidence of leucosome veins, metamorphic<br />

differentiation, or even a granitic composition was observed in the<br />

boulders to confirm an origin as orthogneisses. As such, and because<br />

of the volcanic origin of the host agglomerate, the layered boulders are<br />

re-interpreted here as fragments of underlying sediment entrained in<br />

felsic magma during explosive volcanism.<br />

4.2. Strain<br />

Grain # Mass scans 1<br />

4.2.1. Granite-greenstone contact<br />

In the map area, the Theespruit Pluton is a leucocratic, mediumgrained<br />

trondhjemite that intrudes the greenstones as a series of<br />

little-deformed, bedding-subparallel sheets (sills). Here it represents<br />

the structurally highest and least deformed part of the pluton and has<br />

intruded the highest stratigraphic level in the greenstones. Anhaeusser<br />

(1984) and Kisters and Anhaeusser (1995) described contact<br />

agmatites and ring dykes in other low-strain, high-level parts of the<br />

pluton; the ring dykes are parallel to bedding and thus probably<br />

represent sills turned on edge together with greenstones, during<br />

doming.<br />

Central parts of the pluton to the south of our map area are little<br />

deformed, medium- to coarse-grained leucocratic trondhjemite. Large<br />

Evaporation<br />

temp. in °C<br />

Mean 207 /Pb/ 206 Pb ratio 2<br />

and 2-σm error<br />

207 Pb/ 206 Pb age<br />

and 2-σm error<br />

BA 110 Dark grey, round, 1 88 1597 0.310479 ±52 3523.1±0.3<br />

Detrital 2 103 1594 0.310681±84 3524.3±0.4<br />

Grains 3 65 1595 0.310560±110 3523.7±0.5<br />

Mean of grains 1–3 1–3 0.310580±48 3523.8±0.2<br />

1 207 206 2<br />

Number of Pb/ Pb ratios evaluated for age assessment. Observed mean ratio corrected for non-radiogenic Pb where necessary. Errors based on uncertainties in counting<br />

statistics. ⁎Error of combined mean ages (bold print) is based on reproducibility of internal standard at 0.000027 (2σ).<br />

xenoliths of amphibolite-facies pillow basalt in the core of the pluton<br />

are almost completely undeformed, although they are gently tilted<br />

and face radially away from the core of the pluton, indicating its<br />

domical shape (Fig. 7a). Greenstones along the margins of the pluton<br />

were transformed into strongly foliated and lineated, medium- to<br />

coarsely recrystallised amphibolites and host a transposed network of<br />

granitoid veins emplaced in part during folding. In the narrow septum<br />

of greenstones between the Theespruit and Stolzburg plutons<br />

(Anhaeusser, 1984; Kisters and Anhaeusser, 1995), all rocks have<br />

been intensely deformed into subvertical L-tectonites (Fig. 7b) under<br />

metamorphic conditions up to granulite-facies (9 kbar, 700 °C; <strong>Van</strong><br />

Vuren and Cloete, 1995). The steeply dipping margins of the<br />

Theespruit Pluton are strongly foliated and heterogeneous, with<br />

multiple foliated sheets of variable composition and textural complexity<br />

(Fig. 7c). Such variation was explained by Anhaeusser (1984) as<br />

due to progressive and differential assimilation of marginal greenstones<br />

by successive intrusive trondhjemite sheets and concomitant<br />

deformation of earlier sheets during successive intrusive pulses under<br />

ductile conditions imposed by the heat of the magmas.<br />

4.2.2. Theespruit Formation and KSZ<br />

Rocks in the Tjakastad region vary greatly in strain. In the<br />

southwest, adjacent to the Theespruit Pluton, little deformed rocks<br />

with well-preserved sedimentary and/or volcanic textures are<br />

preserved (e.g. pillows, ocelli, intra-pillow hyaloclastite breccia, and<br />

pillow shelves). Strain increases progressively towards the southern<br />

contact of the KSZ, which is marked by the Theespruit Fault, filled by a<br />

≤12 m wide quartz vein. The southern 300 m of the KSZ is underlain<br />

by high strain ultramylonites derived from mafic, felsic and cherty<br />

precursors (Fig. 8a). Strain then decreases again northeast through the<br />

remainder of the KSZ until the Komati Fault, a locally sharply defined<br />

structure filled by a quartz vein that marks the northern margin of the<br />

KSZ. North of the Komati Fault, only those rocks within ∼20 m of the<br />

fault display a weak foliation, but otherwise, rocks of the Komati<br />

Formation have not been penetratively deformed. Fabric elements<br />

across the area are generally co-axial, characterised by beddingparallel<br />

foliations that are locally axial planar to tight small-scale folds<br />

and lineations which progressively steepen in plunge away from the<br />

Theespruit Pluton (see below). This variation in strain accompanies a<br />

progressive decrease in metamorphic grade from amphibolite-facies<br />

in the southwest, adjacent to the granitoid rocks, to greenschist-facies<br />

in the northeast.<br />

Amphibolite-facies mafic volcanic rocks of the Theespruit<br />

Formation adjacent to the Theespruit Pluton contain a moderate<br />

S–L fabric (3:2:1) in which elongation lineations plunge moderately<br />

to the northwest (Fig. 8b). A narrow high strain zone is developed<br />

along the southern contact of the first felsic horizon, as noted by<br />

de Wit et al. (1983). North of this contact, lineations become more<br />

steeply plunging and more northerly trending in rocks of variable<br />

preservation, but generally the succession becomes more strongly<br />

deformed towards the KSZ. The trend and plunge of extension<br />

lineations across this zone vary between different rock types, as<br />

shown in Fig. 8b, only becoming fully transposed within the<br />

southern part of the KSZ where chloritic ultramylonites contain<br />

subvertical, NNW-SSE striking schistosities and downplunge


M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

mineral elongation lineations. In the northeastern part of the study<br />

area, rocks of the KSZ are pure L-tectonites, with strongly developed<br />

vertical rodding lineations (Fig. 7d).Inthinsectionsofthemost<br />

highly strained schists, kinematic asymmetry could not be identified<br />

due to intense flattening strain. However, in slightly coarser-grained<br />

porphryoblastic and originally porphyritic rocks along the northern<br />

margin of the KSZ, kinematic indicators of SW-side-up displacement<br />

were observed (Fig. 7e). Northeast of the ultramylonitic schists, in<br />

the western part of the KSZ, strain decreases and local sedimentary<br />

features can again be recognised within felsic tuffaceous horizons.<br />

Pillow facing directions to the northeast were identified in basalts.<br />

In the eastern part of the KSZ, the strain is more heterogeneous,<br />

with high strain shear splays near its northern boundary (Fig. 4).<br />

The Komati Fault on the northern margin of the KSZ is exposed in<br />

the western part of the map area, to the south of which SW-side-up<br />

kinematic indicators were observed in serpentinized, schistose metakomatiite<br />

(Fig. 8). In the eastern part of the map area, the Komati Fault<br />

is represented by a quartz vein that separates L-tectonite mylonitic<br />

felsic agglomerates from undeformed quartz-feldspar porphyry at the<br />

base of the Komati Formation (Fig. 4).<br />

4.3. Metamorphism of the Theespruit Formation in the Tjakastad area<br />

Metamorphic grade decreases from amphibolite-facies adjacent to<br />

the Theespruit Pluton (garnet-actinolite pillowed mafic volcanics), to<br />

greenschist- facies in the KSZ, where mafic mylonites contain chloritecarbonate<br />

assemblages, felsic schists consist of sericite and quartz, and<br />

quartz-plagioclase porphyries contain chloritised riebeckite. As<br />

described above, P–T estimates decrease up stratigraphy from the<br />

western margin of the Theespruit Pluton (≤9 kbar, 700 °C; <strong>Van</strong> Vuren<br />

and Cloete, 1995) to the top of the Hooggenoeg Formation (Fig. 2:<br />

1.9 kbar, 320–420 °C; Cloete, 1993, 1999).<br />

The stratigraphically lowest felsic horizon in the Theespruit<br />

Formation contains abundant kyanite porphyroblasts that are elongate<br />

in the lineation direction and syn- to late-kinematic with respect<br />

to the deformation. Andalusite porphyroblasts contain round inclusions<br />

of an altered mineral that now consists of fine-grained quartzsericite<br />

and which may have originally been cordierite. Andalusite<br />

porphyroblasts are commonly overgrown by kyanite, along extension<br />

cracks, indicating replacement of the former by the latter under<br />

conditions of increasing pressure (Fig. 7f). The initial assemblage<br />

andalusite–cordierite indicates P–T conditions of b3.9 kbar, 550 °C<br />

(Spear, 1993), whereas the presence of overgrowing kyanite indicates<br />

an increase in pressure under near-isothermal conditions along part of<br />

an anticlockwise P–T path. As discussed below, these data have<br />

important implications for the tectonic evolution of the BGB.<br />

5. Zircon geochronology and Nd isotopic systematics<br />

Single zircons were dated from 13 felsic Theespruit lithologies and<br />

from one intrusive granite in the Tjakastad area, using the evaporation<br />

technique (Kober, 1986, 1987). The sensitive high-resolution ion<br />

microprobe (SHRIMP II; DeLaeter and Kennedy, 1998) and conventional<br />

analysis by thermal ionisation mass spectrometry (TIMS)<br />

Fig. 9. Histograms showing distribution of Pb isotope ratios derived from evaporation of<br />

single zircons from felsic metavolcanic samples of the Theespruit Formation (a–c, e–h)<br />

and intrusive granite (d), Tjakastad and Songimvelo areas, southern Barberton<br />

Greenstone Belt, South Africa (see Fig. 4 for sample locations): a) Spectrum for 4<br />

grains from felsic sample BA56, integrated from 444 ratios; b) Spectrum for 3 grains<br />

from felsic sample BA57, integrated from 320 ratios; c) Spectrum for 4 grains from felsic<br />

sample BA62, integrated from 363 ratios; d) Spectrum for 3 grains from sheared granite<br />

dyke sample BA65, integrated from 193 ratios; e) Spectrum for 3 grains from felsic tuff<br />

sample BA70, integrated from 220 ratios; f) Spectrum for 3 grains from felsic<br />

agglomerate sample BA71, integrated from 252 ratios; g) Spectrum for 3 grains from<br />

felsic tuff sample BA73, integrated from 220 ratios; h) Spectrum for 3 grains from felsic<br />

tuff sample BA84, integrated from 221 ratios.<br />

129


130 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Fig. 10. Histograms showing distribution of Pb isotope ratios derived from evaporation<br />

of single zircons from felsic metavolcanic samples of the Theespruit Formation<br />

Tjakastad and Songimvelo areas, southern Barberton Greenstone Belt, South Africa (see<br />

Fig. 4 for sample locations): a) Spectrum for 6 grains from felsic tuff sample BA85,<br />

integrated from 581 ratios; b) Spectrum for 3 grains from feldspar porphyry sample BA<br />

105, integrated from 302 ratios; c) Spectrum for 4 grains from felsic tuff sample BA 107,<br />

integrated from 382 ratios; d) Spectrum for 4 grains from felsic agglomerate sample BA<br />

108, integrated from 549 ratios; e) Spectrum for 5 grains from felsic tuff sample BA 109,<br />

integrated from 461 ratios; f) Spectrum for 3 grains from felsic metavolcanic schist<br />

sample BA 110, integrated from 256 ratios.<br />

techniques were used for zircons from one sample. Sample preparation<br />

for the evaporation technique followed methods described in<br />

Kröner et al. (2003). In addition, we determined the Sm–Nd wholerock<br />

isotopic composition of selected samples in order to calculate Nd<br />

model ages and ɛ Nd(t) values. The different analytical procedures are<br />

summarized in Appendix A. Sample locations and short field<br />

characteristics are given in Table 1 and Fig. 4.<br />

Samples BA56, 57, 62, 70, 71, 73, 79, 84, 85, 104, 105, 107, 108, 109<br />

and 110 represent variably foliated felsic volcanic schist layers of the<br />

Theespruit Formation in the Tjakastad and Songimvelo areas (Fig. 4).<br />

We also collected mafic rocks from the base of the succession,<br />

including BA 66, 69 (massive foliated amphibolite), 74 (pillowed<br />

basaltic komatiite) and 75 (gabbroic sill). Samples BA 79, 84, 85, 109<br />

and 110 represent Theespruit felsic tuff or trachytic breccia interlayered<br />

with chert, along strike to the east of the main Tjakastad<br />

samples. The zircons in all samples vary in colour from clear to yellowbrown<br />

and in shape from stubby to long-prismatic. Most are<br />

idiomorphic, occasionally with slight rounding at their terminations.<br />

No core-overgrowth relationships were seen under the microscope.<br />

Between 3 and 6 zircons per sample were evaporated, and the mean<br />

207 Pb/ 206 Pb ages obtained on 13 samples vary within a narrow range<br />

between 3522 and 3533 Ma, with very low errors (generally b1 Ma:<br />

see Table 2 and Figs. 9–10).<br />

Felsic pyroclastic sample 160932 was collected from the middle of<br />

the three main felsic volcaniclastic units identified in the Tjakastad area<br />

(Fig. 4). The horizon is a coarse felsic volcaniclastic rock that contains<br />

irregular-shaped fragments of fine-grained and feldspar-phyric dacite,<br />

and rounded cobbles (to 20 cm diameter) of grey-black and white<br />

layered chert in a fine-grained felsic volcanic matrix (Fig. 5b). The<br />

reasoning behind dating this sample was that it is from the most<br />

heterogeneous felsic layer in the region and thus should have the most<br />

varied detrital zircon population, if derived from a sedimentary<br />

diamictite with basement orthogneiss clasts, as previously suggested.<br />

The sample contained abundant zircon grains with a single<br />

population of dark- to medium-brown, elongate prisms that have<br />

common bipyramidal terminations and internal growth zoning. The<br />

recovered zircon grains are commonly cracked and often broken, but<br />

show no cores or rims. Seven zircon grains were analysed on SHRIMP II<br />

(Table 3 and Fig. 11a). Grain 1 is grossly discordant and probably lost Pb<br />

at unspecified periods in the past; it is therefore not considered for age<br />

assessment. Grains 2–7 are variably discordant but are well aligned<br />

along a chord (MSWD=0.0025) with an upper Concordia intercept at<br />

3539±16 Ma (2-sigma) and a lower intercept at 0±450 Ma. The<br />

weighted mean 207 Pb/ 206 Pb age of the six grains is 3539±2 Ma. We<br />

consider this to reflect the time of eruption of the felsic pyroclastic rock.<br />

TIMS analysis of five single zircons from the same sample (Table 4)<br />

yielded three grossly discordant points (Z2, Z4, Z5) which are not<br />

further considered here, one near-concordant point (Z3) and one<br />

moderately discordant point (Z1) (Fig.11b). A chord through grains Z3<br />

and Z1 suggests recent Pb-loss (Fig. 11b), and we therefore consider<br />

both analyses meaningful. The mean 207 Pb/ 206 Pb age for both Z3 and<br />

Z1 is 3534±4 Ma, which is identical, within error, to the mean<br />

SHRIMP age. Combining all data points (SHRIMP and TIMS) yields a<br />

regression with an upper concordia intercept age of 3534±1 Ma.<br />

In view of the slightly younger age of zircons from felsic<br />

volcaniclastic rocks analysed by evaporation as compared to the<br />

SHRIMP and TIMS results, we cannot completely rule out that nonzero<br />

lead loss has occurred in the younger zircons, although the grains<br />

from individual samples all show near-identical 207 Pb/ 206 Pb ratios,<br />

and our oldest evaporation age of 3533 Ma is identical to a mean<br />

SHRIMP U/Pb zircon age of 3531±10 Ma reported by Armstrong et al.<br />

(1990). We therefore adopt a mean age of c. 3530 Ma for the 14<br />

samples of Theespruit felsic volcanic rocks analysed from the<br />

Tjakastad–Songimvelo area for the purpose of calculating Nd initial<br />

isotopic ratios and T DM model ages. It is important to note that no ca.<br />

3450 Ma zircons were identified during any of these analyses.<br />

Evaporation of three idiomorphic zircon grains from granite dyke<br />

sample BA 65 yielded identical Pb isotopic ratios which combine to a<br />

mean 207 Pb/ 206 Pb age of 3408±0.3 Ma (Table 2, Fig. 9d). We consider<br />

this age to reflect the time of dyke emplacement, and the dyke may<br />

reflect either a late, K-feldspar-rich phase of the Theespruit Pluton, or


Table 3<br />

SHRIMP II analytical data for spot analyses of magmatic zircons from felsic volcaniclastic sample 160932.<br />

Sample no. U ppm Th ppm Th/U 206 Pb/ 204 Pb<br />

208 206<br />

Pb/ Pb<br />

207 206<br />

Pb/ Pb<br />

206 238<br />

Pb/ U<br />

207 235<br />

Pb/ U 206/238 age ±1σ 207/235 age ±1σ 207/206 age ±1σ<br />

BA160932-1 1459 1871 1.28 2969 0.4817±10 0.2110±4 0.2358±28 6.860 ±85 1365±15 2094 ±11 2914±3<br />

BA160932-2 210 140 0.67 22,042 0.1758±8 0.3137±7 0.6851±85 29.636 ±381 3364±32 3475 ±13 3539±3<br />

BA160932-3 573 412 0.72 72,961 0.1830±5 0.3135±4 0.6230 ±76 26.930 ±334 3122±30 3381±12 3538±2<br />

BA160932-4 268 159 0.59 24,413 0.1468±7 0.3135±6 0.6725 ±83 29.068±370 3316±32 3456 ±12 3538±3<br />

BA160932-5 87 42 0.48 86,207 0.1273±11 0.3135±11 0.7122 ±93 30.786 ±428 3467±35 3512±14 3538±5<br />

BA160932-6 527 322 0.61 26,658 0.1517+5 0.3137±5 0.5754±70 24.886±310 2930±29 3304 ±12 3539±2<br />

BA160932-7 187 151 0.81 39,543 0.2123 ±10 0.3137±8 0.7109 ±90 30.750 ±404 3462±34 3511±13 3539±4<br />

part of a phase of granite and aplite emplacement found in all granitic<br />

rocks of the southern Barberton Mountain Land. Importantly, the main<br />

phase of deformation in the dyke and the Theespruit Formation must<br />

be post-3408 Ma.<br />

Whole-rock samples analysed for Sm and Nd isotopes (Table 5)<br />

were not, in all cases, the same as those used for zircon dating, since<br />

some rocks did not contain large or sufficient zircons for evaporation.<br />

In addition to felsic volcanic rocks of the Theespruit Formation, we<br />

also measured isotopic values for mafic/ultramafic samples from the<br />

Theespruit Formation, near outcrops of well-preserved pillowed<br />

basalt that are interlayered with the felsic volcanic rocks.<br />

The ɛNd(t) values for Theespruit felsic rocks (Table 5 and other data in<br />

Hamilton et al., 1979 and Kröner et al., 1996) show a moderate scatter<br />

from −1.4 to + 1.3 (Fig. 12a), suggesting heterogeneous crustal sources,<br />

including juvenile mantle-derived material, as well as recycling of older<br />

crust. Although we cannot preclude disturbance of the Sm–Nd system in<br />

these samples, their high Sm and Nd concentrations probably inhibited<br />

drastic modifications resulting from fluid/rock interaction (e.g., Gruau<br />

et al.,1996). A negative ɛNd value for the pillowed metakomatiite sample<br />

BA 74 may be explained by sample alteration and possibly melt<br />

contamination by older crustal material.<br />

The somewhat lower ɛ Nd(t) in the felsic samples than in the<br />

associated komatiites suggest that these rocks were not comagmatic,<br />

and that the felsic rocks were derived from a crustal source that was<br />

ultimately formed from the same mantle as the associated komatiites.<br />

This is consistent with previous Nd isotopic data for Theespruit felsic<br />

volcanics from the Steynsdorp area farther east that also suggested<br />

involvement of older continental crust (Kröner et al., 1996).<br />

A regression analysis of three komatiitic samples (excluding BA 74,<br />

which has negative ɛ Nd-values of −0.4 at 3.53 Ga that we infer is due<br />

to melt contamination by older crust, or alteration of the Sm–Nd<br />

system by hydrothermal activity that affected the pillow basalts; e.g.,<br />

de Wit et al., 1982; Duchač and Hanor, 1987), yields an age of 3.51±<br />

0.07 Ga (2σ; MSWD=0.02) (Fig. 12b). Enlarging the MSWD to a value<br />

of 1 would reduce the error to c. 10 Ma. This age is identical to the<br />

emplacement age of the Steynsdorp Pluton (3510 Ma; Kröner et al.,<br />

1996) and not much different from the mean zircon age of c. 3530 Ma<br />

for the Theespruit Formation obtained herein. This agreement<br />

suggests that the Sm–Nd system in these rocks behaved as a closed<br />

system during their post-eruption history. We therefore interpret<br />

the initial ɛNd-values of c. +0.5±1.4 (2σ) for the mantle-derived<br />

komatiites as representative of their mantle source. The Theespruit<br />

mantle value is not much different from that of other units of the<br />

Onverwacht Group, which yielded average ɛ Nd values of +2.5<br />

(Lécuyer et al., 1994), +1.9 (Chavagnac, 2004), +1.1 (Lahaye et al.,<br />

1995), and +1.1 (Hamilton et al., 1979), with a total range in values<br />

from +3 to 0 (Fig. 12a).<br />

A regression line calculated for the Theespruit mafic and associated<br />

felsic metavolcanic rocks yields an errochron corresponding to an age<br />

of 3.63±0.13 Ga (2σ, n=9, MSWD=16; initial ɛNd-value=+1.1±<br />

2.1, 2σ). Addition of four samples of the Theespruit Formation from<br />

the Steynsdorp area (Kröner et al., 1996) produces an errorchron age<br />

of 3.60±0.13 Ga (2σ; MSWD=21, initial ɛ Nd-value=+0.8±2, 2σ)<br />

(Fig. 12b). The near-magmatic age derived from the composite<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

regression line indicates that both the crustally-derived felsic rocks<br />

and mantle-derived komatiites were derived from overall isotopically<br />

similar sources. The inclusion of the felsite data in addition to the<br />

komatiitic samples in the regression analysis results in a rotation of<br />

the regression line to a higher age and increase in data scatter. The<br />

scatter of the felsite data may reflect open-system behaviour of the<br />

Sm–Nd system in the felsic rocks or their crustal protoliths. Open-<br />

Fig. 11. a) Concordia diagram showing SHRIMP analyses of zircons from felsic tuff<br />

sample 160932 (Table 3). Data boxes for each analysis are defined by standard errors in<br />

207 Pb/ 235 U, 206 Pb/ 238 U and 207 Pb/ 206 Pb; b) Concordia diagram showing TIMS analyses<br />

of zircons from sample 160932 (Table 4): recent Pb-loss line passes through zircons Z1<br />

and Z3, with an upper intercept age of 3534±4 Ma. Dated sample is shown in Fig. 5b<br />

from locality shown in Fig. 4.<br />

131


132 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Table 4<br />

U–Pb TIMS data for sample 160932.<br />

Fraction Concentration Measured Corrected Atomic ratios a<br />

Ages [Ma]<br />

Weight U Pb R<br />

Common Pb T 206 Pb/<br />

208<br />

Pb/<br />

206 238<br />

Pb/ U<br />

207 235<br />

Pb/ U<br />

207 206<br />

Pb/ Pb<br />

206<br />

Pb/<br />

207<br />

Pb/<br />

207<br />

Pb/<br />

(mg) (ppm) (pg)<br />

204<br />

Pb<br />

206<br />

Pb<br />

238<br />

U<br />

235<br />

U<br />

206<br />

Pb<br />

Z1 brn euh frag 0.001 236 216.7 2 8357 0.1994 0.70423 (156) 30.3641 (698) 0.31271 (26) 3437 3499 3534<br />

Z2 sub prsm 0.001 966 411.1 22 1284 0.2425 0.34605 (72) 8.8451 (176) 0.18538 (26) 1916 2322 2702<br />

Z3 sub prsm 0.010 17 16.6 4 1905 0.2640 0.72643 (208) 31.3197 (884) 0.31270 (38) 3520 3529 3534<br />

Z4 prsm 0.001 267 164.4 7 1600 0.2227 0.48660 (134) 16.6645 (460) 0.24838 (32) 2556 2916 3174<br />

Z5 prsm 0.001 284 171.2 29 331 0.2628 0.46872 (104) 14.7999 (364) 0.22901 (28) 2478 2802 3045<br />

Zircon fractions Z1–Z3 are of single small grains that were well abraded (Krogh, 1982). Pb R refers to radiogenic Pb; Common Pb T refers to total common Pb.<br />

Abbreviations: brn = brown; euh = euhedral; frag = fragment; prsm = prism; sub = subhedral.<br />

a<br />

Ratios corrected for fractionation, laboratory Pb (1 pg) and U (0.25 pg) blank. Initial common Pb calculated using Pb isotopic compositions of Stacey and Kramers (1975). Twosigma<br />

uncertainties on isotopic ratios, calculated with a modified unpublished error propagation program written by L. Heaman, are reported after the ratios (in brackets) and refer to<br />

the final digits.<br />

system behaviour of the Sm–Nd system has been reported in many<br />

studies (Moorbath et al., 1997 and references therein; Gruau et al.,<br />

1996), and it may well be that fluid/rock interaction during<br />

metamorphism of the Theespruit Formation (see above) has affected<br />

the Sm–Nd system. However, any possible secondary disturbance of<br />

the Sm–Nd system had no drastic effects on the samples, as shown by<br />

the reasonable agreement between the whole-rock Sm–Nd ages and<br />

U–Pb zircon data.<br />

6. Discussion<br />

Remapping and dating of the Theespruit Formation and associated<br />

granitic rocks in the Tjakastad area indicates significant differences<br />

from previous tectonic interpretations of this area (de Wit et al., 1983).<br />

In particular, our data question the validity of the thrust-accretion<br />

model because:<br />

1. No tectonic slivers or clasts of basement orthogneiss were found in<br />

the Tjakastad area. Those previously interpreted as such by de Wit<br />

et al. (1983) were found to consist either of undeformed quartzfeldspar<br />

porphyry with igneous flow banding, deformed metasediment,<br />

or clasts of cherty metasediment (metatuff?) within felsic<br />

agglomerate. Furthermore, there is no evidence to support the<br />

presence of an unconformity between “basement orthogneiss” and<br />

“sedimentary diamictite” at locality D in Fig. 3.<br />

2. Rocks within the Theespruit Formation and KSZ that were<br />

interpreted as sedimentary diamictite and intrusion breccia by de<br />

Wit et al. (1983) are re-interpreted here to be felsic volcaniclastic<br />

rocks that form part of a series of felsic volcaniclastic horizons<br />

interlayered with mafic schists of indeterminate intrusive or<br />

extrusive origin. The unique compositional and textural features<br />

of each felsic horizon indicate that they are distinct from one<br />

another and thus do not represent slivers of a thrust-duplicated<br />

unit, as previously suggested (de Wit et al., 1983, 1992).<br />

3. The presence of consistent top-to-the-northeast bedding indicators<br />

across the Theespruit Formation, in combination with the results of<br />

(1) and (2) above, strongly favour a coherent, though strained (see<br />

point 4, below), stratigraphic succession. The presence of local tight<br />

folds and of reversals of bedding in the KSZ (de Wit et al., 1983) is<br />

not considered to be a significant feature, as no large-scale folds of<br />

mappable units have been found.<br />

4. Changes in the intensity and orientation of structures across the<br />

contact between mafic volcanic rocks and the first felsic horizon in<br />

the Theespruit Formation may readily be explained as an example of<br />

strain partitioning between the competent pillowed komatiitic unit<br />

and the well-layered, less competent, and more fully transposed felsic<br />

tuffaceous rocks (cf. figure 26.17 and related text of Ramsay and<br />

Huber, 1987; Hanmer and Passchier, 1991 [pp. 22–24]). Northeast of<br />

this contact, the strain discontinuously increases towards the KSZ, in<br />

concert with more steeply-plunging and more northerly-trending<br />

lineations. These data suggest that the southern contact of the<br />

stratigraphically lowest felsic schist horizon represents the first major<br />

increase in strain approaching the KSZ and that the abrupt change in<br />

lineation orientations across this contact reflects varying degrees of<br />

transposition within a single shear system. Significantly, strain is<br />

highly variable within the KSZ, and largely dependent on lithology,<br />

with the highest strain localised in the softest (talc-carbonate) rocks,<br />

and areas of low strain localised in some pillowed basalts and felsic<br />

volcanic rocks (Fig. 4).<br />

5. The distribution of strain, orientation of fabric elements, and<br />

kinematic data of discrete faults and shear zones within the<br />

Tjakastad area indicate a single set of penetrative fabric elements<br />

developed during greenstone belt-down, granitoid-up extensional<br />

deformation. No evidence of thrust kinematics, fold interference, or<br />

of multiple overprinting sets of structures was observed, and the<br />

period of earlier deformation within orthogneisses as interpreted<br />

by de Wit et al. (1983) has not been substantiated. Compilation of<br />

available data show that the extensional deformation occurred at<br />

ca. 3230 Ma. The KSZ is here interpreted as an extensional<br />

detachment zone, similar to that documented along strike to the<br />

west by Kisters et al. (2003) (see Fig. 1). As discussed below, we<br />

suggest that the most likely origin of the KSZ is as a dislocation<br />

plane to tight, upright, disharmonic folding between deep crustal<br />

levels dominated by the Stolzburg plutonic suite, and high level<br />

rocks of the rest of the Barberton Greenstone Belt.<br />

Table 5<br />

Sm–Nd isotopic data of metavolcanic rocks from the Theespruit Formation.<br />

Sample Age<br />

(Ga)<br />

Sm Nd<br />

147 Sm/<br />

144 Nd<br />

143 144<br />

Nd/ Nd<br />

(meas.)<br />

ɛNd(t) TDM<br />

Felsites<br />

Ba 57 ∼3.53 4.06 26.79 0.09157 0.510113 −1.4 3.69<br />

Ba 70 ∼3.53 7.03 44.50 0.09542 0.510254 −0.4 3.62<br />

Ba 79 ∼3.53 6.35 35.85 0.1071 0.510551 0.1 3.60<br />

Ba 85 ∼3.53 5.84 34.75 0.1016 0.510457 0.8 3.54<br />

Ba 104 ∼3.53 7.29 46.86 0.09407 0.510223 −0.4 3.62<br />

Ba 108 A ∼3.53 1.29 7.093 0.1102 0.510654 0.7 3.55<br />

Komatiites<br />

Ba 66 ∼3.53 4.55 20.07 0.1369 0.511278 0.7 3.60<br />

Ba 69 ∼3.53 3.50 13.74 0.1539 0.511671 0.6 ⁎<br />

Ba 74 ∼3.53 2.65 8.740 0.1833 0.512302 −0.4 ⁎<br />

repeat ∼3.53 2.61 8.682 0.1819 0.512274 −0.4 ⁎<br />

Ba 75 ∼3.53 1.40 4.287 0.1975 0.512685 0.6 ⁎<br />

Standards<br />

BCR-1<br />

6.59 28.84 0.1382 0.512633 −0.1<br />

(n=3; 2σ) ±0.02 ±0.08<br />

±6<br />

BHVO-1 6.15 24.84 0.1496 0.512966 +6.4<br />

6.15 24.85 0.1495 0.512978 +6.6<br />

La Jolla<br />

(n=6; 2σ)<br />

0.511853<br />

±7<br />

−15.3<br />

Within-run error for 143 Nd/ 144 Nd b1.3× 10 −5 (2σ mean), external precision ∼1.3× 10 −5 .<br />

⁎ Sm/Nd-ratio too high for calculation of meaningful model age. Input errors (2σ) for age<br />

calculations: 147 Sm/ 144 Nd=0.3%, 143 Nd/ 144 Nd=0.003%.


Fig. 12. a) Nd isotope evolution diagram for selected Archaean rocks and the felsic and<br />

mafic samples from the Theespruit Formation (diamonds); the average εNd value derived<br />

from a whole-rock regression line of all Theespruit samples is +1.1±2.1 (2σ). Other data<br />

for metavolcanic rocks from the Onverwacht Group (large circles represent average ε Nd<br />

values, bars total range in ɛ Nd values): 1, Chavagnac (2004, calculated for 3.45 Ga); 2,<br />

Hamilton et al. (1979, data derived from a 3.54 Ga whole-rock isochron); 3, Lahaye et al.<br />

(1995, recalculated for 3.49 Ga). Solid diamond = average ε Nd value for Isua (Caro et al.,<br />

2006). Other data points represent average values compiled from the literature and cited<br />

by Caro et al. (2006). Solid squares = North Atlantic Craton, Scotland; cross = North<br />

Atlantic Craton, Labrador; solid triangle = Qianan region (Hebai province), China (Xuan et<br />

al., 1986); solid circles = Yilgarn Craton, Australia; triangles = Superior Craton, Canada;<br />

squares = Indian Craton; open circles = Kaapvaal Craton, South Africa. The Nd-evolution<br />

line for differentiation at 4.4 Ga (stippled line) is from Caro et al. (2006), the other line is<br />

from DePaolo (1981).b)Sm–Nd isochron diagram for komatiite and felsite samples of the<br />

Theespruit Formation. Symbols: solid triangles = komatiites from the Tjakastad area; open<br />

triangles = felsic volcanics from the Tjakastad area; solid circles = felsic volcanics from the<br />

Steynsdorp area (data in Kröner et al., 1996).<br />

6. The petrographic evidence of kyanite overprinting andalusite<br />

presented here is interpreted as a preserved remnant of the<br />

prograde P–T conditions and indicative of near isothermal burial<br />

under moderate temperature conditions. When combined with<br />

previous evidence that metamorphism was syn-kinematic with<br />

regional extension and decreased in temperature away from the<br />

Theespruit Pluton (<strong>Van</strong> Vuren and Cloete, 1995) under low<br />

geothermal gradients (e.g., Dziggel et al., 2002; Diener et al.,<br />

2006), this new observation suggests that metamorphism in the<br />

Theespruit area was not the same as that in more recent<br />

accretionary orogenic belts (Cloete, 1993, 1999), where inverted<br />

metamorphic gradients and clockwise P–T paths are commonly<br />

developed (e.g. Brown, 2006). Instead, the metamorphic data,<br />

when combined with the structural data of weakly deformed cores<br />

to domical granitoids and strong, vertically lineated, greenstone<br />

septae between domical granitoids (e.g. Fig. 7a–c), are more<br />

consistent with episodes of partial convective overturn of green-<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

stone-over-granitoid layered crust, as initially modelled by Mareschal<br />

and West (1980) and discussed by many others (e.g. Collins<br />

et al., 1998; Collins and <strong>Van</strong> Kranendonk, 1999; Sandiford et al.,<br />

2004; <strong>Van</strong> Kranendonk et al., 2004; Bodorkos and Sandiford, 2006).<br />

7. The geometry of weakly strained crests and strongly sheared, diplineated<br />

margins to the TTG domes to the south of the Theespruit<br />

Formation (e.g., Fig. 7a–c) is consistent with partial convective<br />

overturn of a thick, overlying greenstone succession and underlying,<br />

more buoyant and partially melted TTG, as originally proposed by<br />

Viljoen and Viljoen (1969), and subsequently supported by Anhaeusser<br />

(1984) and Kisters and Anhaeusser (1995), possibly within an<br />

overall extensional regime (Kisters et al., 2003). This structural<br />

geometry developed at between c. 3237 and 3216 Ma, which is the<br />

age range of metamorphic zircons and titanite, respectively, from the<br />

Stolzburg and Doornhoek Plutons, and contemporaneous with the<br />

age of the post-tectonic Dalmein Pluton (Kamo and Davis, 1994).<br />

6.1. Significance of the available geochronology<br />

Our new interpretation for the Theespruit–Komati Formations<br />

contact zone contrasts markedly with previous tectonic models based<br />

on geochronological data and this requires some discussion.<br />

Previous tectonic models based on zircon dating suggested that<br />

thrusting occurred at ca. 3230 Ma (Kamo and Davis, 1994). However,<br />

this is negated by the fact that both the higher-grade Theespruit and<br />

Sandspruit Formations structurally beneath the KSZ, and the lowergrade<br />

greenstones above the KSZ (Komati Formation and younger<br />

formations), experienced a common felsic magmatic event at ca.<br />

3450 Ma (Kröner et al., 1991; Kamo and Davis, 1994), and thus were in<br />

contact at, or before, this time. If thrusting did occur, then it must have<br />

been at, or before, ca. 3450 Ma. However, several lines of evidence<br />

preclude such an interpretation. First, only one phase of deformation<br />

is recognised in the area, and that is extensional, and/or related to<br />

partial convective overturn of greenstones and granitoids. Second, an<br />

early thrust scenario is improbable because of the conformable<br />

relations between the Komati Formation and overlying rocks of the<br />

Hooggenoeg (3445 Ma), Kromberg (≤3416–3334 Ma), and Mendon<br />

(3298 Ma) Formations, and Fig Tree (b3260 Ma) Group (Fig. 2: Byerly<br />

et al., 1996; Lowe and Byerly, 2007), which shows that the greenstone<br />

sequences were deposited progressively from at least 3490 to<br />

3260 Ma. Third, the 3408 Ma age of the granite dyke that we dated<br />

from the Tjakastad area (sample BA 65) is significant because it cuts<br />

bedding in the volcanic rocks and is deformed, indicating deformation<br />

occurred after its emplacement.<br />

The main deformation of the belt at ca. 3230 Ma was extensional, as<br />

evidenced by rodding lineations, extensional faults, and kinematic data<br />

from within the KSZ (this paper and Kisters et al., 2003). The 3418±8 Ma<br />

titanite age cited by Dziggel et al. (2001) is most likely related to late stage<br />

felsic magmatism in the area, as indicated by the 3408±2 Ma Pb–Pb<br />

zirconageforthefoliatedgranitedykesampleBA65intheTjakastadarea.<br />

The zircon ages of Armstrong et al. (1990) and Kamo and Davis<br />

(1994) indicate that the unit previously interpreted as a basement<br />

gneiss sliver, but which we re-interpret to represent a felsic<br />

volcaniclastic rock within the Theespruit Formation, was erupted at<br />

3538 Ma. This is the same age as our felsic schist sample 160932<br />

(Tables 3, 4; Fig. 11) and the ages of 13 samples of felsic schist of the<br />

Theespruit Formation located along strike in the map area and to the<br />

east around the Steynsdorp Anticline (Kröner et al., 1996), as well as<br />

for the older population of zircons in the “diamictite” (herein reinterpreted<br />

as felsic agglomerate) from the Tjakastad area (3531±<br />

10 Ma; Armstrong et al., 1990). Based on this remarkable similarity in<br />

age for the same lithostratigraphic unit now dated for 18 samples (this<br />

paper and Kröner et al., 1996), it is clear that 3530 Ma represents the<br />

true age of felsic volcanism in the Theespruit Formation.<br />

The remaining question then becomes what is the origin and<br />

significance of the 3453±6 Ma population of zircons within the<br />

133


134 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Theespruit Formation agglomerate dated by Armstrong et al. (1990)?<br />

The description of zircon morphology by Armstrong et al. (1990),<br />

combined with subsequent information in Kamo and Davis (1994),<br />

allows for a reasonable alternative interpretation.<br />

First, it must be recognised that the dated rock was metamorphosed<br />

under amphibolite-facies conditions, at which grade metamorphic<br />

recrystallisation of zircon and growth of new zircon rims on<br />

older cores are known to occur (e.g., Hoskin and Schaltegger, 2003).<br />

The observation by Armstrong et al. (1990) of core–rim structures in<br />

zircons from their dated sample suggests that the younger population<br />

of zircons may have grown during the contact metamorphic event<br />

associated with the emplacement of the ca. 3450 Ma Stolzburg<br />

plutonic suite (e.g., <strong>Van</strong> Vuren and Cloete, 1995) and associated felsic<br />

intrusions that span the Komati schist zone. Supporting evidence for<br />

such an event was presented by Kamo and Davis (1994) who<br />

identified major felsic igneous activity in the Barberton region at<br />

3476–3440 Ma based on the following ages (listed from structural<br />

base to top): 3460+5/−4 Ma zircon in the older phase of the<br />

Stolzburg Pluton, which is located immediately west of, and connected<br />

to, the 3443+4/−3 Ma Theespruit Pluton (Fig. 1); 3458±2 Ma<br />

titanite in a 3476±10 Ma felsic dyke cutting the Komati Formation;<br />

3457±3 Ma zircon from a felsic sill emplaced into the base of the<br />

Komati Formation; and ca. 3445 ± 5 Ma from the Hooggenoeg<br />

Formation (de Wit et al., 1987b), which conformably overlies, and is<br />

genetically related to, the Komati Formation (Lowe and Byerly, 2007).<br />

We therefore re-interpret the geochronological information from<br />

the Theespruit Formation to indicate that a bimodal, mafic and felsic<br />

volcanic succession was erupted at ca. 3530 Ma, and then intruded by<br />

the Steynsdorp Pluton at 3510 Ma. Following eruption of the Komati<br />

Formation at 3490–3480 Ma, the emplacement of sheets of felsic<br />

plutonic rocks into the volcanic succession at between 3476 and<br />

3443 Ma caused contact metamorphism and the growth of new<br />

zircons and titanite in the host rocks, particularly lower in the section<br />

where the volumetrically significant Stolzburg plutonic suite was<br />

emplaced and resulted in amphibolite-facies contact-style<br />

metamorphism.<br />

6.2. Origin of the Komati schist zone and regional folding<br />

The regional geology of the southwestern Barberton Greenstone Belt<br />

(BGB)isdominatedbyasetoffoldsthatplungeinwardstowardsthecore<br />

of the belt, and are defined by lobate, antiformal granitoids and cuspate,<br />

generally synclinal greenstones (Fig. 13). The strike of fold axial traces<br />

rotates through 130° around the southwestern corner of the BGB, from<br />

almost due north-striking in the southeast, to northeast-striking in the<br />

southwest, to southeast-striking in the northwest (Fig. 13). The lobe-cusp<br />

fold geometry of this area indicates contemporaneous deformation of<br />

competent and incompetent lithologies, respectively. The age of<br />

this folding is tightly constrained by the fact that folds involve the<br />

3229±5 Ma Kaap Valley Tonalite, but are cut by the 3216+2/−1Ma<br />

Dalmein Pluton (Kamo and Davis, 1994: Fig. 1).<br />

Within this broad fold context, the high-strain Komati schist zone<br />

is located within the hinge of a regional, upright anticline, along the<br />

boundary between a thick, upper succession of low grade, low strain,<br />

volcanic flow units of the post-Theespruit Formation rocks of the<br />

Onverwacht Group (≤ca. 3.49 Ga) and a thinner, lower succession of<br />

amphibolite-facies, older greenstones (c. 3.53 Ga Theespruit and<br />

Sandspruit Formations) and ca. 3.45 Ga trondhjemitic intrusions.<br />

Indeed, the KSZ is located directly along the amphibolite-greenschist<br />

isograd (Fig. 4), where ultramafic rocks and greenstones are<br />

transformed into soft talc-carbonate and chlorite schists, respectively,<br />

mineral assemblages perfectly suited to accommodate large degrees<br />

of shear strain. Kinematic data indicate extensional, top to the eastnortheast<br />

shearing along this zone. Kisters et al. (2003) documented<br />

an eastward-dipping, top-to-the-east-northeast extensional detachment<br />

plane at the same structural level as the KSZ along strike to the<br />

west (Figs. 1, 13), which can be considered the continuation of the<br />

same zone. A similar high strain zone occurs at the same structural<br />

level along strike to the east, separating the Steynsdorp Pluton and its<br />

surrounding envelope of Theespruit Formation rocks from lowergrade,<br />

younger rocks of the upper Onverwacht Group in another hinge<br />

of a tight regional anticline (Fig. 1). Given that this zone separates<br />

younger rocks in the north from older, higher-grade rocks in the<br />

southwest, it is most likely that the eastern zone is also extensional;<br />

thus, we consider all of these zones to be a continuous structural<br />

dislocation zone across the southwestern Barberton Greenstone Belt,<br />

intimately related to fold development.<br />

Significantly, these dislocation zones developed at the same time<br />

as folding of the low-grade greenstones of the post-Theespruit<br />

Formation units of the Onverwacht Group, and also at the same time<br />

as the dome-and-keel geometry and peak metamorphic conditions<br />

were being developed in the higher grade rocks (see 2.3; Metamorphism:<br />

Dziggel et al., 2001, 2002; Diener et al., 2006).<br />

Studies of fold mechanics in buckled multilayers with units of<br />

different competencies show that high-strain dislocation zones<br />

commonly develop along unit interfaces and that folding may be<br />

accommodated by a variety of processes, including disharmonic<br />

folding if units are of significantly different thicknesses (e.g., Ramsay<br />

and Huber, 1987). Considering the contemporaneity of fold formation<br />

and shear zone development, the most likely origin of the KSZ and its<br />

equivalents along strike to the east and west, is as an extensional<br />

detachment surface during regional folding. That folds in the upper<br />

Onverwacht Group are tighter than folds of the KSZ and its equivalents<br />

along strike, but that folding also affected the shear zones themselves,<br />

suggests that folding was contemporaneous with shearing. As<br />

discussed below, the fact that the rocks above the detachment surface<br />

display one style (upright, large amplitude and large wavelength<br />

folds) and those below display another style (dome-and-keel<br />

geometry: see below) suggests disharmonic folding and/or a different<br />

response at different structural levels to similar overall driving forces.<br />

In the upper-crustal, lower-grade rocks, the inward-plunging<br />

nature of the folds around the southwestern BGB into the core of<br />

the greenstone belt, together with the greenstone-down kinematics<br />

across the shear zones, strongly suggest that deformation was driven<br />

by greenstone sinking, as previously suggested for this area (Viljoen<br />

and Viljoen, 1969; Anhaeusser, 1984). Greenstone sinking was<br />

accompanied by the exhumation of deeper level rocks beneath the<br />

detachment zones, as indicated by the presence of higher pressure<br />

metamorphic mineral assemblages in these rocks. Exhumation along<br />

the Komati schist zone may have excised a considerable thickness of<br />

stratigraphy between the Theespruit and Komati Formations,<br />

although this is unconstrained.<br />

In the higher-grade rocks, the presence of vertically-rodded Ltectonites<br />

in strongly deformed greenstone keels between low-strain,<br />

domed granitoid cores is consistent with deformation driven by<br />

greenstone sinking. This is in contrast to, and different from,<br />

deformation generated as a result of granite doming (diapirism),<br />

where greenstones in the roof zones of diapric granitoids should be<br />

extensively flattened (e.g. Ramberg, 1967): however, this is clearly not<br />

the case, as shown in Fig. 7a. Greenstone sinking is confirmed by the<br />

occurrence of high-P metamorphic mineral assemblages of 9 kbar,<br />

700 °C in highly transposed greenstone septae between domes<br />

(Fig. 13: <strong>Van</strong> Vuren and Cloete, 1995). The low geothermal gradients<br />

indicated by metamorphic studies in this area are also consistent with<br />

greenstone sinking (cf. Collins and <strong>Van</strong> Kranendonk, 1999).<br />

Some aspects of the geology of this area have similarities with<br />

metamorphic core complexes, as suggested by Kisters et al. (2003).<br />

However, the fact that folds and metamorphic mineral elongation and<br />

stretching lineations point inwards towards the core of BGB contrasts<br />

with the typically unidirectional trend of linear fabric elements within<br />

metamorphic core complexes, as does the dome-and-keel geometry of<br />

folding within the higher-grade footwall in the Tjakastad area. Other


differences between the geology of the Barberton Mountain Land and<br />

Phanerozoic metamorphic core complexes include the lack of<br />

evidence in the former for thinning of the lithosphere during<br />

extension, and for a prior episode of structural thickening of the<br />

lithosphere through orogenesis (see also Hickman and <strong>Van</strong> Kranendonk,<br />

2004; <strong>Van</strong> Kranendonk et al., 2004).<br />

6.3. Regional orogeny<br />

Of critical importance relating to interpretations of the tectonic<br />

evolution of the BGB, is the contrasting evidence for contemporaneous<br />

extensional deformation (KSZ and associated zones along strike) and<br />

compressional deformation at c. 3230 Ma. Extension is indicated by<br />

the KSZ and associated zones along strike, and by the dome-and-keel<br />

architecture of c. 3.45 Ga trondhjemitic rocks along the southwestern<br />

margin of the belt. Compressional deformation is indicated by the<br />

presence of upright folds around the southwestern BGB, synsedimentary<br />

contractional folds (including recumbent isoclinal<br />

folds) in the Moodies Group (e.g. Lamb, 1987), and horizontaltectonics<br />

style structures in the Fig Tree Group (e.g., de Wit, 1982).<br />

Significantly, extensional structures are restricted to the higher-grade,<br />

lower structural and stratigraphic parts of the BGB below the KSZ<br />

detachment, whereas compressional structures are restricted to<br />

higher structural and stratigraphic levels within the lower-grade<br />

core of the BGB, above the KSZ detachment. Whereas this contempor-<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Fig. 13. Simplified litho-structural map of the southwestern Barberton Greenstone Belt and surrounding area, showing the position of high strain shear zones within the regional<br />

lobe-cusp fold pattern. Lower greenstones = Sandspruit and Theespruit Formations; Upper greenstones = main part of Onverwacht Group (Komati, Hooggenoeg, Kromberg and<br />

Mendon Formations) and Fig Tree and Moodies Groups. DP = Doornhoek Pluton; KSZ = Komati schist zone; SP = Stolzburg Pluton; TP = Theespruit Pluton. Metamorphic pressure–<br />

temperature conditions also shown (see text for references). Towns: B = Barberton, L = Lochiel, T = Tjakastad.<br />

aneity of contrasting styles does occur in Phanerozoic plate-collisional<br />

orogens, extension in these terranes follows significant crustal<br />

thickening (crustal doubling) through thrusting, caused by the<br />

collision of tectonic plates (e.g. Searle, 2007).<br />

We contend that our new data, combined with the evidence of an<br />

upward-younging, thick greenstone stratigraphy, precludes an origin<br />

of the BGB through subduction–accretion, as widely supposed (e.g. de<br />

Wit et al., 1992; Lowe, 1994; Stevens et al., 2002; Moyen et al., 2006).<br />

Rather, we suggest that the available data supports an origin of the<br />

BGB through partial convective overturn of a dense upper crust and<br />

more buoyant middle crust at c. 3230 Ma, similar to models previously<br />

proposed by Viljoen and Viljoen (1969) and Anhaeusser (1984).<br />

We envisage a two-stage development of ancient Kaapvaal Craton<br />

crust.<br />

1) Long-lived, predominantly mantle-derived magmatism from<br />

3530–3416 Ma, which resulted in the construction of a thick greenstone<br />

pile, based on a substrate of older continental crust, perhaps including<br />

the ≤c. 3.64 Ga Ancient Gneiss Complex (AGC) in Swaziland, on the<br />

southeastern side of the BGB (Kröner, 2007). This was further thickened<br />

by the emplacement of c. 3470–3437 Ma TTG melts, for which<br />

geochemical data indicates melting under a range of pressures from a<br />

dominantly infracrustal basaltic source (Moyen et al., 2007): this implies<br />

that the Kaapvaal crust was already thick by this time. Emplacement of<br />

TTG into the mid-crust caused low-pressure metamorphism of adjacent<br />

greenstones (e.g. andalusite in the present study area).<br />

135


136 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

2) Subsequent addition of mantle-derived melts at 3334–3298 Ma<br />

(Kromberg and Mendon Formations) resulted in magmatic overthickening<br />

of the crust and development of a gravitational instability<br />

represented by a N10 km thick, largely komatiitic volcanic pile on top of<br />

granitic (TTG and AGC) middle crust. Modelling studies show that this<br />

type of gravitationally unstable Palaeoarchaean crustal architecture,<br />

when combined with the input of conductive heat from the mantle<br />

melts and radiogenic heat in TTGs buried to mid-crustal levels by the<br />

associated eruptives, results in partial melting of the TTG middle crust,<br />

which, in turn leads to sinking of greenstones and doming of the<br />

partially melted TTG during partial convective overturn (Rey et al., 2003;<br />

Sandiford et al., 2004; Bodorkos and Sandiford, 2006). We suggest that<br />

this mechanism, rather than subduction–accretion, was responsible for<br />

the deep burial of greenstones along the flanks of doming granitoids in<br />

the lower structural parts of the terrain, where high-pressure<br />

metamorphic assemblages have been recorded (e.g. Dziggel et al.,<br />

2002; Diener et al., 2006). Partial convective overturn was rapid, with<br />

associated isothermal decompression of the lower structural greenstones<br />

during extension-related exhumation at ca. 3230 Ma (Kisters<br />

et al., 2003; Dziggel et al., 2005). Uplift of the granitoid margins to the<br />

BGB shed detritus into the sinking greenstone core (Fig Tree and<br />

Moodies Groups), which became deformed as the sediments accumulated.<br />

Horizontal translation of the greenstones into the zone of sinking<br />

from adjacent to the rising granitoids was effected across mid-crustal<br />

detachment zones such as the KSZ and caused horizontal-style<br />

deformation of the upper greenstones (including recumbent isoclinal<br />

folds: e.g., Ramberg, 1967).<br />

The model of partial convective overturn for the development of<br />

the BGB outlined above should be considered as a working, testable<br />

hypothesis and that it is not necessarily exclusive of subduction/<br />

accretion tectonics in the development of the craton as a whole, nor of<br />

Archaean plate tectonics in general (e.g., <strong>Van</strong> Kranendonk, 2004,<br />

2007; Smithies et al., 2005).<br />

7. Conclusions<br />

Remapping of the southwestern part of the Barberton Greenstone<br />

Belt, combined with zircon dating and Sm–Nd data, has revealed several<br />

important inconsistencies with the previous horizontal tectonic model<br />

of de Wit et al. (1983, 1992).Mostsignificantly, the kinematics of shear<br />

deformation in the Theespruit Formation is extensional rather than<br />

contractional, as previously proposed. Rocks previously interpreted as<br />

tectonic slivers of basement orthogneiss were found instead to be<br />

composed of little deformed quartz-feldspar porphyry or sheared felsic<br />

volcaniclastic metasediment, and those interpreted as sedimentary<br />

diamictite were found to be felsic agglomerates. Consistent facing<br />

directions of bedding to the northeast, the recognition of distinct<br />

stratigraphic variations between several felsic volcanic and volcaniclastic<br />

horizons across strike, and strain data all combine to suggest that the<br />

Theespruit Formation in this area is a ca. 3530 Ma old, moderately<br />

strained, coherent lithostratigraphic succession and not a tectonostratigraphic<br />

assemblage as proposed by de Wit et al. (1983, 1992).<br />

We re-interpret a 3458 Ma population of zircons dated from an<br />

amphibolite-facies Theespruit Formation felsic schist by Armstrong<br />

et al. (1990) as being of metamorphic, rather than detrital origin,<br />

based on these authors' observation of core–rim structures in zircons,<br />

the amphibolite-facies metamorphic grade of the dated rocks, and our<br />

observations for the widespread emplacement of felsic plutonic rocks<br />

of that age across the Komati Schist Zone, which we suggest provided<br />

heat for contact metamorphism of the greenstones.<br />

In combination with previous work (Byerly et al., 1996; Lowe and<br />

Byerly, 2007), the data presented here show that the Barberton<br />

Supergroup is an upward-younging, autochthonous succession<br />

deposited over ∼300 Ma (3530–3230 Ma). Deformation commenced<br />

during deposition of the Moodies Group at ca. 3230 Ma, and involved<br />

partial convective overturn of upper and middle crust.<br />

Acknowledgements<br />

MVK thanks Thinus Cloete for help with fieldwork logistics and for an<br />

introductory fieldtrip through the Barberton Mountain Land in 1998. A.K.<br />

acknowledges mass spectrometric analytical facilities in the Max-Planck-<br />

Institut für Chemie in Mainz. Zircon analyses of sample 160932 were<br />

carried out on the Sensitive High Resolution Ion Microprobe Mass<br />

Spectrometer (SHRIMP II) operated by a consortium consisting of Curtin<br />

University of Technology, the Geological Survey of Western Australia and<br />

the University of Western Australia with the support of the Australian<br />

Research Council. We thank M.T.D. Wingate for assistance during SHRIMP<br />

analysis. A.K. also appreciates financial support from the Institute of<br />

Geoscience Research (TIGeR), Curtin University of Technology, Perth. Paul<br />

Evins and two anonymous reviewers provided comments that helped<br />

clarify parts of the manuscript. This paper is published with the<br />

permission of the Executive Director of the Geological Survey of Western<br />

Australia and is contribution no. 436 of the Geocycles Cluster of Mainz<br />

University, funded by the State of Rhineland-Palatinate, Germany.<br />

Appendix A. Laboratory procedures<br />

A.1. Single zircon evaporation<br />

Our laboratory procedures, as well as comparisons with conventional<br />

and ion-microprobe zircon dating, are detailed in Kröner et al.<br />

(1991) and Kröner and Hegner (1998). Isotopic measurements were<br />

carried out on a Finnigan-MAT 261 mass spectrometer at the Max-<br />

Planck-Institut für Chemie in Mainz.<br />

The calculated ages and uncertainties are based on the means of all<br />

ratios evaluated and their 2σ mean errors. Mean ages and errors for several<br />

zircons from the same sample are presented as weighted means of the<br />

entire population. During the course of this study we repeatedly analysed<br />

fragments of large, homogeneous zircon grains from the Palaborwa<br />

Carbonatite, South Africa. Conventional U–Pb TIMS analyses of six separate<br />

grain fragments from this sample yielded a 207 Pb/ 206 Pb age of 2052.2±<br />

0.8 Ma (2σ, W.Todt,unpublisheddata),whereasthemean 207 Pb/ 206 Pb<br />

ratio for 19 grains, evaporated individually over a period of 12 months, was<br />

0.126634±0.000027 (2σ error of the population), corresponding to an<br />

age of 2051.8±0.4 Ma, which is identical to the U–Pb TIMS age, within<br />

error. The above 2σ error of the population of evaporated zircons is<br />

considered the best estimate for the reproducibility of our evaporation<br />

data and corresponds approximately to the 2σ (mean) error reported for<br />

individual analyses in this study (Table 2). In the case of combined data<br />

sets, the 2σ (mean) error may become very low, and whenever this error<br />

was less than the reproducibility of the internal standard, we have used<br />

the latter value (that is, an assumed 2σ error of 0.000027).<br />

The analytical data are presented in Table 2, andthe 207 Pb/ 206 Pb<br />

spectra are shown in histograms that permit visual assessment of the<br />

data distribution from which the ages are derived. The evaporation<br />

technique provides only Pb isotopic ratios, and there is no aprioriway to<br />

determine whether a measured 207 Pb/ 206 Pb ratio reflects a concordant<br />

age. Thus, all 207 Pb/ 206 Pb ages determined by this method are<br />

necessarily minimum ages. However, many studies have demonstrated<br />

that there is a very strong likelihood that these data represent true zircon<br />

crystallisation ages when (1) the 207 Pb/ 206 Pb ratio does not change with<br />

increasing temperature of evaporation and/or (2) repeated analyses of<br />

grains from the same sample at high evaporation temperatures yield the<br />

same isotopic ratios within error. Comparative studies by evaporation,<br />

conventional U–Pb dating, and ion-microprobe analysis have shown this<br />

to be correct (Kröner et al.,1991,1999; Cocherie et al.,1992; Jaeckel et al.,<br />

1997; Karabinos, 1997).<br />

A.2. SHRIMP II analysis<br />

U–Pb ion microprobe data for zircons from sample 160932 were<br />

obtained on the SHRIMP II (B) of the John de Laeter Centre of Mass


Spectrometry at Curtin University, Australia (DeLaeter and Kennedy,<br />

1998). Clear euhedral zircons some 80 to 150 µm in length were<br />

handpicked and mounted in epoxy resin together with chips of the<br />

Perth zircon standard CZ3. For data collection, six scans through the<br />

critical mass range were made, and details on the analytical procedure<br />

and data reduction can be found in Compston et al. (1992), Stern<br />

(1997), Nelson (1997), and Williams (1998). Primary beam intensity<br />

was about 2.0 nA, and a Köhler aperture of 100 µm diameter was used,<br />

giving a slightly elliptical spot size of about 30 µm. Sensitivity was<br />

about 26 cps/ppm/nA Pb on the standard CZ3. Analyses of samples<br />

and standards were alternated to allow assessment of Pb + /U +<br />

discrimination. Common-Pb corrections have been applied using the<br />

204 Pb-correction method, and because of low counts on 204 Pb it was<br />

assumed that common lead is surface-related (Kinny, 1986), and the<br />

isotopic composition of Broken Hill lead was used for correction. The<br />

analytical data are presented in Table 3. Errors on individual analyses<br />

are given at the 1-sigma level and are based on counting statistics and<br />

include the uncertainty in the standard U/Pb age (Nelson, 1997).<br />

Errors for pooled analyses are at the 2-sigma or 95% confidence<br />

interval. The data are graphically presented on a conventional<br />

concordia plot (Fig. 11a).<br />

A.3. TIMS analysis of zircons<br />

Rock samples were crushed to mineral size under clean conditions<br />

using a jaw crusher and disc pulveriser and initial heavy mineral<br />

concentrates were made using a Wilfley table at The University of Texas<br />

at Austin. Wilfley table concentrates were sieved to remove particles<br />

larger than 70 mesh and zircon was concentrated using standard heavy<br />

liquid and magnetic separation techniques. Zircons were characterised<br />

using a binocular reflected-light microscope, transmitted light petrographic<br />

microscope (with condenser lens inserted to minimise edge<br />

refraction) and a scanning cathodoluminescence (CL) imaging system<br />

on a JEOL 730 scanning electron microscope.<br />

Single grains of zircon were strongly abraded following Krogh<br />

(1982), subsequently re-evaluated optically and then washed successively<br />

in distilled 4 N nitric acid, water and acetone. They were loaded<br />

dry into TEFLON capsules with a mixed 205 Pb/ 235 U isotopic tracer<br />

solution and dissolved with HF and HNO 3. Chemical separation of U and<br />

Pb from zircon using minicolumns (0.055 ml resin volume; after Krogh<br />

1973) resulted in total Pb and U procedural blanks that are estimated to<br />

be 1 and 0.25 pg, respectively. Pb and U were loaded together with silica<br />

gel and phosphoric acid onto an outgassed filament of zone-refined<br />

rhenium ribbon and analysed on a multi-collector MAT 261 thermal<br />

ionisation mass spectrometer operating in dynamic mode with all<br />

masses measured sequentially by the secondary electron multiplier<br />

(SEM) — ion counting system. Ages were calculated using decay<br />

constants of Jaffey et al. (1971). Errors on isotopic ratios were calculated<br />

by propagating uncertainties in measurement of isotopic ratios,<br />

fractionation and amount of blank with a program written by JNC.<br />

Results are reported in Table 4 with 2σ errors.<br />

A.4. Sm–Nd isotopic analysis<br />

Whole-rock isotopic analyses were carried out at Tübingen and<br />

Munich Universities using a technique described in Hegner et al.<br />

(1995) and Hegner and Kröner (2000). The sample powders were<br />

totally spiked with a 150 Nd– 149 Sm tracer and dissolved initially in an<br />

open-beaker overnight, followed by dissolution of accessory phases in<br />

steel jacket-lined TEFLON bombs at 160 °C over 5 days, ensuring<br />

decomposition of e.g. zircon and garnet. Total procedural blanks were<br />

b100 pg for Nd and Sm and are not significant for the samples of the<br />

present study.<br />

143 Nd/ 144 Nd isotopic ratios were measured using MAT 262 and MAT<br />

261 mass spectrometers and a dynamic-quadruple mass collection<br />

mode. Nd and Sm concentrations were calculated from static collection<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

routines. 143 Nd/ 144 Nd ratios are normalized to 146 Nd/ 144 Nd=0.7219,<br />

and Sm isotopic ratios to 147 Sm/ 152 Sm=0.56081. During the course of<br />

this study the La Jolla Nd standard yielded 143 Nd/ 144 Nd=0.511853±7<br />

(2σ, n=6) and Ames-Nd, prepared at Munich 0.512139±11. Analyses<br />

of the rock standards BCR-1 and BHVO-1 are included in Table 5.Withinrun<br />

precision (2σ mean) for Nd was ≤1.3× 10 −5 ; the external precision<br />

inferred from a large number of analyses is about 1.3×10 −5 (2σ),<br />

resulting in an error of ca.±0.3 ɛNd units.<br />

The Sm–Nd isotopic data are listed in Table 5.TheɛNd(t) values have<br />

been calculated for an age of 3.53 Ga using present-day Bulk Earth values<br />

of 143 Nd/ 144 Nd=0.512638 (normalized to 146 Nd/ 144 Nd=0.7219) and<br />

147 Sm/ 144 Nd=0.1967 (Jacobsen and Wasserburg, 1980; Wasserburg<br />

et al.,1981). The Nd model ages were calculated using the parameters of<br />

DePaolo (1981). We interpret the Nd model ages in terms of mean<br />

crustal residence ages (Arndt and Goldstein, 1987).<br />

References<br />

Anhaeusser, C.R., 1984. Structural elements of Archaean granite-greenstone terranes as<br />

exemplified by the Barberton Mountain Land, southern Africa. In: Kröner, A.,<br />

Greiling, R. (Eds.), Precambrian Tectonics Illustrated. E. Schweizerbart'sche<br />

Verlagsbuchhandlung, Stuttgart, pp. 57–78.<br />

Anhaeusser, C.R., Mason, R., Viljoen, M.J., Voljoen, R.P., 1969. A reappraisal of some<br />

aspects of Precambrian shield geology. Geological Society of America Bulletin 80,<br />

2175–2200.<br />

Armstrong, R.A., Compston, W., de Wit, M.J., Williams, I.S., 1990. The stratigraphy of the<br />

3.5–3.2 Ga Barberton Greenstone Belt revisited: a single zircon ion microprobe<br />

study. Earth and Planetary Science Letters 101, 90–106.<br />

Arndt, N.T., Goldstein, S.L., 1987. Use and abuse of crust-formation ages. Geology 15,<br />

893–895.<br />

Bodorkos, S., Sandiford, M., 2006. Thermal and mechanical controls on the evolution of<br />

Archean crustal deformation: examples from Western Australia. In: Benn, K.,<br />

Mareschal, J.-M., Condie, K. (Eds.), Archean Geodynamics and Environments.<br />

American Geophysical Union. Geophysical Monograph Series, vol. 64, pp. 131–147.<br />

Brandl, G., Cloete, M., Anhaeusser, C.R., 2006. Archean greenstone belts. In: Johnson, M.R.,<br />

Anhaeusser, C.R., Thomas, R.J., (eds.). The Geology of South Africa. Geological Society of<br />

South Africa, Johannesburg/Council for Geoscience, Pretoria, 9–56.<br />

Brown, M., 2006. Duality of themal regimes is the distinctive characteristic of plate<br />

tectonics since the Neoarchean. Geology 34, 961–964.<br />

Byerly, G.R., Kröner, A., Lowe, D.R., Todt, W., Walsh, M.M., 1996. Prolonged magmatism<br />

and time constraints for sediment deposition in the early Archaean Barberton<br />

greenstone belt: evidence from the Upper Onverwacht and Fig Tree Groups.<br />

Precambrian Research 78, 125–138.<br />

Byerly, G.R., Lowe, D.R., Wooden, J.L., Xie, X., 2002. An Archean impact layer from the<br />

Pilbara and Kaapvaal cratons. Science 297, 1325–1327.<br />

Caro, G., Bourdon, B., Birck, J.L., Moorbath, S., 2006. High-precision 142 Nd/ 144 Nd<br />

measurements in terrestrial rocks: constraints on the early differentiation of the<br />

Earth's mantle. Geochimica et Cosmochimica Acta 70, 164–191.<br />

Chavagnac, V., 2004. A geochemical and Nd isotopic study of Barberton komatiites<br />

(South Africa): implication for the Archean mantle. Lithos 75, 253–281.<br />

Cloete, M., 1993. Burial metamorphism in the 3.47 Ga Barberton Greenstone Belt, South<br />

Africa. In: Vergara, M., Klem, U. (Eds.), Low Temperature Metamorphism: Processes,<br />

Products and Economic Significance, Abstracts. IGCP Project 294. Department of<br />

Geology, University of Santiago, pp. 1–7. Chile.<br />

Cloete, M., 1999. Aspects of volcanism and metamorphism of the Onverwacht Group<br />

lavas in the southwestern portion of the Barberton Greenstone Belt. Mem.<br />

Geological Survey of South Africa 84, 232p.<br />

Cloete, M., King, V., 1991. Structure and metamorphism of the Komati Formation at<br />

Spinifex Stream and Hooggenoeg Formation in the Songimvelo game reserve. In:<br />

Ashwal, L.D. (Ed.), Two Cratons and an Orogen — Excursion Guidebook and Review<br />

Articles for a Field Workshop through Selected Archean Terranes of Swaziland,<br />

South Africa and Zimbabwe, I.G.C.P. Project 280. Department of Geology, University<br />

of the Witwatersrand, Johannesburg, South Africa, pp. 123–139.<br />

Cocherie, A., Guerrot, C., Rossi, P., 1992. Single-zircon dating by stepwise Pb evaporation:<br />

comparison with other geochronological techniques applied to the Hercynian<br />

granites of Corsica, France. Chemical Geology 101, 131–141.<br />

Collins, W.J., <strong>Van</strong> Kranendonk, M.J., 1999. Model for the development of kyanite during<br />

partial convective overturn of Archaean granite-greenstone terranes: the Pilbara<br />

Craton, Australia. Journal of Metamorphic Geology 17, 145–156.<br />

Collins, W.J., <strong>Van</strong> Kranendonk, M.J., Teyssier, C., 1998. Partial convective overturn of<br />

Archaean crust in the east Pilbara Craton, Western Australia: driving mechanisms<br />

and tectonic implications. Journal of Structural Geology 20, 1405–1424.<br />

Compston, W., Williams, I.S., Kirschvink, J.L., Zhang, Z., Ma, G., 1992. Zircon U–Pb ages for<br />

the Early Cambrian time scale. Journal of the Geological Society of London 149,<br />

171–184.<br />

Dann, J.C., 2000. The Komati Formation, Barberton greenstone belt, South Africa, Part I:<br />

new map and magmatic architecture. South African Journal of Earth Sciences 103,<br />

47–68.<br />

Dann, J.C., Grove, T.L., 2007. Volcanology of the Barberton greenstone belt, South Africa:<br />

inflation and evolution of flow fields. In: <strong>Van</strong> Kranendonk, M.J., Smithies, R.H., Bennet,<br />

137


138 M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

V. (Eds.), Earth's Oldest Rocks. Developments in Precambrian Geology, vol.15. Elsevier,<br />

Amsterdam, pp. 527–570.<br />

DeLaeter, J.R., Kennedy, A.K., 1998. A double focusing mass spectrometer for<br />

geochronology. International Journal of Mass Spectrometry 178, 43–50.<br />

DePaolo, D.J., 1981. Neodymium isotopes in the Colorado Front Range and crust–mantle<br />

evolution in the Proterozoic. Nature 291, 193–196.<br />

de Vries, S.T., Nijman, W., Armstrong, R.A., 2006. Growth-fault structure and<br />

stratigraphic architecture of the Buck Ridge volcano-sedimentary complex, upper<br />

Hooggenoeg Formation, Barberton Greenstone Belt, South Africa. Precambrian<br />

Research 149, 77–98.<br />

de Wit, M.J., 1982. Gliding and overthrust nappe tectonics in the Barberton Greenstone<br />

Belt. Journal of Structural Geology 4, 117–136.<br />

de Wit, M.J., 1983. Notes on a preliminary 1:25 000 geological map of the southern part<br />

of then Barberton greenstone belt. Geological Society of South Africa, Special<br />

Publication 9, 185–187.<br />

de Wit, M.J., 1991. Archaean greenstone belt tectonism and basin development: some<br />

insights from the Barberton and Pietersburg greenstone belts, Kaapvaal Craton,<br />

South Africa. Journal of African Earth Sciences 13, 45–63.<br />

de Wit, M.J., 1998. On Archean granites, greenstones, cratons and tectonics: does the<br />

evidence demand a verdict? Precambrian Research 91, 181–226.<br />

de Wit, M.J., Hart, R., Martin, A., Abbott, P., 1982. Archean abiogenic and probable<br />

structures associated with mineralized hydrothermal vent systems and regional<br />

metasomatism, with implications for greenstone belt studies. Economic Geology<br />

77, 1783–1802.<br />

de Wit, M.J., Fripp, R.E.P., Stanistreet, I.G., 1983. Tectonic and stratigraphic implications<br />

of new field observations along the southern part of the Barberton greenstone belt.<br />

Geological Society of South Africa, Special Publication 9, 21–30.<br />

de Wit, M.J., Hart, R.A., Hart, R.J., 1987a. The Jamestown Ophiolite Complex, Barberton<br />

Mountain Belt: a section through 3.5 Ga oceanic crust. Journal of African Earth<br />

Sciences 5, 681–730.<br />

de Wit, M.J., Hart, R.J., Wilson, A.H., 1987b. Felsic igneous rocks within the 3.3.–3.5 Ga<br />

Barberton greenstone belt: high crustal level equivalents to the surrounding<br />

tonalite–trondhjemite terrane, emplaced during thrusting. Tectonics 6, 529–549.<br />

de Wit, M.J., Roering, C., Hart, R.J., Armstrong, R.A., de Ronde, C.E.J., Green, R.W.E., Tredoux, M.,<br />

Peberdy, E., Hart, R.A., 1992. Formation of an Archaean continent. Nature 357, 553–562.<br />

Diener, J.G., Stevens, G., Kisters, A.F.M., 2006. High-pressure, low-temperature<br />

metamorphism in the southern Barberton granitoid greenstone terrain, South<br />

Africa: a record of overthickening and collapse of Mid-Archean continental crust.<br />

In: Benn, K., Mareschal, J.-C., Condie, K.C. (Eds.), Archean Geodynamic Processes.<br />

Monograph, vol. 164. American Geophysical Union, pp. 239–254.<br />

Dixon, J.M., Summers, J.M., 1983. Patterns of total and incremental strain in subsiding<br />

troughs: experimental centrifuged models of inter-diapir synclines. Canadian<br />

Journal of Earth Sciences 20, 1843–1861.<br />

Duchač, K.C., Hanor, J.S., 1987. Origin and timing of the metasomatic silicification of an<br />

early Archean komatiite sequence, Barberton Mountain land, South Africa.<br />

Precambrian Research 37, 125–146.<br />

Dziggel, A., Stevens, G., Armstrong, R.A., Droop, G.T.R., Poujol, M., Anhaeusser, C.R., 2001.<br />

Metamorphism during terrane accretion in the Barberton greenstone belt. In: Cassidy,<br />

K.F., Dunphy, J.M., <strong>Van</strong> Kranendonk, M.J. (Eds.), 4th International Archean Symposium,<br />

Extended Abstracts. AGSO-Geoscience Australia, pp. 39–41. Record 2001/37.<br />

Dziggel, A., Stevens, G., Poujol, M., Anhaeusser, C.R., Armstrong, R.A., 2002.<br />

Metamorphism of the granite-greenstone terrane south of the Barberton greenstone<br />

belt, South Africa: an insight into the tectono-thermal evolution of the ‘lower’<br />

portions of the Onverwacht Group. Precambrian Research 114, 221–247.<br />

Dziggel, A., Armstrong, R.A., Stevens, G., Nasdala, L., 2005. Growth of zircon and titanate<br />

during metamorphism of the Barberton greenstone belt, South Africa. Mineralogical<br />

Magazine 69, 1019–1036.<br />

Gruau, G., Rosing, M., Bridgwater, Gill, R.C.O., 1996. Resetting of Sm–Nd systematics<br />

during metamorphism of N3.7-Ga rocks: implications for isotopic models of early<br />

Earth differentiation. Chemical Geology 133, 225–240.<br />

Hamilton, W.B., 1998. Archean magmatism and deformation were not the products of<br />

plate tectonics. Precambrian Research 91, 143–180.<br />

Hamilton, P.J., Evenson, N.M., O'Nions, R.K., 1979. Sm–Nd dating of Onverwacht Group<br />

Volcanics, southern Africa. Nature 279, 298–300.<br />

Hanmer, S., Passchier, C., 1991. Shear Sense Indicators: A Review. Geological Survey of<br />

Canada, Paper 90–17.<br />

Hegner, E., Kröner, A., 2000. Review of Nd isotopic data and xenocrystic and detrital<br />

zircon ages from the pre-Variscan basement in the eastern Bohemian Massif:<br />

speculations on palinspastic reconstructions. In: Franke, W., Altherr, R., Haak, V.,<br />

Oncken, O., Tanner, D. (Eds.), Orogenic Processes — Quantification and Modelling in<br />

the Variscan Belt of Central Europe. Special Publication, vol. 179. Geological Society<br />

of London, pp. 113–129.<br />

Hegner, E., Walter, H.J., Satir, M., 1995. Pb–Sr–Nd isotopic compositions and trace<br />

element geochemistry of megacrysts and melilitites from the Tertiary Urach<br />

volcanic field: source composition of small volume melts under SW Germany.<br />

Contributions to Mineralalogy and Petrology 122, 322–335.<br />

Hickman, A.H., <strong>Van</strong> Kranendonk, M.J., 2004. Diapiric processes in the formation of<br />

Archaean continental crust, East Pilbara Granite-Greenstone Terrane, Australia. In:<br />

Eriksson, P.G., Altermann, W., Nelson, D.R., Mueller, W.U., Catuneau, O. (Eds.), The<br />

Precambrian Earth: Tempos and Events. Elsevier, Amsterdam, pp. 54–75.<br />

Hoskin, P.W.O., Schaltegger, U., 2003. The composition of zircon and igneous and<br />

metamorphic petrogenesis. Review of Mineralogy and Geochemistry 53, 27–62.<br />

Huebeck, C.E., Lowe, D.R., 1994. Depositional and tectonic setting of the Archean Moodies<br />

Group, Barberton Greenstone Belt, South Africa. Precambrian Research 68, 257–290.<br />

Jacobsen, S.B., Wasserburg, G.J., 1980. Sm–Nd isotopic evolution of chondrites. Earth and<br />

Planetary Science Letters 50, 139–155.<br />

Jaeckel, P., Kröner, A., Kamo, S.L., Brandl, G., Wendt, J.I., 1997. Late Archaean to early<br />

Proterozoic granitoid magmatism and high grade metamorphism in the central<br />

Limpopo Belt, South Africa. Journal of the Geological Society of London 154, 25–44.<br />

Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bentley, W.C., Essling, A.M., 1971. Precision<br />

measurements of half-lives and specific activities of 238 U and 235 U. Physical Reviews<br />

4, 1889–1906.<br />

Kamo, S.L., Davis, D.W., 1994. Reassessment of Archean crustal development in the<br />

Barberton Mountain Land, South Africa, based on U–Pb dating. Tectonics 13,<br />

167–192.<br />

Karabinos, P., 1997. An evaluation of the single-grain zircon evaporation method in<br />

highly discordant samples. Geochimica et Cosmochimica Acta 61, 2467–2474.<br />

Kinny, P.D., 1986. 3820 Ma zircons from a tonalitic Amitsoq gneiss in the Godthab<br />

district of southern West Greenland. Earth and Planetary Science Letters 79,<br />

337–347.<br />

Kisters, A.F.M., Anhaeusser, C.R., 1995. Emplacement features of Archaean TTG plutons<br />

along the southern margin of the Barberton greenstone belt, South Africa.<br />

Precambrian Research 75, 1–15.<br />

Kisters, A.F.M., Stevens, G., Dziggel, A., Armstrong, R.A., 2003. Extensional detachment<br />

faulting and core complex formation in the southern Barberton granite-greenstone<br />

terrain, South Africa: evidence for a 3.2 Ga orogenic collapse. Precambrian Research<br />

127, 355–378.<br />

Kober, B., 1986. Whole-grain evaporation for 207 Pb/ 206 Pb-age investigations on single<br />

zircons using a double-filament thermal ion source. Contributions to Mineralogy<br />

and Petrology 93, 482–490.<br />

Kober, B., 1987. Single-zircon evaporation combined with Pb + emitter-bedding for<br />

207 Pb/ 206 Pb-age investigations using thermal ion mass spectrometry, and implications<br />

to zirconology. Contributions to Mineralogy and Petrology 96, 63–71.<br />

Krogh, T.E., 1973. A low-contamination method for hydrothermal decomposition of<br />

zircon and extraction of U and Pb for isotopic age determination. Geochimica et<br />

Cosmochimica Acta 37, 485–494.<br />

Krogh, T.E., 1982. Improved accuracy of U–Pb zircon ages by the creation of more<br />

concordant systems using an abrasion technique. Geochimica et Cosmochimica<br />

Acta 46, 637–649.<br />

Kröner, A., 2007. The Ancient Gneiss Complex of Swaziland and environs: records of<br />

Early Archean crustal evolution in southern Africa. In: <strong>Van</strong> Kranendonk, M.J.,<br />

Smithies, R.H., Bennett, V.C. (Eds.), Earth's Oldest Rocks. Developments in<br />

Precambrian Geology, vol. 15. Elsevier, Amsterdam, pp. 465–480.<br />

Kröner, A., Hegner, E., 1998. Geochemistry, single zircon ages and Sm–Nd systematics of<br />

granitoid rocks from the Góry Sowie (Owl) Mts., Polish West Sudetes: evidence for<br />

early Palaeozoic arc-related plutonism. Journal of the Geological Society of London<br />

155, 711–724.<br />

Kröner, A., Todt, W., 1988. Single zircon dating constraining the maximum age of the<br />

Barberton greenstone belt, southern Africa. Journal of Geophysical Research 93,<br />

15,329–15,337.<br />

Kröner, A., Byerly, G.R., Lowe, D.R., 1991. Chronology of early Archaean granitegreenstone<br />

evolution in the Barberton Mountain land, South Africa, based on<br />

precise dating by single zircon evaporation. Earth and Planetary Science Letters 103,<br />

41–54.<br />

Kröner, A., Hegner, E., Wendt, J.I., Byerly, G.R., 1996. The oldest part of the Barberton<br />

granitoid-greenstone terrain, South Africa: evidence for crust formation between<br />

3.5 and 3.7 Ga. Precambrian Research 78, 105–124.<br />

Kröner, A., Jaeckel, P., Brandl, G., Nemchin, A.A., Pidgeon, R.T., 1999. Single zircon ages for<br />

granitoid gneisses in the Central Zone of the Limpopo belt, southern Africa, and<br />

geodynamic significance. Precambrian Research 93, 299–337.<br />

Kröner, A., Muhongo, S., Hegner, E., Wingate, M.T.D., 2003. Single-zircon geochronology<br />

and Nd isotopic systematics of Proterozoic high-grade rocks from the Mozambique<br />

belt of southern Tanzania (Masasi area): implications for Gondwana assembly.<br />

Journal of teh Geological Society of London 160, 745–757.<br />

Lahaye, Y., Arndt, N., Byerly, G., Chauvel, C., Fourcade, S., Gruau, G., 1995. The influence of<br />

alteration on the trace-element and Nd isotopic compositions of komatiites.<br />

Chemical Geology 126, 43–64.<br />

Lamb, S.H., 1987. Archean synsedimentary tectonic deformation — a comparison with<br />

the Quaternary. Geology 15, 565–568.<br />

Lécuyer, C., Gruau, G., Anhaeusser, Fourcade, S., 1994. The origin of fluids and the effects<br />

of metamorphism on the primary chemical compositions of Barberton komatiites:<br />

new evidence from geochemical (REE) and isotopic (Nd, O, H, 39 Ar/ 40 Ar) data.<br />

Geochimica et Cosmochimica Acta 58, 969–984.<br />

Lopez-Martinez, M., York, D., Hanes, J.A., 1992. A 40 Ar/ 39 Ar geochronological study of<br />

komatiites and komatiitic basalts from the lower Onverwacht volcanics: Barberton<br />

Mountain Land, South Africa. Precambrian Research 57, 481–526.<br />

Lowe, D.R., 1994. Accretionary history of the Archean Barberton Greenstone Belt (3.55–<br />

3.22 Ga), southern Africa. Geology 22, 1099–1102.<br />

Lowe, D.R., Byerly, G.R. (Eds.), 1999. Geological Evolution of the Barberton Greenstone<br />

Belt, South Africa. Special Paper, vol. 329. Geological Society of America. 319 pp.<br />

Lowe, D.R., Byerly, G.R., 2007. An overview of the geology of the Barberton greenstone<br />

belt and vicinity: implications for early crustal development. In: <strong>Van</strong> Kranendonk,<br />

M.J., Smithies, R.H., Bennett, V.C. (Eds.), Earth's Oldest Rocks. Developments in<br />

Precambrian Geology, vol. 15. Elsevier, Amsterdam, pp. 481–526.<br />

Mareschal, J.-C., West, G.F., 1980. A model for Archean tectonism. Part 2. Numerical<br />

models of vertical tectonism in greenstone belts. Canadian Journal of Earth Sciences<br />

17, 60–71.<br />

Moorbath, S., Whitehouse, M.J., Kamber, B.S., 1997. Extreme Nd-isotope heterogenietiy<br />

in the early Archaean — fact or fiction? Case histories from northern Canada and<br />

West Greenland. Chemical Geology 135, 213–231.<br />

Moyen, J.-F., Stevens, G., Kisters, A.F.M., 2006. Record of mid-Archaean subduction from<br />

metamorphism in the Barberton terrain, South Africa. Nature 442, 559–562.


Moyen, J.-F., Stevens, G., Kisters, A.F.M., Belcher, R.W., 2007. TTG plutons of the<br />

Barberton granitoid-greenstone terrain, South Africa. In: <strong>Van</strong> Kranendonk, M.J.,<br />

Smithies, R.H., Bennet, V. (Eds.), Earth's Oldest Rocks. Developments in Precambrian<br />

Geology, vol. 15. Elsevier, Amsterdam, pp. 607–668.<br />

Myers, J.S., Kröner, A., 1994. Archaean tectonics. In: Hancock, P. (Ed.), Continental<br />

deformation. Pergamon Press, Oxford, pp. 355–369.<br />

Nelson, D.R., 1997. Compilation of SHRIMP U–Pb Zircon Geochronology Data, 1996.<br />

Geological Survey of Western Australia. Record 1997/2, 189 pp.<br />

Ramberg, H., 1967. Gravity Deformation and the Earth's Crust. Academic Press, London.<br />

214 pp.<br />

Ramsay, J.G., Huber, M.I., 1987. The Techniques of Modern Structural Geology. Volume 2:<br />

Folds and Fractures. Academic Press. 700 pp.<br />

Rey, P.F., Philippot, P., Thébaud, N., 2003. Contribution of mantle plumes, crustal<br />

thickening and greenstone blanketing to the 2.75–2.65 Ga global crisis. Precambrian<br />

Research 127, 43–60.<br />

Sandiford, M., <strong>Van</strong> Kranendonk, M.J., Bodorkos, S., 2004. Conductive incubation and the<br />

origin of dome-and-keel structure in Archean granite-greenstone terrains: a model<br />

based on the eastern Pilbara Craton, Western Australia. Tectonics 23, TC1009.<br />

doi:10.1029/2002TC001452.<br />

Searle, M.P., 2007. Diagnostic features and processes in the construction and evolution<br />

of Oman-, Zagros-, Himalayan-, Karakoram-, and Tibetan-type orogenic belts. In:<br />

HatcherJr. Jr., R.D., Carlson, M.P., McBride, J.H., Martinez Catalan, J.R. (Eds.), 4-D<br />

Framework of Continental Crust. Memoir, vol. 200. Geological Society of America,<br />

Boulder, Colorado, pp. 41–62.<br />

Smithies, R.H., Champion, D.C., <strong>Van</strong> Kranendonk, M.J., Howard, H.M., Hickman, A.H.,<br />

2005. Modern-style subduction processes in the Mesoarchaean: geochemical<br />

evidence from the 3.12 Ga Whundo intraoceanic arc. Earth and Planetary Science<br />

Letters 231, 221–237.<br />

Spear, F.S., 1993. Metamorphic phase equilibria and pressure–temperature–time paths.<br />

Mineralogical Society of America Monograph, Washington, D.C. 799 pp.<br />

Stacey, J.S., Kramers, J.D., 1975. Approximation of terrestrial lead isotope evolution by a<br />

two-stage model. Earth and Planetary Science Letters 26, 207–221.<br />

Stern, R.A., 1997. The GSC sensitive high resolution ion microprobe (SHRIMP): analytical<br />

techniques of zircon U–Th–Pb age determinations and performance evaluation.<br />

Radiogenic Age and Isotope Studies. Geological Survey of Canada, Report 10,<br />

Current Research. F1–31.<br />

Stevens, G., Moyen, J.-F., 2007. Metamorphism in the Barberton granite-greenstone terrain:<br />

a record of Paleoarchean accretion. In: <strong>Van</strong> Kranendonk, M.J., Smithies, R.H., Bennet, V.<br />

(Eds.), Earth's Oldest Rocks. Developments in Precambrian Geology, vol. 15. Elsevier,<br />

Amsterdam, pp. 669–698.<br />

M.J. <strong>Van</strong> Kranendonk et al. / Chemical Geology 261 (2009) 115–139<br />

Stevens, G., Droop, G.T.R., Armstrong, R.A., Anhaeusser, C.R., 2002. Amphibolite facies<br />

metamorphism in the Schapenburg schist belt: a record of the mid-crustal response<br />

to ∼3.23 Ga terrane accretion in trhe Barberton greenstone belt. South African<br />

Journal of Geology 105, 271–284.<br />

Tegtmeyer, A., Kröner, A., 1987. U–Pb zircon ages bearing on the nature of early<br />

Archaean greenstone belt evolution, Barberton Mountainland, South Africa.<br />

Precambrian Research 36, 1–20.<br />

<strong>Van</strong> Kranendonk, M.J., 2004. Archaean tectonics 2004: a review. Precambrian Research<br />

131, 143–151.<br />

<strong>Van</strong> Kranendonk, M.J., 2007. Tectonics of early Earth. In: <strong>Van</strong> Kranendonk, M.J., Smithies,<br />

R.H., Bennet, V. (Eds.), Earth's Oldest Rocks. Developments in Precambrian Geology,<br />

vol. 15. Elsevier, Amsterdam, pp. 1105–1116.<br />

<strong>Van</strong> Kranendonk, M.J., Collins, W.J., Hickman, A.H., Pawley, M.J., 2004. Critical tests of<br />

vertical vs horizontal tectonic models for the Archaean East Pilbara Granite-<br />

Greenstone Terrane, Pilbara Craton, Western Australia. Precambrian Research 131,<br />

173–211.<br />

<strong>Van</strong> Kranendonk, M.J., Smithies, R.H., Hickman, A.H., Champion, D.C., 2007. Secular<br />

tectonic evolution of Archaean continental crust: interplay between horizontal and<br />

vertical processes in the formation of the Pilbara Craton, Australia. Terra Nova 19,<br />

1–38.<br />

<strong>Van</strong> Vuren, C.F.J., Cloete, M., 1995. A preliminary P–T–t path for the emplacement of the<br />

Theespruit pluton, Barberton Greenstone Belt. In: BartonJr. Jr., J.M., Copperthwaite,<br />

Y.E. (Eds.), Centennial Geocongress. Extended Abstracts, vol. 1. Geological Society of<br />

South Africa, pp. 315–318.<br />

Viljoen, M.J., Viljoen, R.P., 1969. The geology and geochemistry of the lower ultramafic<br />

unit of the Onverwacht Group and a proposed new class of igneous rocks.<br />

Geological Society of South Africa, Special Publication 2, 55–86.<br />

Wasserburg, G.J., Jacobsen, S.B., DePaolo, D.J., McCulloch, M.T., 1981. Precise determination<br />

of Sm/Nd ratios, Sm and Nd isotopic abundances in standard solutions.<br />

Geochimica et Cosmochimica Acta 45, 2311–2323.<br />

Williams, I.S., 1998. U–Th–Pb geochronology by ion microprobe. In: McKibben, M.A.,<br />

ShanksIII III, W.C., Ridley, W.I. (Eds.), Applications of Microanalytical Techniques to<br />

Understanding Mineralizing Processes. Reviews in Economic Geology, vol. 7, pp. 1–35.<br />

Windley, B.F., 1995. The Evolving Continents, third edition. John Wiley and Sons, New<br />

York. 526 pp.<br />

Xuan, H., Ziwai, B., de Paolo, D., 1986. Sm–Nd isotope study of early Archean rocks,<br />

Qianan, Hebi Province, China. Geochimica et Cosmochimica Acta 50, 625–631.<br />

139

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!