You are on page 1of 21

Precambrian Research 99 (2000) 91111

www.elsevier.com/locate/precamres

The Dharwar craton, southern India, interpreted as the result


of Late Archaean oblique convergence
B. Chadwick a, *, V.N. Vasudev b, G.V. Hegde c
a Earth Resources Centre, University, Exeter EX4 4QE, UK
b 120/45(A) III Block, Thyagarajanagar, Bangalore 560 028, India
c Department of Mines and Geology, Government of Karnataka, 16/3-5 S. P. Complex, Lalbagh Road, Bangalore 560 027, India
Received 6 March 1999; accepted 2 July 1999

Abstract
The Dharwar craton comprises two distinct parts separated by a steep NS sinistral shear zone. In the western
part a pre-2900 Ma complex of orthogneisses, granodiorites and narrow tracts of supracrustal rocks (Sargur Group)
forms the basement to volcanic and sedimentary basins of the Dharwar Supergroup (ca. 28002550 Ma). Late
Archaean deformation is characterised by NESW crustal shortening and steep NS sinistral shear zones. The eastern
part is underlain by parallel, steep NS or NWSE linear belts of calc-alkaline, anatectic and juvenile granites and
granodiorites (Dharwar batholith, ca. 27502510 Ma) with intervening volcanic and sedimentary schist belts (ca.
28002550 Ma). The plutonic belts are 1525 km wide, up to 150 km long, and bounded by steep NWSE high-strain
zones up to 2 km wide with sinistral shear sense (except one which is dextral ). Magmatic-state fabrics and structures
in the plutonic rocks are parallel to solid-state sinistral shear fabrics in the high-strain zones, but diffuse magmatic
banding is commonly oblique to these zones and coincides with the plane of instantaneous shortening in sinistral
shear. Magmatic-state structures, swarms of vertical NWSE dykes of granite, and the vertical wedge shape of the
linear belts are consistent with emplacement of the batholith during sinistral shear when magma pressure exceeded
regional horizontal compressive stress. Upright folds and schistosity, steep reverse faults and effects of regional HT/LP
metamorphism show that deformation was partitioned into NESW shortening in the schist belts during emplacement
of the plutonic belts in the sinistral shear regime. The western part of the craton is interpreted as the foreland to an
accretionary arc represented by the batholith and schist belts (intra-arc basins) in the east. NESW shortening and
sinistral transcurrent displacements in the foreland and arc are consistent with arc-normal and arc-parallel displacements
during oblique convergence analogous to MesozoicCenozoic convergent settings. 2000 Elsevier Science B.V. All
rights reserved.
Keywords: Batholith emplacement; Dharwar craton; Foreland deformation; Late Archaean; Oblique convergence

1. Introduction
Convergent settings analogous to those of the
Phanerozoic, but with smaller plates and longer
* Corresponding author. Tel.: +44-1392-263916;
fax: +44-1392-263342.
E-mail address: b.chadwick@exeter.ac.uk (B. Chadwick)

ocean ridge lengths, are widely regarded as the


tectonic environment of Archaean terrains
( Windley, 1995; Condie and Sloan, 1998; de Wit,
1998). In contrast, a minority holds that nonuniformitarian processes related primarily to the
thermal vigour of the Archaean Earth were predominant. The conflict in views is exemplified by
a spirited attack on the plate tectonic consensus

0301-9268/00/$ - see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S0 3 0 1- 9 2 68 ( 9 9 ) 0 00 5 5 -8

92

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

by Hamilton (1998). The Dharwar craton in southern India is a critical example in the current debate.
On the one hand, Choukroune et al. (1995, 1997)
and Chardon et al. (1996, 1998) contend that its
evolution was controlled by non-uniformitarian
sagduction (Goodwin and Smith, 1980), that is,
passive sinking of volcanic and sedimentary basins
into basement gneisses (softened lithosphere) with
no crustal shortening. In this paper we present
new evidence of magmatic- and solid-state structures and fabrics which point to control of the
Late Archaean history of the Dharwar craton
(Fig. 1, Table 1) not by non-uniformitarian processes but by emplacement of a calc-alkaline batholith, NWSE shortening and NWSE sinistral
transcurrent displacements with close analogies in
oblique convergent settings, for example, in the
East Indies, the Andes and the Cordillera of western North America. Of the widely different proposals for the Late Archaean plate tectonic setting of
the Dharwar craton (Chadwick et al., 1997), our
new interpretation is closest to models of Hanson
et al. (1988) and Newton (1990) who envisaged
Cordilleran or Andean margin intrusion and NS
lateral shearing.

2. Geology of the Dharwar craton


The craton comprises a western domain
underlain by orthogneisses and granodiorites [ca.
29003300 Ma, collectively termed Peninsular
Gneiss: Beckinsale et al. (1980); Taylor et al.
(1984); Bhaskar Rao et al. (1991); Peucat et al.
(1993)] interspersed with older tracts of metasedimentary and metamorphosed igneous suites
[Sargur Group, Swami Nath et al. (1976)]. This
association forms the basement to mixed-mode
[sensu Gibbs (1987)], volcanic and sedimentary
basins (schist belts) of the low-grade Late
Archaean Dharwar Supergroup [ca. 2900
2600 Ma; Swami Nath et al. (1976); Chadwick
et al. (1991, 1992); Nutman et al. (1996); Kumar
et al. (1996); Trendall et al. (1997a,b)]. The western domain is locally intruded by granite plutons
[ca. 2600 Ma, Taylor et al. (1984); Rogers (1988)].
The eastern part of the craton is dominated by
a Late Archaean calc-alkaline complex of juvenile

and anatectic granites, granodiorites, monzonites


and diorites. They are interspersed with schist belts
which are lithologically similar to the Dharwar
Supergroup in the west, but a lack of precise ages
precludes direct correlation (Chadwick et al., 1996;
Nutman et al., 1996). The plutonic complex was
termed the Dharwar batholith by Chadwick et al.
(1996) on the grounds of the close lithological
similarity, structural coherence and emplacement
of its principal components in the period ca. 2750
2510 Ma ( Table 1). High-grade metamorphism
leading to charnockites took place close to
2500 Ma in the south of the batholith (Hansen
et al., 1997): lower grade metamorphic recrystallisation of limestone in the Sandur belt took place
at about the same time [2475+65 Ma; Russell
et al. (1996)].
The Dharwar batholith includes the so-called
Closepet Granite, a term introduced in 1901
( Fig. 1; Smeeth, 1915; B.P. Radhakrishna, personal communication, 1994) for a narrow linear
tract of granites and gneisses trending NS or
NWSE close to the western boundary of the
batholith. The tract has figured prominently on
published maps for many years. Allen et al. (1986)
and Jayananda et al. (1995) described it as the
Closepet batholith, but with no justification. We
regard the terms Closepet Granite and Closepet
batholith as unsatisfactory because the belt is not
a single granite and, moreover, it is no more than
a small part of the much larger system of plutonic
rocks forming the Dharwar batholith. This larger
grouping is better described as a batholith (Pitcher,
1979; Bateman, 1992) in the sense that it is a
cluster of plutons and plutonic belts whose magma
generation and intrusion were controlled by a
crustal scale event. Chadwick et al. (1996) attributed this event to Late Archaean oblique convergence that is addressed in this paper.
The southern part of the batholith has been the
prime target of petrogenetic studies. Early work
viewed metasomatism (granitisation) of older
gneisses as the principal genetic mechanism
(Radhakrishna, 1956), but more recent studies in
the southwestern marginal zone revealed the prevalence of partial melting of older gneisses,
>2900 Ma (Friend, 1983; Friend and Nutman,
1991; Oak, 1990; Jayananda and Mahabaleshwar,

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

93

Fig. 1. (A) Regional map of the Dharwar batholith and part of the foreland in the Dharwar craton in Karnataka and southern
Andhra Pradesh, showing the principal schist belts and the sites of Figs. 2 and 4. (B) Position of the Dharwar craton in
southern Peninsular India.

94

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

Table 1
Salient isotopic age data (with error bars) related to principal geological divisions and Late Archaean oblique convergence in the
Dharwar craton in Karnataka and Andhra Pradesha

a Notes: (a) Nutman et al. (1996): 1, acid volcanic rocks, Sandur schist belt; 2, Joga granite; 3, Belagallu Tanda gneiss; 4, acid
volcanic rocks, Daginkatte. (b) A.P. Nutman, personal communication (1998): 1, clast of granodiorite in polymict conglomerate,
Hutti schist belt; 2, Koppal syenite; 3, granodiorite west of Gooty. (c) Zacharaiah et al. (1995): metabasalts, Ramagiri schist belt;
(d) Balakrishnan et al. (1990): tholeiite, Kolar schist belt; (e) Peucat et al. (1993): granulite facies metamorphism, Krishnagiri, in
the extreme south of the Dharwar batholith; (f ) Friend and Nutman, 1991: 1, anatectic granite, Ramnagaram; 2, gneiss, Ramnagaram.
(g) Krogstad et al. (1991, 1995): 1, Dod Gneiss; 2, Dosa Gneiss; 3, Patna Granite; 4, Kambha Gneiss; all adjacent to the Kolar
schist belt. (h) Subba Rao et al. (1998): 1. Hampi granite; 2. Lepakshi granite. i. Taylor et al. (1984): 1, Chitradurga granite; 2, acid
volcanic rocks, Daginkatte, E of Honnali; ( j) Bhaskar Rao et al. (1992): 1, Chitradurga granite; 2, Volcanic suite, Shimoga district.
(k) Trendall et al. (1997a): acid volcanic rocks, Daginkatte, E of Honnali. ( l ) Trendall et al. (1997b): ash-fall tuffs, Bababudan. (m)
Kumar et al. (1996): volcanic suites, Bababudan. (n) Stroh et al. (1983): Halekote trondhjemite; (o) Monrad (1983); (p) Beckinsale
et al. (1980); (q) Nutman et al. (1992): 1, metamorphic zircon. (r) Peucat et al. (1995): rhyolite, Holenarsipur. (s) Ramakrishnan
et al. (1994): detrital zircon.

1991), which generated granitic melts that


interacted to various degrees with contemporaneous intermediate magmas of mantle origin

(Jayananda et al., 1995; Moyen et al., 1997). In


contrast, Krogstad et al. (1991, 1995), Peucat et al.
(1993) and Hansen et al. (1995) identified major

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

episodes of juvenile crustal accretion, ca. 2550


2510 Ma, that took place in association with granulite facies metamorphism in the extreme south of
the batholith. These isotopic and geochemical data
point to the predominance of granites derived by
partial melting of older gneisses (>2900 Ma) in
the western portion of the batholith in contrast
with juvenile granites and related suites which
predominate in the east (Chadwick et al., 1996).
The eastern and western parts of the Dharwar
craton are separated by a steep sinistral shear zone
which includes mylonitised batholith granites and
volcanic and sedimentary rocks of the Dharwar
Supergroup: the zone is up to 23 km wide and
extends for ca. 400 km along the eastern boundary
of the GadagChitradurga schist belt ( Fig. 1;
Chadwick et al., 1992).

3. Late Archaean basin development and


deformation in the western part of the craton
(foreland)
Facies distributions and well-preserved volcanic
and sedimentary structures show that basin development of the Dharwar Supergroup between
Chikmagalur and Ranibennur (Fig. 1) had much
in common with mixed-mode basins (Gibbs, 1987)
generated in transpression (Chadwick et al., 1989,
1992). Initial crustal extension led to shallow
marine+fluvial sedimentation (quartzpebble conglomerates, quartzites, argillites, banded iron formations) and eruption of basalts with back-arc
affinity ca. 29002700 Ma ( Table 1; Bhaskar Rao
and Drury, 1982; Drury, 1983; Anantha Iyer and
Vasudev, 1985). Later fluvial and marine sedimentation (greywackes, polymict conglomerates
with intra-basinal and basement clasts, banded
iron formations, stromatolithic limestones) was
accompanied by further basaltic and rhyolitic volcanism. Some of the later sequences are at least
7000 m thick and reflect variable uplift and substantial syn-depositional subsidence of the older
Dharwar cover and basement gneisses (Chadwick
et al., 1985a, 1988, 1989, 1991).
Deformation took place at ca. 2600 Ma as
indicated by fabrics younger than rhyolites and
the cross-cutting relationship of the Chitradurga

95

Granite ( Table 1). Reverse faults and thrust


wedges of basement gneisses with top-to-SW SC
fabrics, large-scale asymmetric folds with SW
vergence, lobe and cusp contacts between the
Dharwar cover and its basement (Chadwick et al.,
1991), and a previously unreported top-to-SW
thrust stack of mylonitised basement gneisses and
Dharwar cover rocks in the Gadag region (Figs. 2
and 3) are all consistent with NESW crustal
shortening. Apart from zones of ductile mylonites,
brittle-ductile deformation of the bulk of the basement gneisses on myriad fractures and small retrograde shear zones contrasts with the pervasive
ductile behaviour of the cover.
Structures formed during NESW shortening
are deformed by N- or NE-trending folds and
crenulation fabrics that are consistent with sinistral
displacements in the steep NS shear zones within
the foreland and along its eastern boundary
( Fig. 1; Chadwick et al., 1985b, 1988, 1989, 1991,
1992). The link between folding, schistosity and
transcurrent shear is evident in transitions from
the mylonitised cover into mylonitised batholith
granites in the steep sinistral shear zone flanking
the east of the Gadag-Chitradurga schist belt.
Calcite in veins in the basement and as porphyroblasts in the cover suggests CO was an important
2
volatile phase during NESW shortening and N
S sinistral shearing.
Contrary to the widespread evidence of NE
SW shortening, Chardon et al. (1996, 1998)
claimed that the Late Archaean structure in the
west of the craton resulted from passive sinking
of the Dharwar Supergroup into its basement
gneisses with no crustal shortening. They envisaged
two distinct episodes of diapirism, ca. 3000 and
ca. 2600 Ma, under the influence of mantle plumes
and sagduction. They based their case for Late
Archaean sagduction on structures at the base of
the Dharwar Supergroup in the Kibbanahalli Arm
and the Bababudan schist belt (Fig. 1), namely,
mylonitic extensional structures at the boundary
between the basement gneisses and the Dharwar
cover, cleavage more shallow than bedding in
quartzpebble conglomerates at the base of the
supergroup, and patterns of fabric trajectories.
However, cleavage steeper than bedding immediately above the basal contact, and the form and

96

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

Fig. 3. Shallow mylonitic basement gneiss showing top-to-SW


shear sense. Outcrop approximately midway between Lakkundi
and Dambal (Fig. 2).

Fig. 2. Map and NESW section across the north of the Gadag
schist belt showing thrust stacking of the Dharwar Supergroup
and its basement gneisses.

asymmetry of widespread folds older than those


generated by NS sinistral shear in the
Kibbanahalli and Bababudan belts are consistent
with NESW shortening. Moreover, in the south
of Bababudan the shallow cleavage at the base of
the supergroup is deformed by small-scale contractional folds with SW vergence consistent with
NESW shortening. Chardon et al. (1998) claimed

that the reverse fault forming the northern boundary of the Bababudan schist belt is a refolded
extensional structure, contrary to its overthrust
relationship with the underlying cover which is
consistent with a similar reverse fault north of the
Kaldurga syncline (Chadwick et al., 1985b). The
sagduction model implies that the Sirankatte dome
of basement gneisses NE of Kibbanahalli (Fig. 1)
is also a diapiric structure, but we have found
abundant NW-trending, upright contractional
crenulations superimposed on migmatitic banding
which are consistent with ENEWSW shortening.
Although we agree with Naha et al. (1995) that
the dome is not a diapir, we do not share their
view that the basement was extensively remobilised. On the basis of fabric trajectories, Chardon
et al. (1996) modelled the Kibbanahalli syncline
as widening with depth and having a flat base, but
neither of these characteristics are indicated by its
steep wedge shape at its NW extremity. There is
also no exposed evidence of a flat floor to the
Bababudan basin (see Chardon et al., 1998).
The claim that the Dharwar Supergroup sank
passively into its basement is thus open to question
because it neglects widespread evidence of NE
SW regional shortening. Moreover, the sagduction
model takes no account of the NS sinistral shearing and refolding of regional folds developed

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

during NESW shortening, for example, the main


Bababudan syncline, the Kaldurga syncline in NE
Bababudan and the Nidnegal anticline (Chadwick
et al., 1985b; Mukhopadhyay, 1986; Chadwick
et al., 1991). Substantial syn-depositional subsidence and intra-basinal clasts with pre-depositional
schistose fabrics in the polymict conglomerates in
the upper part of the Dharwar stratigraphy
(Chadwick et al., 1991) point to the local extensional structures at the base of parts of the
Dharwar Supergroup being effects not of sagduction but basin subsidence that occurred during
accumulation of major thicknesses of younger
sequences in the supergroup, for example, the
Kaldurga conglomerate (Chadwick et al., 1985a).

4. The Dharwar batholith and its schist belts


Emplacement of the Dharwar batholith coincided with the transition from predominantly
tonalitetrondhjemitegranodiorite ( TTG) suites
in earlier periods of crustal evolution to greater
abundances of granite and granodiorite in new
additions to continental crust in the Palaeoproterozoic and thereafter (Martin, 1994;
Sylvester, 1994). This transition marked the beginning of a persistent uniformitarian pattern of
granitegranodiorite batholith construction at
convergent plate boundaries ( Windley, 1995).
Insight into the mechanisms of batholith development has been revolutionised by recent radical
changes in ideas of granite generation, melt
transfer and emplacement (Brown, 1994; Pitcher,
1997), based on numerous illustrations of magma
ascent and emplacement controlled by crustal
scale, compressional and extensional shear zones
and fractures (Clemens and Mawer, 1992; Hutton,
1988, 1992, 1996, 1997; Ingram and Hutton, 1997;
Tikoff and Teyssier, 1992; Petford, 1996). At the
same time, the genesis of magmatic- and solidstate fabrics featured prominently in new theoretical and field studies of granite rheology (Hutton,
1988; Paterson et al., 1989; Fernandez and
Barbarin, 1991; Fernandez and Gasquet, 1994;
Saint Blanquet and Tikoff, 1997). In this paper
we use these new developments as the basis for

97

the interpretation of the growth of the Dharwar


batholith.
Based on the foregoing references, recognition
of magmatic-state fabrics in the batholith hinged
on preferred shape orientations of subhedral plagioclase, K-feldspar, biotite and hornblende, and
interstitial aggregates of coarse-grained quartz
with minimal preferred orientation and few recrystallisation textures. In contrast, solid-state (crystalplastic) fabrics are indicated primarily by winged
porphyroclasts of K-feldspar and ribbons of
fine-grained quartz. Magmatic-state fabrics are
widespread, but common transitions from
magmatic- to solid-state fabrics are indicated by
shape modification and internal distortion of feldspars, biotite and quartz aggregates,
Our evidence is drawn mainly from an area of
ca. 26 000 km2 in NE Karnataka and SW Andhra
Pradesh ( Fig. 4) where the batholith and schist
belts crop out on a rolling plain with an altitude
of ca. 600 m. Exposure is extremely variable, but
linear ranges of granite tors and numerous quarries
opened in the last 20 years for aggregate and
dimensional stone provided abundant, previously
unreported data for structural analysis and classification of the plutonic suites.

4.1. Plutonic belts


The batholith comprises a series of parallel,
steep elongate belts analogous to, but smaller than,
the plutonic belts of subduction-related Mesozoic
granites in Sumatra (McCourt et al., 1996). The
plutonic belts in the Dharwar batholith are elongate vertical wedges ca. 1525 km wide, ca. 75
150 km long, trending NWSE, but swinging to
NS further south ( Figs. 1 and 4). They taper
upward ( Fig. 5) and along strike into interdigitated
or rounded contacts with the schist belts. Outcrop
areas and steep magmatic fabrics suggest vertical
dimensions of at least a few kilometres, but magnetic and gravity data (Mahadevan, 1994) and a
deep seismic profile ( Kaila et al., 1979; Kaila and
Sain, 1997) provide no clues to the vertical form:
the seismic profile reveals only sub-horizontal
reflectors of uncertain significance. Each belt is
bounded by steep contacts with schist belts or

98

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

Fig. 4. Principal trends of magmatic banding ( long dashes) and steep zones of high-strain orthogneiss and mylonite (short dashes
with shear sense) in the Dharwar batholith in NE Karnataka and SW Andhra Pradesh. Blank areas in the plutonic belts indicate
uncertainties in linking dispersed outcrops of granite and related suites. Cross-section XY is shown in Fig. 5.

zones of mylonitised granite and high-strain orthogneisses. Some belts are homogeneous, but most
are mixtures of multipulse granites and diorites.
Plutonic compositions are based on petrography
and our unpublished chemical analyses. Isotopic
age data are meagre, but available data point to
emplacement of the bulk of the batholith ca. 2600
2500 Ma (Table 1).

Syntectonic metamorphic PT conditions related


to granite intrusion adjacent to the Sandur schist
belt (Hanuma Prasad et al., 1996) indicate depth
of emplacement of 1317 km, broadly in line with
depths of 1015 km indicated by assemblages in
the SW of the belt ( Harris and Jayaram, 1982).
Deeper levels of emplacement in the south of the
batholith are indicated by the close association of

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

Fig. 5. SWNE section across the central part of the Sandur


schist belt ( Fig. 4) showing the saw-tooth boundary resulting
from interdigitation of steep sheets of granite with metabasalts,
metagabbros and tremolite schists of the Sandur Group in the
crest of a plutonic wedge (after Chadwick et al., 1996).

emplacement and granulite facies metamorphism


(Peucat et al., 1993; Hansen et al., 1995).
4.1.1. Western margin of the batholith
Sub-horizontal migmatitic orthogneisses form a
belt ca. 25 km wide and extending NS for ca.
400 km immediately east of the shear zone separating the two parts of the craton. On the grounds
of their TTG composition (our unpublished data)
and limited isotopic ages ( Friend and Nutman,
1991), the gneisses are correlated with the
pre-2900 Ma gneisses of the basement in the west
of the craton. In contrast with the gneisses immediately west of the shear zone, they display variable
degrees of partial melting with transitions to pink
granites with partly digested remnants of the old
gneiss percursor. Swarms of criss-crossing pegmatite dykes and steep NS wedges of granite with
linear and planar magmatic-state and crystalplastic fabrics intrude the belt of shallow gneisses.
For these reasons, the 25 km wide tract of subhorizontal older gneisses and metatexites is
included in the Dharwar batholith.
4.1.2. Older orthogneiss with mafic dykes
Other older orthogneisses are found in a few of
the plutonic belts. They are similar to the highstrain, granodioritic, Belagallu Tanda gneiss
[2719+40 Ma; Nutman et al. (1996); Chadwick
et al. (1996)] which at the type locality west of
Bellary is a streaky, banded, biotitic migmatite

99

with an intense tectonic S fabric and coplanar


seams of pegmatite. It is cut by deformed mafic
dykes and includes xenoliths of quartzite and
amphibolite. Grey gneisses of Belagallu Tanda
type occur in other belts as xenoliths in granites
and as extensive tracts intersheeted with younger
granites. They also form a major part of the
batholith south of Anantapur and west of Kolar.
Similar orthogneisses east of Ramagiri are folded
with the schist belt in a large S-plunging antiform.
Appinite dykes (our unpublished data) cut the
migmatised gneisses, but were disrupted during
partial melting of their host.
4.1.3. Diorites
Enclaves of diorite are common in the grey
granodiorites described below, and a major body
of quartz monzonite to diorite composition, here
called Kallur diorite, occurs SW of the Raichur
schist belt which defines part of its eastern boundary. The western boundary of the Kallur diorite is
defined by a steep zone of mylonitised granites
and high-strain orthogneisses whose protoliths
were a mixture of the diorite and younger intersheeted granites. Diorite sensu lato in the SE of
the belt grades into porphyritic hornblende granites s.l.: there appears to be a similar gradation in
the NW of the belt. The diorite is cut by steep
NWSE dykes of aplogranite and pegmatite.
At the type locality at Kallur the diorite is dark,
medium-grained with common ellipsoidal enclaves
of coarse-grained hornblende diorite and finegrained diorite with blue blebs of plagioclase: the
enclaves and their host have an intense vertical S
fabric trending NW. Elsewhere the Kallur diorite
is fine-grained with enclaves of mafic diorite up to
2 m in size. Hornblende and plagioclase (commonly sericitised ) have a magmatic-state preferred
orientation, but a superimposed disequilibrium
texture appears to be of post-magmatic tectonic
origin.
4.1.4. Granodiorite
Grey hornblende granodiorite (including monzonitic facies) with mafic enclaves of feldsparphyric
or even-grained, hornblende diorite equivalent to
the microgranular mafic enclaves (MME) and polygenic swarms of Barbarin and Didier (1991) pre-

100

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

dominates in the largest plutonic belt between


Guntakal and the Hutti schist belt. Similar granodiorite is found W of the Hutti schist belt, NW of
Raichur, and S of Anantapur. Granodiorite W of
Gooty has yielded a SHRIMP UPb zircon age of
2580+31 Ma ( Table 1; A.P. Nutman, personal
communication, 1998). Large tracts of hornblendite, hornblende gabbro and gabbroic diorite are
common in the granodiorites W of the Hutti belt,
and appinite dykes displaying synmagmatic disruption, including the Sederholm effect, feature elsewhere. West of Anantapur the older orthogneisses
are cut by granodiorite with appinite dykes.
Mafic enclaves in the granodiorites commonly
have prolate ellipsoidal shapes with long axes
coinciding with magmatic-state hornblende lineations plunging gently NW or SE. Their relatively
smooth, rounded forms and lobate boundaries are
consistent with commingling of diorite and granodiorite melts. Some with angular shapes appear to
be disrupted synmagmatic dykes or detached fragments of solid diorite. Solid-state deformation
superimposed on the magmatic fabrics in the
granodiorites is indicated by distortion and sericitisation of plagioclase. Chloritisation of hornblende
and biotite appears to be most pronounced in the
granodiorite SE of the Hutti schist belt compared
with elsewhere.
4.1.5. Granites
Granites sensu stricto range from grey, through
grey-pink to brick red, porphyritic and evengrained types. Microcline megacrysts, zoned plagioclase and quartz with variable proportions of
biotite (commonly chloritised) and hornblende
occur with accessory apatite, monazite, epidote,
allanite, titanite, zircon, magnetite, pyrite and
molybdenite: trace amounts of fluorite and carbonate are almost universal, but some granites are
rich in coarse-grained fluorite. Muscovite granites
are rare. Tourmaline occurs in the groundmass of
some granites or in late pegmatites, and spodumene has been recorded in pegmatites NW of the
Hutti schist belt (Devaraju et al., 1990) and
elsewhere.
Steep NWSE wedges of uniform granite are
evident in some plutonic belts, but most granites
are multipulse mixtures of different pale and dark

grey, biotite- or hornblende-rich phases. Granites


NW and SE of the Sandur schist belt include
mixtures of homogeneous grey granite with large
enclaves of porphyritic granite and rounded immiscible inclusions of quartz monzonite. This variety
of granite [ Toranagallu type, Chadwick et al.
(1996)] has yielded a whole rock RbSr age of
2452+50 Ma (Bhaskar Rao et al., 1992) which
appears to be a cooling age on the grounds of the
older SHRIMP age of the Joga granite described
below ( Table 1). Similar granites are found at least
as far south as Kalyandurg. Although some granites include commingled immiscible dark phases,
mafic dioritic enclaves and synmagmatic dykes of
the appinite and lamprophyre suites are less
common in the granites s.s. compared with the
granodiorites. Most granites are younger than the
granodiorites ( Table 1), but available isotopic age
data are limited to granites in the western part of
the batholith.
Magmatic banding, manifest by variations in
biotite, feldspar and quartz abundances, is prominent in granites in many plutonic belts (Fig. 6). It
varies from almost imperceptible, through diffuse
to sharply defined layering a few centimetres to
ca. 1 m thick which persists along strike for many
metres. Some banding is enhanced by coplanar
seams of pegmatite or is impersistent, with gradations into diffuse, elongate schlieren. Outcrop evidence suggests the banding was the result of
different processes that acted together or independently, including smearing of commingled pale
and dark granite melts into layers and lenses
during magmatic flow; dispersion of enclaves and
their mixing to yield schlieric lenses and wisps (cf
Blake and Koyaguchi, 1991); injections of successive sheets of pale and dark granites yielding welldefined or diffuse, persistent or impersistent layering; injection or diatexis leading to banding defined
by thin seams of pegmatite; concentration of crystals by mechanical or chemical effects associated
with magmatic flow s.l. which included sorting of
minerals by magmatic currents, separation of minerals and melt by filter pressing, and transposition
of pre-existing banding. Sporadic felsic ocelli with
cores of titanite are consistent with precipitation
of specific minerals consequent on magma replen-

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

101

Fig. 7. Magmatic-state NWSE shear zone with sinistral displacement in diffusely banded grey granite; arrows highlight
displaced banding. Chandrabanda, 10 km E of Raichur.

Fig. 6. Diffuse magmatic banding in grey granite trending NW


SE, 40 km E of Raichur.

ishment, mixing and other changes in PT conditions (Hibbard, 1995).


Magmatic-state fabrics are of S, LS or L type.
Fabrics and banding commonly coincide, but magmatic-state fabrics were also impressed across
banding or other compositional boundaries such
as occur in plutons in the Cascades Mountains,
Washington (Paterson et al., 1996). Magmaticstate S fabrics and banding in some belts trend
NWSE, although banding is also irregular at
outcrop scale as a result of magmatic-state folding.
Mafic enclaves have a shallow NWSE elongation
parallel to the magmatic linear fabric, for example
in the granite at Bellary. Banding and S fabrics
also dip steeply, with NESW strike oblique to
the NWSE trend of the belts (Fig. 4). Steep axial
surfaces of magmatic-state, open to isoclinal folds
of banding with coplanar S fabrics trending NW
SE indicate NESW shortening, but others are
oblique to the regional NWSE trend. Steep magmatic shear zones ( Fig. 7), equivalents of intramagmatic shear zones (isz) of Fernandez et al.

(1997), are up to 1 m wide with displacements up


to 6 m indicated by displaced banding and trains
of small-scale asymmetric folds. Shear zone trends
and displacement directions vary, but NWSE
zones with sinistral shear sense consistent with that
in the steep border zones of mylonite predominate.
4.1.6. Late plutons of granite and syenite
Sporadic small plutons form late components
of the plutonic belts. The post-tectonic, porphyritic
Joga granite, with scattered rounded enclaves of
mafic syenite or monzonite rich in fluorite, cuts
schistosity in metabasalts with interbedded cherts
in the central NW of the Sandur schist belt
(Chadwick et al., 1996). The granite has a
SHRIMP UPb zircon age of 2570+62 Ma
( Table 1; Nutman et al., 1996).
The Koppal syenite with microcline, albitic plagioclase, hornblende and pale green clinopyroxene
has an elliptical outcrop elongated NESW
( Fig. 4). Its SE boundary is concordant with steep
banding in the host granites: boundary relationships in the NW are not exposed. The upright
asymmetric wedge shape of the pluton is indicated
by variations in the orientation of magmatic banding and parallel S fabrics with lineated hornblende

102

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

plunging down dip. The syenite has a SHRIMP


UPb zircon age of 2526+8 Ma ( Table 1; A.P.
Nutman personal communication, 1998).
4.2. High-strain orthogneisses and mylonites
The steep NWSE high-strain zones between
the plutonic belts are characterised by orthogneisses and mylonites derived from adjacent undeformed granites or granodiorites. Mylonitised
granites are common adjacent to the schist belts.
The high-strain zones are up to 2 km wide and
traceable along strike for up to 110 km: strike
lengths may be much longer, but exposure is
incomplete. Steep SC fabrics, winged porphyroclasts of microcline, shallow lineations (mainly
quartz), small-scale S-folds and extensional shear
bands (Fig. 8) are consistent with sinistral shear
in all the high-strain zones apart from that east of
the Kushtagi schist belt where the displacement
was dextral. The reason for the different shear
sense is uncertain.
Stable hornblende and biotite in the mylonites
and high-strain orthogneisses show that ductile
deformation in the shear zones took place at

Fig. 8. Extensional shear band indicating sinistral shear in steep


NWSE high-strain orthogneiss derived from banded granite;
a sub-horizontal mineral lineation lies on the gneiss banding.
15 km NE of Adoni.

relatively high temperatures. Accessory chlorite


replacing biotite may be the result of later ductile
movement or younger brittle fracturing that led to
NE joints with veneers of epidote and chlorite
which are common throughout the craton.
Magmatic-state folds indicating NESW shortening, sinistral NWSE magmatic shear zones, NW
SE magmatic-state L and S fabrics parallel to
solid-state fabrics in the shear zones, and transitions from NESW magmatic-state S fabrics to
NWSE solid-state fabrics ( Fig. 4) point to close
links between emplacement in the plutonic belts
and solid-state sinistral shear.
4.3. Schist belts
Volcanic and sedimentary rocks in the schist
belts ( Fig. 4) comprise greywackes, polymict conglomerates, banded iron formations, subordinate
orthoquartzites and limestones (Chadwick et al.,
1996), and bimodal basaltrhyolite arc associations in which basalts predominate (Anantha Iyer
and Vasudev, 1979; Zacharaiah et al., 1996;
Hanuma Prasad et al., 1997). The schist belts were
deformed and metamorphosed at greenschist to
amphibolite facies ( Roy, 1979; Roy and Biswas,
1979; Harris and Jayaram, 1982; Hanuma Prasad
et al., 1996). Lower grade assemblages characterise
the interiors of larger belts, whereas their margins
adjacent to the plutonic belts are at amphibolite
facies. Syntectonic minerals in the Sandur belt
include cordierite, gedrite, garnet, andalusite, staurolite, biotite, muscovite and chlorite in pelites,
and hornblende, garnet and plagioclase in metabasalts. Metamorphic T 550600C and P 45.2 kbar
are correlated with granite emplacement (Hanuma
Prasad et al., 1996).
With a few exceptions (Roy, 1979; Roy
and Biswas, 1983; Mukhopadhyay, 1989;
Mukhopadhyay and Matin, 1993), the detailed
structure of most schist belts is largely unknown.
Recent work in the Sandur belt has revealed
thickening of the stratigraphy on reverse faults
dipping steeply NE which are marked by narrow
zones of mylonite and low-angle discordances with
stratigraphic markers such as banded iron formations (Chadwick et al., 1996). Orientations of L
and S fabrics show that faulting was broadly

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

contemporaneous with large-scale, tight upright


folds. Associated schistosity contains a steep finite
elongation lineation marked by the long axes of
ellipsoidal clasts and pillows. Outcrop patterns
and curved hingelines suggest the folds are regional
sheath folds.
Magmatic- and solid-state fabrics in the granites
immediately adjacent to the Sandur schist belt
have trends in common with those in the schist
belt, but shallow linear fabrics predominate in the
plutonic belts compared with steep lineations in
the schist belts. Since deformation and metamorphism in the schist belts were contemporaneous
with emplacement of the plutonic belts, we attribute the difference in lineation plunge to partitioning of regional deformation into NESW
shortening in the schist belts during sinistral
transcurrent displacements in the plutonic rocks.
The shape of the Jonnagiri schist belt (Fig. 4)
suggests it was controlled by sinistral shear along
the eastern boundary of the granodiorite between
Gooty and the Hutti schist belt. Moreover,
Zachariah et al. (1996) implied sinistral shearing
to account for the collage of distinct blocks of
amphibolites and intersheeted granites in the
Ramagiri schist belt. Every schist belt is intruded
to different degrees by syn- and post-tectonic sheets
of granite and pegmatite.

103

planar fabrics in common with those in the pale


grey granite host ( Fig. 9). These characteristics
indicate emplacement while the host was hot and
ductile [crystallinity >ca. 70%; Fernandez and
Barbarin (1991)], but with sufficient rigidity to
fracture. The quarry also exposes vertical sheets
of granite breccia trending NWSE, ca. 20 m wide
with exposed lengths of ca. 500 m (Fig. 10). The
grey granite matrix with diffuse magmatic fabrics
is host to packed inclusions of pale granite a few
metres in size. The form of thin apophyses of
matrix granite along the sheet boundaries points
to emplacement during shear fracture of the hot
host: accessory fluorite and carbonate suggest that
fracturing was facilitated by high volatile pressure.
Vertical granite dykes elsewhere, for example, S of
Raichur, have pull-apart structures consistent with
sinistral shear during intrusion ( Fig. 11).
Vertical NWSE dykes of pale granite, dark
microgranite, aplite, pegmatite and commingled
granite and diorite with widths 0.3ca. 100 m were
also emplaced late in the growth of the plutonic
belts when the host rocks were cooler and more
brittle. Some dykes occupy vertical conjugate shear
fractures formed during EW shortening, but most
were controlled by NWSE fractures. NWSE
dykes have sharp boundaries and little internal
deformation: irregular lobate masses of immiscible

4.4. Emplacement of the plutonic belts


Magmatic-state planar fabrics and banding are
steep with either NWSE or NESW trends,
whereas magmatic-state linear fabrics mostly have
shallow NW or SE plunges. Apart from localised
curvature of mylonite fabrics adjacent to the
central northern boundary of the Sandur schist
belt, the plutonic belts include neither radial patterns of magmatic-state lineations nor concentric
patterns of banding and magmatic-state planar
fabrics consistent with diapirism or ballooning.
4.4.1. Inflation of plutonic belts by vertical NWSE
dyking
Inflation by dyking is common in every plutonic
belt. For example, in the large quarries at
Sanganakallu, NE of Bellary, vertical NWSE
dykes of dark grey granite, 12 m thick, have
irregular boundaries and magmatic-state, NWSE

Fig. 9. Vertical NWSE dykes of dark facies of granite intruded


into hot host of pale granite, Sanganakallu, 4 km NE of Bellary.

104

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

Fig. 10. Part of a vertical NWSE sheet of granite breccia in


pale granite (scale: laundrywoman bottom right); the pale
blocks are similar to the host granite and the matrix is similar
to the dark dykes in Fig. 9. Sanganakallu, 4 km NE of Bellary.

dark facies show no distortion subsequent to their


magmatic disruption ( Fig. 12). In contrast, some
dykes were mylonitised as a result of strike-parallel
shearing whose effects were restricted to individual
dykes within swarms.
Granite dykes in the Dharwar batholith are
consistent with emplacement in fractures developed either during the magmatic consolidation of
the host (cf Clemens and Mawer, 1992) or postconsolidation brittle fracturing (cf Petford et al.,
1994). Widths of dykes in hot and cool fracture
systems in the Dharwar plutonic belts are consistent with the range quoted by Petford (1996) for
rapid magma ascent in dykes. Inflation of the
plutonic belts by granite dyking is also compatible
with the contention that magma pressure is an
integral part of the regional stress system analogous to the effective stress concept of soil
mechanics (Hutton, 1997), that is, NWSE dykes
in the Dharwar batholith are consistent with
emplacement into hot and cool plutonic host rocks
in a sinistral strikeslip regime when magma pressure was enhanced by high volatile content and
exceeded regional horizontal compressive stresses.
4.4.2. Magmatic banding and synmagmatic shear
zones
Magmatic banding in many belts trends NE
SW and has curving transitions into steep NW
SE high-strain zones (Fig. 4). This curvature is
common to S fabrics generated in ductile shear

Fig. 11. Pull-apart indicating sinistral shear contemporaneous


with emplacement of NS, dark grey granite dyke into pale
granite host, ca. 10 km S of Raichur.

Fig. 12. Vertical NWSE dyke of commingled, immiscible dark


facies (diorite?) in pale granite intruded into grey granite.
Devinagara, 40 km N of Bellary.

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

zones (Ramsay and Graham, 1970) and it implies


that the magmatic- and crystal-plastic fabrics were
part of a continuum of emplacement and deformation like that in more recent intrusions, for example, the Mono Creek Granite in the Cretaceous
Sierra Nevada batholith, California (Saint
Blanquet and Tikoff, 1997). Oblique relationships
between magmatic phenomena and regional shear
zones led Tikoff and Teyssier (1992) to propose
that tensional bridges between en echelon P shear
arrays facilitated emplacement in the Sierra
Nevada batholith. However, the oblique trends in
the Dharwar plutonic belts bear no relationship to
P shears. Instead they coincide with, or are nearly
parallel to, the plane of instantaneous shortening
in sinistral simple shear and transtension based on
theoretical analyses of extension, shortening and
Riedel shears in strikeslip regimes (Sanderson
and Marchini, 1984, Fig. 5).
The regular NESW trend of magmatic banding
indicates a lack of progressive rotation concomitant with increasing shear strain like that predicted
by the theoretical analyses of Ramsay and Graham
(1970) and Sanderson and Marchini (1984). This
lack of rotation suggests relatively rapid development of the banding prior to crystallinity exceeding
ca. 70% in accord with granites being relatively
short-lived geological events (Clemens and Mawer,
1992; Hanson and Glazner, 1995; Saint Blanquet
and Tikoff, 1997). However, deformation of NE
SW banding by common magmatic-state shear
zones and folds trending NWSE shows that crystallinity <ca. 70% persisted after banding had
formed.
Increase in width of the plutonic belts by additions of melt in NWSE dykes and the transtensional setting implied by the orientation of
NESW magmatic banding are consistent with a
protracted, punctuated continuum in a sinistral
strikeslip regime wherein pulses of melt with
pressures in excess of the regional horizontal maximum compression (after Hutton, 1997) led to
transtensional phases interspersed with transpression that gave rise to the high-strain zones of
mylonites and orthogneisses. The relatively regular
NWSE trend of magmatic banding in the plutonic
belt NW of the Sandur schist belt and in parts of
other plutonic belts ( Fig. 4) contrasts with the
common NESW trend, but vertical wedges of

105

granite with NWSE banding are consistent with


Huttons (1992) model of emplacement broadly
parallel to the trend of regional sinistral shear
planes with magma pressure exceeding regional
horizontal compressive stress (Hutton, 1997).
Syntectonic HT/LP metamorphism, fabrics, folds
and reverse faults in the schist belts show that
NESW shortening was intimately related to
growth of the plutonic belts during regional sinistral shear. The broadly contemporaneous NESW
shortening and NWSE sinistral shear contrasts
with the structural chronology in the west of the
craton where NESW shortening preceded NS
sinistral shearing. Intrusion of the ca. 2600 Ma
Chitradurga granite ( Table 1) into deformed metabasalts low in the Dharwar Supergoup (Seshadri
et al., 1981) suggests that NESW shortening
occurred earlier in the west, but the lack of isotopic
age data precludes detailed correlation of the structural chronology in each part of the craton.
5. Plate tectonic setting of the Dharwar batholith
Apart from the gravity-driven diapiric model
based on sagduction (Choukroune et al., 1995;
1997; Chardon et al., 1996, 1998), previous views
of the Late Archaean tectonic setting of the
Dharwar craton hinged on uniformitarian principles (Chadwick et al., 1997). At least two of the
earlier proposals likened the setting to that of the
North American Cordillera ( Krogstad et al., 1989)
and Andean margin and island arc magmatism
(Newton, 1990). Since 1989 we have consistently
drawn attention to the oblique-slip mobile regime
of the foreland region of the craton with analogies
in MesozoicCenozoic active continental margins
and micro-continental arcs (Chadwick et al., 1989,
1992). Here we integrate this oblique convergent
setting of the foreland with that of the Dharwar
batholith.
5.1. Plutonic belts and batholith accretion
5.1.1. Comparison with batholiths in Phanerozoic
convergent settings
Development of the Dharwar batholith as a
series of steep linear belts has close parallels with
batholith accretion in younger sites of oblique
convergence and mid-crustal strikeslip shearing

106

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

described, for example, by DLemos et al. (1992),


Hutton and Reavy (1992) and McCaffrey (1992).
With specific reference to Phanerozoic convergence
in the East Indies, for example, the structure of
the Dharwar batholith compares closely with the
sub-parallel, NWSE belts in which emplacement
of subduction-related, mainly I-type, Late Triassic
to Pliocene granites and granodiorites was focused
along deep faults acting as magma conduits
(McCourt et al., 1996). The deeply rooted, strike
slip, dextral faults that controlled magmatism in
the Sumatran plutonic belts were a consequence
of partitioning of displacements during oblique
convergence into arc-normal and arc-parallel components (Malod and Kemal, 1996; McCaffrey,
1996).
Similarities with the Dharwar batholith are also
found in the Andean mobile belt and North
American Cordillera where Mesozoic batholiths
are steep linear arrays of plutons (Pitcher, 1997).
JurassicCretaceous magmatism and volcanism in
northern Chile, for example, took place during
five distinct episodes within ca. 100 Ma when
extensional deformation in the foreland was
replaced by sinistral strikeslip displacement,
probably as a result of a change in the convergence
vector (Dallmeyer et al., 1996). However, although
Mesozoic to Early Tertiary Cordilleran plutonism
as a whole was associated with significant arc-

parallel transcurrent displacements (Condie and


Chomiak, 1996), large-scale relationships between
emplacement and tectonism appear to be more
complex in the Andean and Cordilleran systems
compared with the Dharwar batholith.
5.1.2. Significance of polymict conglomerates with
plutonic clasts in schist belts within the Dharwar
batholith
Polymict conglomerates with mixtures of plutonic and intrabasinal clasts are common, but
clasts of plutonic rocks are not universal. They are
intensely deformed, with steep schistosity and clast
elongation having moderate or steep plunge.
Zircons in a clast of gneiss from Kolar with ages
ca. 2800 Ma ( Krogstad et al., 1991) suggest the
erosional provenance included either early phases
of the batholith or late phases of the foreland
basement. In contrast, a SHRIMP UPb zircon
age of 2576+12 Ma from a deformed clast of
hornblende granodiorite in the Palkanmardi conglomerate in the Hutti schist belt compares closely
with the SHRIMP UPb zircon age of
2580+31 Ma from similar granodiorite in the plutonic belt extending south of the schist belt ( Fig. 4,
Table 1; A.P. Nutman personal communication,
1998). Granodiorite intrudes schistose metabasalts
in the south of the Hutti belt. The SHRIMP ages
indicate rapid unroofing and erosion of granodio-

Fig. 13. Approximate NESW section across the Dharwar craton showing relationships between plutonic belts and schist belts in
the batholith terrain in the northeast and SW-vergence of structures in the Dharwar Supergroup and its basement in the foreland
in the southwest.

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

rite after its emplacement as part of the growing


batholith. Deposition was closely followed by
deformation and further additions of granodiorite.
These relationships are identical to those
of intra-arc and fore-arc basins within and
adjacent to the calc-alkaline batholith in the
Palaeoproterozoic Ketilidian orogen in South
Greenland (Chadwick and Garde, 1996; Hamilton
et al., 1996). They reinforce our interpretation of
the schist belts in the Dharwar batholith as intraarc basins that were intimately involved with batholith accretion (Chadwick et al., 1996). The oblique
convergent setting of the Ketilidian orogen has other
aspects in common with the Dharwar craton,
namely, a batholith dominated by steep plutonic
belts of granite and granodiorite and a foreland
with an Archaean gneiss basement unconformably
overlain by Palaeoproterozoic volcanic and sedimentary basins which have much in common with
the Dharwar Supergroup. A salient difference is the
lack of a fore-arc in the Dharwar craton, but on
the grounds of composition and isotopic age data,
albeit limited (Dasgupta, 1995), we suggest that
psammitic and pelitic paragneisses (khondalites) in
the Proterozoic Eastern Ghats Mobile Belt east of
the craton are a likely candidate.
5.2. Late Archaean basin development and
deformation in the foreland
Basin development of the Dharwar Supergroup
and the pattern of ensuing Late Archaean deformation, which was the result of NESW shortening and NS sinistral transpression on steep,
brittle-ductile shear zones, compare closely with
basin development and deformation in the schist
belts and batholith in the east of the craton
(Fig. 13). Whereas sinistral shear outlasted NE
SW shortening in the west as indicated by refolded
large-scale folds [e.g. NW of Shimoga,
Mukhopadhyay (1986); Chadwick et al. (1991)],
there appears to have been a closer link between
NWSE sinistral shear and NESW shortening in
the east of the craton.
5.3. Late Archaean oblique convergence in the
Dharwar craton
Partitioning of convergence into two components, parallel and perpendicular to subduction, is

107

Fig. 14. Model of slip-partitioning in the Dharwar craton


during Late Archaean oblique convergence based on slip vectors
in the East Indies (Malod and Kemal, 1996). Note reduction
of arc-normal compression and increase in sinistral arc-parallel
shear from south to northwest. F, Foreland on the margin of
the overriding continental plate of the Dharwar convergent
system; DB, Dharwar batholith with juvenile plutonic suites in
the east and granites derived from partial melting of older continental crust adjacent to the foreland; the bulk of the subducting
plate is presumed to have been oceanic.

well known in Phanerozoic convergent boundary


systems (McCaffrey, 1996). McCaffrey showed,
for example, that slip-partitioning can occur on
relatively small spatial scales, and arc-parallel displacements are evident even where subduction is
normal to an arc. Application of slip-partitioning
of the type seen in the East Indies to the Dharwar
craton accounts for the NESW shortening in the
foreland and the intra-arc basins in the Dharwar
batholith as the result of arc-normal compression
during oblique convergence of an oceanic plate
subducting towards the WNW ( Fig. 14). Arcparallel displacements are manifest in the NWSE
and NS sinistral transcurrent shear zones in the
foreland and batholith.
6. Conclusions
The Late Archaean history of the Dharwar
craton is interpreted as the consequence of:
1. accretion of the Dharwar batholith, ca. 2750
2510 Ma, in the east of the craton with close

108

2.

3.

4.

5.

6.

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

structural and geochemical similarities to calcalkaline batholiths in Proterozoic and


Phanerozoic convergent settings;
control of batholith accretion by multipulse
emplacement of steep plutonic belts during
regional sinistral transcurrent displacements;
formation of schist belts in the east of the
craton as intra-arc basins during batholith
accretion;
overlap of arc accretion onto a foreland continental margin and its incipient back-arc basins
in the west;
WNW-directed convergence of oceanic lithosphere oblique to an overriding continental
lithospheric plate represented by the foreland
in the west; and
partitioning of the oblique convergence vector
into arc-parallel NWSE and NS sinistral displacements and arc-normal NESW shortening.

Acknowledgements
The authors are deeply indebted to the
Department of Mines and Geology, Government
of Karnataka, Bangalore, for logistic facilities and
steadfast support. BC gratefully acknowledges
funding from the Royal Society, London, the
Indian National Science Academy, New Delhi,
and the University of Exeter. They are also grateful
to A. P. Nutman for SHRIMP age data, and
thank A. A. Garde, J. Grocott and K. McCaffrey
for their constructive comments on an early version
of the paper and W.S. Pitcher for guidance through
the labyrinth of pluton and batholith terminology.
The authors also acknowledge an original idea of
M.A. Hamilton (Garde et al., 1998) for the construction of Table 1.

References
Allen, P., Condie, K.C., Bowling, G.P., 1986. Geochemical
characteristics and possible origins in the southern Closepet
batholith, South India. J. Geol. 94, 283299.
Anantha Iyer, G.V., Vasudev, V.N., 1979. Geochemistry of
Archaean metavolcanic rocks of Kolar and Hutti Gold
Fields, Karnataka, India. J. Geol. Soc. India 20, 419432.
Anantha Iyer, G.V., Vasudev, V.N., 1985. Copper metallogeny
in the Jogimardi volcanics, Chitradurga greenstone belt.
J. Geol. Soc. India 26, 580598.

Balakrishnan, S., Hanson, G.N., Rajamani, V., 1990. Pb and


Nd isotope constraints on the origin of high Mg and tholeiitic amphibolites, Kolar Schist Belt, South India. Contrib.
Miner. Petrol. 107, 279292.
Barbarin, B., Didier, J., 1991. Macroscopic features of mafic
microgranular enclaves. In: Didier, J., Barbarin, B. ( Eds.),
Enclaves and Granite Petrology. Elsevier, Amsterdam,
pp. 253262.
Bateman, P.C., 1992. Plutonism in the central part of the Sierra
Nevada batholith, California. U.S.G.S. Prof. Paper 1483,
186.
Beckinsale, R.D., Drury, S.A., Holt, R.W., 1980. 3,360-Myr
old gneisses from the South Indian Craton. Nature 283,
469470.
Bhaskar Rao, Y.J., Drury, S.A., 1982. Incompatible trace element geochemistry of Archaean metavolcanic rocks from
the Bababudan VolcanicSedimentary Belt, Karnataka.
J. Geol. Soc. India 23, 112.
Bhaskar Rao, Y.J., Naha, K., Srinivasan, R., Gopalan, K.,
1991. Geology, geochemistry and geochronology of the
Archaean Peninsular Gneiss around Gorur, Hassan District,
Karnataka, India. Ind. Acad. Sci. ( Earth and Planet. Sci.)
Proc. 100, 399412.
Bhaskar Rao, Y.J., Sivaram, T.V., Pantulu, G.V.C., Gopalan,
K., Naqvi, S.M., 1992. RbSr ages of Late Archaean metavolcanics and granites, Dharwar craton, South India, and
evidence for Early Proterozoic thermotectonic event(s).
Precambrian Res. 59, 145170.
Blake, S., Koyaguchi, T., 1991. Insights on magma mixing
model of volcanic rocks. In: Didier, J., Barbarin, B. ( Eds.),
Enclaves and Granite Petrology. Elsevier, Amsterdam,
pp. 403413.
Brown, M., 1994. The generation, segregation, ascent and
emplacement of granite magma: the migmatite-to-crustallyderived granite connection in thickened orogens. Earth-Sci.
Rev. 36, 83130.
Chadwick, B., Garde, A.A., 1996. Palaeoproterozoic oblique
plate convergence in South Greenland: a reappraisal of the
Ketilidian Orogen. Geol. Soc. Spec. Publ. 112, 179196.
Chadwick, B., Ramakrishnan, M., Vasudev, V.N., Viswanatha,
M.N., 1989. Facies distributions and structure of a Dharwar
volcanosedimentary basin: evidence for late Archaean transpression in southern India? J. Geol. Soc. Lond. 146,
825834.
Chadwick, B., Ramakrishnan, M., Viswanatha, M.N., 1985a.
Bababudan a Late Archaean intracratonic volcanosedimentary basin, Karnataka, South India. Part I: Stratigraphy
and Basin Development. J. Geol. Soc. India 26, 769801.
Chadwick, B., Ramakrishnan, M., Viswanatha, M.N., 1985b.
Bababudan a Late Archaean intracratonic volcanosedimentary basin, Karnataka, South India. Part II: Structure.
J. Geol. Soc. India 26, 802821.
Chadwick, B., Vasudev, V.N., Jayaram, S., 1988. Stratigraphy
and structure of Late Archaean, Dharwar volcanic and sedimentary rocks and their basement in a part of the Shimoga
Basin east of Bhadravathi, Karnataka. J. Geol. Soc., India
32, 119.
Chadwick, B., Vasudev, V.N., Krishna Rao, B., Hegde, G.V.,

B. Chadwick et al. / Precambrian Research 99 (2000) 91111


1991. The Stratigraphy and Structure of the Dharwar Supergroup adjacent to the Honnali Dome: Implications for Late
Archaean Basin Development and Regional Structure in the
Western Part of Karnataka. J. Geol. Soc. India 38, 457484.
Chadwick, B., Vasudev, V.N., Krishna Rao, B., Hegde, G.V.,
1992. The Dharwar Supergroup: basin development and
implications for Late Archaean tectonic setting in western
Karnataka, Southern India. Univ. Western Australia Publ.
22, 315.
Chadwick, B., Vasudev, V.N., Ahmed, N., 1996. The Sandur
Schist Belt and its Adjacent Plutonic Rocks: Implications
for Late Archaean Crustal Evolution in Karnataka. J. Geol.
Soc. India 47, 3757.
Chadwick, B., Vasudev, V.N., Hegde, G.V., 1997. The Dharwar
craton, southern India, and its Late Archaean plate tectonic
setting: current interpretations and controversies. Ind. Acad.
Sci. ( Earth and Planet. Sci.) Proc. 106 (4), 110.
Chardon, D., Choukroune, P., Jayananda, M., 1996. Strain
patterns, decollement and incipient sagducted greenstone
terrains in the Archaean Dharwar craton (south India).
J. Struct. Geol. 18, 9911004.
Chardon, D., Choukroune, P., Jayananda, M., 1998. Sinking
of the Dharwar Basin (South India): implications for
Archaean tectonics. Precambrian Res. 91, 1539.
Choukroune, P., Bouhallier, H., Arndt, N.T., 1995. Soft lithosphere during periods of Archaean crustal growth or crustal
reworking. Geol. Soc. Spec. Publ. 95, 6786.
Choukroune, P., Ludden, J.N., Chardon, D., Calvert, A.J.,
Bouhallier, H., 1997. Archaean crustal growth and tectonic
processes: a comparison of the Superior Province, Canada,
and the Dharwar craton, India. Geol. Soc. Spec. Publ.
121, 6398.
Clemens, J.D., Mawer, C.K., 1992. Granitic magma transport
by fracture propagation. Tectonophysics 204, 339360.
Condie, K.C., Chomiak, B., 1996. Continental accretion: contrasting Mesozoic and Early Proterozoic tectonic regimes in
North America. Tectonophysics 265, 101126.
Condie, K.C., Sloan, R.E., 1998. Origin and Evolution of Earth.
Prentice Hall, USA.
DLemos, R.S., Brown, M., Strachan, R.A., 1992. Granite generation, ascent and emplacement within a transpressional
orogen. J. Geol. Soc. Lond. 149, 487490.
Dallmeyer, R.D., Brown, M., Grocott, J., Taylor, G.K.,
Treloar, P.J., 1996. Mesozoic magmatic and tectonic
events within the Andean Plate boundary zone, 262730S,
North Chile: Constraints from 40Ar/39Ar mineral ages.
J. Geol. 104, 1940.
Dasgupta, S., 1995. Pressuretemperature evolutionary history
of the eastern Ghats Granulite province: recent advances
and some thoughts. Geol. Soc. India Mem. 34, 101110.
Devaraju, T.C., Rajshekar, N., Srikantappa, C., Khandali,
S.D., Subba Rao, G., 1990. Lithium pegmatites of Amareshwar, Raichur District, Karnataka, India. In: Naqvi, S.M.
( Ed.), Precambrian Continental Crust and its Economic
Resources. Elsevier, Amsterdam, pp. 653669.
Drury, S.A., 1983. The petrogenesis and setting of Archaean
metavolcanics from Karnataka State, South India.
Geochim. Cosmochim. Acta 47, 317329.

109

Fernandez, A., Barbarin, B., 1991. Relative rheology of coeval


mafic and felsic magmas: nature of resulting interaction process and shape and mineral fabrics of mafic microgranular
enclaves. In: Didier, J., Barbarin, B. ( Eds.), Enclaves and
Granite Petrology. Elsevier, Amsterdam, pp. 243275.
Fernandez, A., Castro, A., de la Rosa, J.D., Moreno-Ventas,
I., 1997. Rheological aspects of magma transport inferred
from rock structures. In: Bouchez, J.L., Hutton, D.G.W.,
Stephens, W.E. (Eds.), Granite: From Segregation of Melt
to Emplacement Fabrics. Kluwer Academic, Dordrecht,
pp. 7591.
Fernandez, A.N., Gasquet, D.R., 1994. Relative rheological
evolution of chemically contrasted coeval magmas: example
of the Tichka plutonic complex (Morocco). Contrib. Miner.
Petrol. 116, 316326.
Friend, C.R.L., 1983. The link between granite production and
the formation of charnockites: evidence from Kabbaldurga,
Karnataka. In: Atherton, M.P., Gribble, C.D. ( Eds.),
Migmatities, Melting and Metamorphism. Shiva Press,
Nantwich, pp. 264276.
Friend, C.R.L., Nutman, A.P., 1991. SHRIMP UPb Geochronology of the Closepet Granite and Peninsular Gneiss,
Karnataka, South India. J. Geol. Soc. India 38, 357368.
Garde, A.A., Chadwick, B., Grocott, J., Hamilton, M.A.,
McCaffrey, K., Swager, C., 1998. An overview of the Paleoproterozoic Ketilidan Orogen, South Greenland. Eastern
Canadian Shield OnshoreOffshore Transect ( ECSOOT ),
Wardle, R.J., Hall, J. ( Eds.), Univ. British Columbia, Lithoprobe Rep. 68, 5066.
Gibbs, A.D., 1987. Development of extension and mixed-mode
sedimentary basins. Geol. Soc. Spec. Publ. 28, 1933.
Goodwin, A.M., Smith, I.E.M., 1980. Chemical discontinuities
in Archean metavolcanic terrains and the development of
Archean crust. Precambrian Res. 10, 301311.
Hamilton, M.A., Garde, A.A., Chadwick, B., Swager, C., 1996.
Observations on Palaeproterozoic fore-arc sedimentation
and deformation: preliminary UPb results from the Ketilidian Orogen, South Greenland, Eastern Canadian Shield
OnshoreOffshore Transect ( ECSOOT ), Wardle, R.J., Hall,
J. ( Eds.), Univ. British Columbia, Lithoprobe Rep. 57,
112122.
Hamilton, W.B., 1998. Archean magmatism and deformation
were not products of plate tectonics. Precambrian Res. 91,
143179.
Hansen, E.C., Newton, R.C., Janardhan, A.S., Lindenberg, S.,
1995. Differentiation of Late Archean Crust in the Eastern
Dharwar Craton, KrishnagiriSalem Area, South India.
J. Geol. 103, 629651.
Hansen, E.C., Stern, R.J., Devaraju, T.C., Mahabaleshwar, B.,
Kinny, P.J., 1997. RubidiumStrontium Whole-Rock Ages
of Banded and Incipient Charnockites from Southern
Karnataka. J. Geol. Soc. India 50, 267275.
Hanson, R.B., Glazner, A.F., 1995. Thermal requirements for
extensional emplacement of granitoids. Geology 23,
213216.
Hanson, G.N., Krogstad, E.J., Rajamani, V., 1988. Tectonic
setting of the Kolar Schist Belt, Karnataka, India. J. Geol.
Soc. India 31, 4042.

110

B. Chadwick et al. / Precambrian Research 99 (2000) 91111

Hanuma Prasad, M., Krishna Rao, B., Vasudev, V.N., Basavalingu, B., 1996. PT Conditions of Metamorphism of
Supracrustal Rocks in the Sandur Schist Belt, Dharwar
Craton, Southern India. J. Geol. Soc. India 48, 617628.
Hanuma Prasad, M., Krishna Rao, B., Vasudev, V.N., Srinivasan, R., Balaram, V., 1997. Geochemistry of Archaean
Bimodal Volcanic Rocks of the Sandur Supracrustal Belt,
Dharwar Craton, Southern India. J. Geol. Soc. India 49,
307322.
Harris, N.B.W., Jayaram, S., 1982. Metamorphism of cordierite
gneisses from the Bangalore region of the Indian Archean.
Lithos 15, 8998.
Hibbard, M.J., 1995. Petrography to Petrogenesis. PrenticeHall, New Jersey, USA.
Hutton, D.H.W., 1988. Granite emplacement mechanism and
tectonic controls: inferences from deformation studies.
Trans. R. Soc. Edin.: Earth Sci. 79, 245255.
Hutton, D.H.W., 1992. Granite sheeted complexes: evidence for
the dyking ascent mechanism. Trans. R. Soc. Edin.: Earth
Sci. 83, 377382.
Hutton, D.H.W., 1996. The space problem in the emplacement
of granite. Episodes 19, 114119.
Hutton, D.H.W., 1997. Syntectonic granites and the principle
of effective stress: a general solution to the space problem?
In: Bouchez, J.L., Hutton, D.G.W., Stephens, W.E. ( Eds.),
Granite: From Segregation of Melt to Emplacement Fabrics. Kluwer Academic, Dordrecht, pp. 189197.
Hutton, D.H.W., Reavy, R.J., 1992. Strikeslip tectonics and
granite petrogenesis. Tectonics 11, 960967.
Ingram, G., Hutton, D.H.W., 1997. The Great Tonalite sill:
emplacement into a contractional shear zone and implications for late Cretaceous to early Eocene tectonics in southeastern Alaska and British Columbia. Geol. Soc. Am. Bull.
106, 715728.
Jayananda, M., Mahabaleshwar, B., 1991. Relationship
between shear zones and igneous activity: the Closepet granite of southern India. Ind. Acad. Sci. ( Earth and Planet.
Sci.) Proc. 100, 3136.
Jayananda, M., Martin, H., Peucat, J-J., Mahabaleshwar, B.,
1995. Late Archaean crust-mantle interactions: geochemistry of LREE-enriched mantle derived magmas. Example of
the Closepet batholith, South India. Contrib. Miner. Petrol.
119, 314329.
Kaila, K.L., Sain, K., 1997. Variation of Crustal Velocity Structure in India as Determined from DSS Studies and Their
Implications on Regional Tectonics. J. Geol. Soc. India
49, 395407.
Kaila, K.L., Roy Choudhury, K., Reddy, P.R., Krishna, V.G.,
Narain, H., Subbotin, S.I., Sollugub, V.B., Chekunov, A.V.,
Kharetchiko, G.E., Lazarenko, M.A., Ilchenko, T.V., 1979.
Crustal structure along KavaliUdipi profile in the Indian
Peninsular shield from deep seismic soundings. J. Geol. Soc.
India 20, 307333.
Krogstad, E.J., Balakrishnan, S., Mukhopadhyay, D.K., Rajamani, V., Hanson, G.N., 1989. Plate tectonics 2.5 Billion
Years ago: evidence at Kolar, South India. Science 243,
13371340.

Krogstad, E.J., Hanson, G.N., Rajamani, V., 1991. UPb ages


of zircon and sphene for two gneiss terranes adjacent to the
Kolar Schist Belt, South India: evidence for separate crustal
evolution histories. J. Geol. 99, 801816.
Krogstad, E.J., Hanson, G.N., Rajamani, V., 1995. Sources of
continental magmatism adjacent to the late Archean Kolar
Suture Zone, south India: distinct isotopic and elemental
signatures of two late Archaean magmatic series. Contrib.
Miner. Petrol. 122, 159173.
Kumar, A., Bhaskar Rao, Y.J., Sivaraman, T.V., Gopalan, K.,
1996. SmNd ages of Archaean metavolcanics of the
Dharwar craton, South India. Precambrian Res. 80,
206215.
Mahadevan, T.M., 1994. Deep continental structure of India:
a review. Geol. Soc. India, Mem. 28.
Malod, J.A., Kemal, B.M., 1996. The Sumatra margin: oblique
subduction and lateral displacement of the accretionary
prism. Geol. Soc. Spec. Publ. 106, 1928.
Martin, H., 1994. The Archean grey gneisses and the genesis of
continental crust. In: Condie, C. ( Ed.), Archaean Crustal
Evolution. Elsevier, Amsterdam, pp. 205259.
McCaffrey, K.J.W., 1992. Igneous emplacement in the transpressive shear zone: Ox Mountains Igneous complex.
J. Geol. Soc. Lond. 149, 221235.
McCaffrey, R., 1996. Slip partitioning at convergent plate
boundaries of SE Asia. Geol. Soc. Spec. Publ. 106, 318.
McCourt, W.J., Crow, M.J., Cobbing, E.J., Amin, T.C., 1996.
Mesozoic and Cenozoic plutonic evolution of SE Asia: evidence from Sumatra, Indonesia. Geol. Soc. Spec. Publ.
106, 321335.
Monrad, J.R., 1983. Evolution of sialic terranes in the vicinity
of the Holenarsipur belt, Hassan District, Karnataka, India.
Geol. Soc. Ind. Mem. 4, 343364.
Moyen, J.F., Martin, H., Jayananda, M., 1997. Origine du granite fini-Archeen de Closepet (Inde du Sud): appropos de la
modelisation geochimique du comportement des element en
trace. C. R. Acad. Sci. Paris 325, 659664.
Mukhopadhyay, D., 1986. Structural pattern in the Dharwar
craton. J. Geol. 94, 167186.
Mukhopadhyay, D., Matin, A., 1993. The structural anatomy
of the Sandur schist belt a greenstone belt in the Dharwar
craton of South India. J. Struct. Geol. 15, 309322.
Mukhopadhyay, D.K., 1989. Significance of small-scale structures in the Kolar Schist Belt, south India. J. Geol. Soc.
India 33, 291308.
Naha, K., Mukhopadhyay, D., Dastidar, S., Mukhopadhyay,
R.P., 1995. Basement-cover relations between a granite
gneiss body and its metasedimentary envelope: a structural
study from the Early Precambrian Dharwar tectonic province, southern India. Precambrian Res. 72, 283299.
Newton, R.C., 1990. The Late Archaean High-grade Terrain of
South India and the deep structure of the Dharwar Craton.
In: Salisbury, M.H., Fountain, D.M. (Eds.), Exposed CrossSections of the Continental Crust. Kluwer Academic,
Dordrecht, pp. 305326.
Nutman, A.P., Chadwick, B., Ramakrishnan, M., Viswanatha,
M.N., 1992. SHRIMP UPb ages of detrital zircon in Sargur

B. Chadwick et al. / Precambrian Research 99 (2000) 91111


supracrustal rocks in Western Karnataka, southern India.
J. Geol. Soc. India 39, 367374.
Nutman, A.P., Chadwick, B., Krishna Rao, B., Vasudev, V.N.,
1996. SHRIMP U/Pb zircon ages of acid volcanic rocks in
the Chitradurga and Sandur Groups, and granites adjacent
to the Sandur Schist Belt, Karnataka. J. Geol. Soc. India
47, 153164.
Oak, K.A., 1990. The geology and geochemistry of the Closepet
granite, Karnataka, South India. Unpublished PhD Thesis.
Council for National Academic Awards, Oxford Polytechnic, UK.
Paterson, S.R., Vernon, R.H., Tobisch, O.T., 1989. A review of
criteria for the identification of magmatic and tectonic foliations in granitoids. J. Struct. Geol. 11, 349363.
Paterson, S.R., Fowler, T.K., Miller, R.B., 1996. Pluton
emplacement in arcs: a crustal-scale exchange process.
Trans. R. Soc. Edin.: Earth Sci. 87, 115123.
Petford, N., 1996. Dykes or diapirs? Trans. R. Soc. Edin.: Earth
Sci. 87, 105114.
Petford, N., Lister, J.R., Kerr, R.C., 1994. The ascent of felsic
magmas in dykes. Lithos 32, 161168.
Peucat, J.J., Mahabaleshwar, B., Jayananda, M., 1993. Age of
younger tonalitic magmatism and granulitic metamorphism
in the South India transition zone ( Krishnagiri area):
comparison with older Peninsular gneisses from the Gorur
Hassan area. J. Metam. Geol. 11, 879888.
Peucat, J.J., Bouhallier, H., Fanning, C.M., Jayananda, M.,
1995. Age of the Holenarsipur Greenstone belt, relationships
with the surrounding Gneisses ( Karnataka, South India).
J. Geol. 103, 701710.
Pitcher, W.S., 1979. The nature, ascent and emplacement of
granitic magmas. J. Geol. Soc. Lond. 136, 627662.
Pitcher, W.S., 1997. The Nature and Origin of Granite. 2nd ed.
Chapman and Hall, London.
Radhakrishna, B.P., 1956. The Closepet granites of Mysore
State, India. Mysore Geological Association, Bangalore.
Ramakrishnan, M., Venkata Dasu, S.P., Kroner, A., 1994.
Middle Archaean age of Sargur Group by single grain zircon
dating and geochemical evidence for the Clastic Origin of
Metaquartzite from J. C. Pura Greenstone belt, Karnataka.
J. Geol. Soc. India 44, 605616.
Ramsay, J.G., Graham, R.H., 1970. Strain variation in shear
belts. Can. J. Earth Sci. 7, 786813.
Rogers, J.J.W., 1988. The Arsikere granite of southern India:
magmatism and metamorphism in a previously depleted
crust. Chem. Geol. 67, 155163.
Roy, A., 1979. Polyphase folding deformation in the Hutti
Maski Schist Belt, Karnataka. J. Geol. Soc. India 20,
598607.
Roy, A., Biswas, S.K., 1979. Metamorphic history of the
Sandur schist belt, Karnataka. J. Geol. Soc. India 20,
179187.
Roy, A., Biswas, S.K., 1983. Stratigraphy and structure of the
Sandur schist belt, Karnataka. J. Geol. Soc. India 24, 1929.
Russell, J., Chadwick, B., Krishna Rao, B., Vasudev, V.N.,
1996. Whole-rock Pb/Pb isotopic ages of Late Archaean

111

limestones, Karnataka, India. Precambrian Res. 78,


261272.
de Saint Blanquet, M., Tikoff, B., 1997. Development of magmatic to solid-state fabrics during syntectonic emplacement
of the Mono Creek Granite, Sierra Nevada Batholith. In:
Bouchez, J.L., Hutton, D.G.W., Stephens, W.E. ( Eds.),
Granite: from Segregation of Melt to Emplacement Fabrics.
Kluwer Academic, Dordrecht, pp. 189197.
Sanderson, D.J., Marchini, W.R.D., 1984. Transpression.
J. Struct. Geol. 6, 449458.
Seshadri, T.S., Chauduri, A., Harinadha Babu, P., Chayapathi,
N., 1981. Chitradurga belt. Geol. Surv. India Mem. 112,
163198.
Smeeth, W.F., 1915. Geological Map of Mysore. Department
Mines and Geology, Mysore.
Stroh, P.T., Monrad, J.R., Fullagar, P.D., Naqvi, S.M., Hussain, S.M., Rogers, J.J.W., 1983. 3000 m.y.-Old Halekote
Trondhjemite: a record of stabilization of the Dhwarar
Craton. Geol. Soc. Ind. Mem. 4, 365376.
Subba Rao, M.V., Rama Rao, P., Divakara Rao, V., 1998. The
advent of the Proterozoic a trigger for extensive intracrustal processes in the South Indian Shield? Gondwana Res.
1, 275283.
Swami Nath, J., Ramakrishnan, M., Viswanatha, M.N., 1976.
Dharwar stratigraphic model and Karnataka craton evolution. Geol. Surv. India Records 107, 149175.
Sylvester, P.J., 1994. Archean granite plutons. In: Condie, C.
( Ed.), Archaean Crustal Evolution. Elsevier, Amsterdam,
pp. 261314.
Taylor, P.N., Chadwick, B., Moorbath, S., Ramakrishnan, M.,
Viswanatha, M.N., 1984. Petrography, chemistry and isotopic ages of Peninsular Gneiss, Dharwar acid volcanic
rocks and the Chitradurga granite with special reference
to the late Archaean evolution of the Karnataka craton.
Precambrian Res. 23, 349375.
Tikoff, B., Teyssier, C., 1992. Crustal-scale en echelon Pshear tensional bridges: a possible solution to the batholithic room problem. Geology 20, 927930.
Trendall, A.F., de Laeter, J.R., Nelson, D.R., Bhaskar Rao,
Y.J., 1997a. Further zircon UPb age data for the Daginkatte Formation, Dharwar Supergroup, Karnataka Craton.
J. Geol. Soc. India 50, 2530.
Trendall, A.F., de Laeter, J.R., Nelson, D.R., Mukhopadhyay,
D., 1997b. A precise zircon UPb age for the base of the
BIF of the Mulaingiri Formation (Bababudan Group,
Dharwar Supergroup) of the Karnataka Craton. J. Geol.
Soc. India 50, 161170.
Windley, B.F., 1995. The Evolving Continents. 3rd ed., Wiley.
de Wit, M.J., 1998. On Archean granites, greenstones, cratons
and tectonics: does the evidence demand a verdict? Precambrian Res. 91, 181226.
Zacharaiah, J.K., Hanson, G.N., Rajamani, V., 1995. Postcrystallization disturbance in the neodymium and lead isotope
systems of metabasalts from the Ramagiri schist belt, southern India. Geochim. Cosmochim. Acta 59, 31893203.
Zacharaiah, J.K., Mohanta, M.K., Rajamani, V., 1996. Accretionary evolution of the Ramagiri schist belt, Eastern
Dharwar Craton. J. Geol. Soc. India 47, 279291.

You might also like