Geological Society of America Special Paper 380 2004
Gneiss domes and crustal flow Donna L. Whitney* Christian Teyssier Department of Geology & Geophysics, University of Minnesota, Minneapolis, Minnesota 55455, USA Olivier Vanderhaeghe Université Henri Poincaré Nancy 1, Géologie et Gestion des Ressources Minérales et Energétiques, 54506 Vandoeuvre-lès-Nancy Cedex, France ABSTRACT Gneiss domes are ubiquitous structures in all exhumed orogens, and their formation represents a first order thermal-tectonic process that has operated from the Archean to the present. The vertical flow of crust to create domal structures is a significant factor in the redistribution of heat and material in orogens and therefore in the evolution of continents. Worldwide, gneiss domes display many similarities in geometry (aspect ratio), petrology, and structure, and these similarities transcend differences in tectonic setting. Gneiss domes are cored by high-grade metamorphic rocks (including migmatite) ± granitoids, and the core rocks commonly record a component of isothermal decompression, in contrast to mantling schists, and may exhibit a late, low-pressure–high-temperature metamorphic assemblage. Rapid cooling typically follows isothermal decompression, as hot rocks are rapidly emplaced at higher structural levels. Most gneiss domes are elongate parallel to the strike of the orogen. Domes with long dimension ≤90 km have a ratio of long to short axes of ~2:1–3:1. The elliptical shape of gneiss domes worldwide suggests that their morphology, and therefore genesis, is controlled by crustal flow dynamics, including the magnitude of vertical versus lateral crustal flow. The conditions and mechanisms involved in dome formation inform the relative rates of vertical and lateral crustal flow during orogeny. Keywords: [[pending from author; pending from author; pending from author; pending form author.]] SIGNIFICANCE OF GNEISS DOMES IN OROGENY
(20–30 km) layer of partially molten crust (Nelson et al., 1996) that may flow laterally to build an orogenic plateau (Royden, 1996) and drive or accommodate gravitational collapse (Rey et al., 2001). Regional-scale migmatite complexes are exhumed manifestations of these phenomena involving the flow of partially molten crust and are commonly characterized by domal structures. The link between gneiss domes and fundamental orogenic processes such as melting, flow, and exhumation requires further examination. Despite the spatial association of gneiss domes and migmatites/granite, a major question is whether metamorphism, melting, and doming are coeval, or whether doming is instead a
Gneiss domes are structural domes cored primarily by gneissic rocks (migmatite, orthogneiss) and granite and mantled by high-grade schist and gneiss. These domes occur in every exhumed orogen (Fig. 1; Appendix). Their ubiquity indicates that they form by the thermal-mechanical processes that characterized continental tectonics from the Archean to the Cenozoic. Because of the close link between gneiss domes and migmatites/magma, the origin and significance of gneiss domes directly relates to some of the most important recent advances in understanding orogeny. Orogens are characterized by a thick *
[email protected]
Whitney, D.L, Teyssier, C., and Vanderhaeghe, O., 2004, Gneiss domes and crustal flow, in Whitney, D.L, Teyssier, C., and Siddoway, C.S., Gneiss domes in orogeny: Boulder, Colorado, Geological Society of America Special Paper 380, p. xxx–xxx. For permission to copy, contact
[email protected]. © 2004 Geological Society of America
1
2
D.L. Whitney, C. Teyssier, and O. Vanderhaeghe
��������������������������� �������������� �����������
�� �� �� �� �� �� �� �� �� ��������� �� � �� �� �� �� �� ���������� �� ����������� �� �� ��������������� �� �� �� ���������� �� �� �������������� �� �� �� �� �� �� �� �������� �� �������� �� �� �� �� �� ������������ ������� �������������� �� �� �� �������� �� � �� �
��
��
��
�� �������������
�� ��
��
� ��
��
�� ��������� �� �� ����� �� �������������
�� ��
�
�� ������� �������
�
��
�
�� ��������� �� �����������
������ ����������������
Figure 1. World map showing locations of gneiss domes (black dots) that have been documented in the literature. In belts with numerous closely spaced gneiss domes, the black dots are representative of the region. The number of dots does not correspond to the number of domes; there are many more gneiss domes than are shown on this map, particularly in Precambrian terranes. Numbers on map are keyed to Table A1 (Appendix), which contains supplementary references to the literature on gneiss domes.
subsolidus syn- to post-metamorphic process. Some domes are considered to have formed late relative to high-grade metamorphism and melting and are viewed as folding phenomena, such as the result of fold interference or folded nappes (Ramsay, 1967; Thompson et al., 1968; Talbot, 1974; van Staal and Williams, 1983; Holdsworth, 1989; Brown et al., 1991; Steck et al., 1998; Fowler and Osman, 2001). However, if doming and melting (or the presence of melt) are coeval, the contribution of gravitational upwelling (diapirism) in gneiss dome formation should be considered. Diapirism was long favored as a model for gneiss dome genesis (e.g., Eskola, 1949; Ramberg, 1980; Brun, 1980; Coward, 1981; Soula, 1982; Faure et al., 1986). For much of the 1980–1990s, however, contraction (folding) and extension (core complex development) were more common interpretations, although it is possible that the processes of buoyant flow (diapirism) and crustal buckling or core complex formation are linked. The large number of studies of metamorphic core complexes in the past 20 years has obscured the difference between domes and core complexes. Some metamorphic core complexes are domes (Bitterroot complex, Montana, United States; Naxos, Greece) or contain gneiss domes within them (Shuswap com-
plex, British Columbia, Canada; the Albion-Raft River-Grouse Creek complex, Idaho and Utah, United States). Studies of these and other metamorphic core complexes commonly focus on the mechanisms of detachment faulting and exhumation of middle crustal rocks rather than the occurrence and formation of domal structures within the core complex (Davis and Lister, 1988; Saltzer and Hodges, 1988; Chen et al., 1990). However, the P-T-t history of gneiss domes in core complexes cannot be completely described by low angle normal faulting (Fayon et al., this volume), so metamorphic core complexes that are or contain domes require the consideration of additional thermal-tectonic processes to explain the origin of the domes. Of particular interest is the question of whether diapirism has a primary (driving force) or secondary (passive) role in relation to extension in core complex development. Domes have their own dynamics and may form during contraction or extension. More generally, orogenic crust may flow in a channel independent of boundary conditions (e.g., Rey et al., 2001; Vanderhaeghe and Teyssier, 2001), and domes represent the vertical component of this flow. Therefore, what is most fundamental about domes is what they contribute to the transfer
Gneiss Domes and Crustal Flow of mass and heat during orogeny, their role in the P-T-t evolution of the orogen, and the long-term differentiation of continental crust. In this paper, we examine gneiss domes to evaluate general features that can be used to understand dome formation in particular and fundamental thermal-mechanical processes of continental tectonics in general. This review is not intended to be a thorough summary of any orogen or tectonic regime. We aim to highlight the importance of the domal structure and P-T-t history of gneiss domes for understanding their origin and significance in the context of orogeny and to emphasize the importance of gneiss dome formation as a mechanism for material and heat transfer in the evolution of continental crust. A GNEISS DOME IS MORE THAN A STRUCTURE Gneiss domes have been called mantled gneiss domes, gneissic domes, granite-gneiss domes, metamorphic domes, and migmatite domes. We strongly suspect that crustal melting is a fundamental process in the origin of most gneiss domes (Teyssier and Whitney, 2002), and therefore most domes are in fact anatectic migmatite domes. For this paper, however, we use the simple term gneiss dome, recognizing that the core of some gneiss domes—in particular, Precambrian domes—are dominated by granitic plutons. The term gneiss dome can be applied to any structure that contains high-grade metamorphic rocks or plutonic rocks displaying a domal foliation; the term implies no genetic of temporal relationship between metamorphism and doming, and can be used to describe domal structures dominated by granitic rocks or metamorphic rocks other than gneiss (e.g., schist, marble). This paper focuses on anatectic migmatite domes.
3
Structural Geology Gneiss domes are defined by foliation or gneissic layering dipping radially away from the core of the dome (Figs. 2 and 3A). At deep levels, foliation trajectories may exhibit a funnel shape; however, we use the term “dome” because it is the shape typically encountered at exposure level. This domal structure is a fundamental feature that must be explained in order to understand the origin and significance of gneiss domes. Gneiss domes are typically characterized by a core and an overlying metamorphic sequence (mantling gneiss or schist) with the primary distinction between the two being the amount of former melt in the core. Depending on the level of exposure and other processes related to denudation and weathering, domes may consist of the metamorphic mantle only or of both the core and metamorphic mantle. In cases where both core and mantling rocks are exposed, the transition between the two may be gradual (e.g., defined by increasing amount of leucosome in migmatites into the core, such as a metatexite-diatexite
���������� ����� ����������
���������� ����
�������� ����������� �����
������������
Figure 2. Schematic view of a gneiss dome showing some characteristic features: migmatitic core, metamorphic mantling rocks, and cascading (spruce tree) folds. Arrows indicate sense of shear that might be expected for diapiric ascent of the dome.
Figure 3. A. The contact between the migmatitic core and the marble-dominated metamorphic cover of the Naxos dome. B. Outcrop photograph of migmatite in the core of the Naxos dome showing that the granitic fraction of the rocks is greater than the metasedimentary component, which typically occurs as thin layers and wispy lenses surrounded by foliated granite.
4
D.L. Whitney, C. Teyssier, and O. Vanderhaeghe
transition) or abrupt (e.g., migmatite/granite structurally overlain by non-migmatitic rocks; Fig. 3A). To date, clues to the origin of gneiss domes have typically been looked for in the rocks that mantle dome cores. These features include cascading (“spruce tree”) folds (Skehan, 1961; Miller, 1980) (Fig. 2), patterns of foliation-lineation, and specific patterns of finite strain expected from various dome-forming mechanisms, maturity of the doming, and levels of exposure (i.e., structural level of dome). In dome cores dominated by diatexite or granitic rocks, foliation at outcrop scale is variable owing to folding and flow of partially molten rocks (Fig. 3B); yet, at the scale of the map, foliation is rather well organized. Foliation trajectories may outline a second-order structure in the form of subdomes—smaller scale domes within the larger domal structure (Fig. 4). Examples of gneiss domes containing subdomes include Thor-Odin (British Columbia, Canada; Vanderhaeghe et al. [1999]), Naxos (Cycladic Islands, Greece; Vanderhaeghe, this volume), Menderes (western Turkey), Velay (Massif Central, France; Lagarde et al. [1994]), Entia (central Australia; Arnold et al. [1995]), and many Precambrian domes. Lineation may be difficult to discern in the field owing to high-temperature recrystallization. Because the core regions of domes contain a large proportion of migmatite or granite in which (linear) fabrics and strain history are not easily identified, new methods must be developed to recover fabric orientation, strain intensity, and shape of the finite strain ellipsoid. The method based on the anisotropy of magnetic susceptibility (AMS), developed to study the fabric of granite (Bouchez, 1997), has also proved valuable in the analysis of migmatites (Ferré et al., 2003); it is a promising new way to address regional strain patterns in gneiss domes.
���������� �����������
�������������� ��������
����� ����
����� ����
Other than the domal shape of foliation or layering that is common to and defines all domes, structures are highly variable. Lineation patterns, where documented, are not necessarily radial, and finite strain patterns do not show clear trends. For example, domes characterized by flattening strain at their apex and plane or constriction strain on their limbs are rare, except perhaps in some Precambrian domes (Brun et al., 1981; Collins et al., 1998). This structural complexity may reflect the fact that several mechanisms are responsible for doming or that finite strain in domes reflects not only the internal dynamics of domes but also their interaction with crustal flow on a time scale similar to that of dome formation. If a dome is emplaced diapirically, specific structures and strain/kinematic patterns are expected. For example, experiments by Dixon (1975) and Cruden (1988, 1990) demonstrated the distribution of fabrics and strain around diapirs. Figure 2 shows that foliation forms a complex pattern around a diapir. Early-formed foliation is folded into antiforms and synforms with axes wrapping around the dome. Strain intensity in a diapir shows great variations, from regions that essentially escape deformation to zones that are intensely deformed as the material is squeezed by the ascent and ballooning of the diapir. A diapir also shows a large variation in kinematic vorticity, from zones characterized by dominant simple shear on the flanks of the diapir to zones that are deformed mainly by coaxial strain. Perhaps the most compelling signature of diapiric flow is the variation of the shape of the strain ellipsoid in space and time. With physical models of a rising diapir, Cruden (1988) demonstrated how the strain ellipsoid changes from flattening in the material above the diapir toward plane strain as the material is sheared on the flanks of the diapir and finally to constriction as the material is dragged toward the tail (overturned portion) of the diapir.
� ������
����� ����
������������ �������� ������������
����� ����
���������
���������
��������������������
�
����� �����
����� ����
���������� ������� ������������ �������������
������������� �������������
�
�
Figure 4. Domes and subdomes; maps shown at approximately same scale. A. The Entia dome in central Australia. Map modified from satellite photograph and mapping by M. Hand (Arnold et al., 1995). B. The Thor-Odin dome, Shuswap metamorphic core complex, British Columbia, Canada (map modified from Vanderhaeghe et al., 1999). C. Naxos dome, Aegean Sea, Greece (map modified from Vanderhaeghe, this volume).
Gneiss Domes and Crustal Flow Petrology Most gneiss domes are characterized by a core of upper amphibolite to granulite facies gneiss, typically including anatectic migmatites and orthogneiss/granite. Many are structurally overlain (mantled) by high-grade metasedimentary or metavolcanic rocks (Eskola, 1949). In cases where the mantling rocks are migmatitic, the volume percent melt inferred from anatectic migmatites and granites is higher in the core of the dome than in the cover rocks. Many gneiss dome cores record higher pressures than their mantling rocks and may record the highest pressures in a metamorphic terrane or orogen (e.g., northern Appalachians, Spear et al., [2002]; Naxos, Buick and Holland, [1989]; ThorOdin, British Columbia, Norlander et al., [2002]). In some gneiss domes, a common difference between the dome core and mantling rocks is that the dome rocks may record near-isothermal decompression from kyanite to sillimanite-cordierite zone conditions. Mantling metamorphic rocks may also have experienced high-temperature metamorphism, but P-T paths involved cooling during decompression (Teyssier and Whitney, 2002). The contact between the core and metamorphic mantling rocks is commonly represented by a distinctive metasedimentary unit such as a metaconglomerate (e.g., some Precambrian domes in Canada and Finland), quartzite (Baltimore Gneiss, Appalachians; North American Cordillera; Yangtze block, China), or marble (Aegean/Anatolian domes). In some cases, these distinctive units, which are comprised of lithologies that do not melt during orogenesis, can be traced continuously over tens of kilometers around the domes (Fig. 3A), but the same lithologies are discontinuous within the migmatitic core (e.g., quartzite in the Cordilleran domes; marble in the Aegean Naxos dome). Particularly in examples where metaconglomerate is present, the core/cover contact has been interpreted as a deformed unconformity (Eskola, 1949). These distinctive layers also have geologic and geodynamic significance because they are convenient stratigraphic markers that can be used to track the structural evolution of the dome versus mantling rocks and to document the rheology of the dome and surrounding rocks. The occurrence of these markers both outside and inside the domes can be interpreted in various ways: the dome did not move significantly relative to the mantling gneisses; the dome brought up stratigraphic units that were part of lower thrust sheets (e.g., developed during crustal thickening) containing the same stratigraphic markers; or the dome brought up rocks from the mantling gneisses that were first involved in downflow around a diapir before being incorporated in the upward flow within the dome (partial convection). Spatial Characteristics of Gneiss Domes and Gneiss Dome Suites Typically, more than one gneiss dome is present in a region. In some orogens, the domes occur in approximately linear belts along the strike of the orogen (e.g., northern North American Cordillera; southern and eastern Tibet). In other regions, the
5
domes are distributed more broadly (e.g., Karakoram region of the Himalaya; Massif Central of the French Variscides), with either apparently random distribution (Karakoram) or the occurrence of overlapping domes in en echelon arrays (Pyrenees of Spain and France; eastern Finland) (see Yin, this volume, for further discussion of gneiss dome systems). Gneiss domes worldwide vary considerably in dimension. Although a few domes are approximately equant, most are elongate and elliptical and are aligned parallel to the strike of the orogen. Dramatic examples of this along-orogen elongation are some of the northern Appalachian domes. Domes within wrench zones (e.g., Red River shear zone; Leloup et al. [2001]) are also highly elongate, with aspect ratios exceeding 10:1. A survey of 205 gneiss domes worldwide, including Precambrian and Phanerozoic examples, documents an approximately linear correlation between the map-view long and short axes of domes for gneiss domes with long dimensions up to 90 km (Fig. 5). Beyond 90 km, the correlation is weak to absent. For gneiss domes with principal axes 90 km do not follow this trend. B. Axial ratio (long/short) versus an areal parameter (long × short) for the domes graphed in A, showing that the elliptical shape is independent of overall dome size (see similar graph in Brun [1980] for domes in Finland). Many of the domes that exceed an axial ratio of 3.5 are narrow, elongate domes in the northern Appalachians.
• the Hercynian/Variscan belt in Spain (Pyrenees), France (Pyrenees, Massif Central, Brittany), and central/eastern Europe (Bohemian Massif: Erzgebirge Massif and Sudeten Mountains, Germany, Czech Republic, Poland); • the Bering Sea region (Alaska, Russia); • the Appalachians/Grenville orogens of eastern North America; • the Caledonides of Greenland, Ireland, and Norway; • the Paleozoic of central Australia (e.g., Entia dome); • Antarctica (Shackleton Range; Grove Mountains; Marie Byrd Land); • and all Precambrian cratons/orogens: Australia (Pilbara, Yilgarn, E Australia), Brazil (Sao Francisco, Borborema Province, Goias), Canada (e.g., Superior, Slave, Grenville), Scandinavia (southern and southeast Finland), India (e.g., Dharwar), Africa (including the Pan-African orogen and Archean cratons), China, Russia, Saudi Arabia, and the United States. The above list is geographically organized, but it is worth noting that gneiss domes also commonly occur in highpressure to ultrahigh-pressure terranes such as those in China (Qinling-Dabieshan-Sulu domes; e.g., Faure et al. [2003]), Kazakhstan, the Himalaya (Tso Morari; de Sigoyer et al. [1997]), and Norway (Labrousse et al., this volume).
ORIGIN OF GNEISS DOMES The origin of gneiss domes has been much debated (Eskola, 1949), with preferred explanations for their formation and unroofing changing by the decade. Up to the early 1980s, many domes were thought to form by diapirism (Berner et al., 1972; Fletcher, 1972; Dixon, 1975; Brun et al., 1981; Soula, 1982), and this idea has reappeared in recent years (Bouhallier et al., 1995; Burg and Vanderhaeghe, 1993; Weinberg and Podladchikov, 1994; Chardon et al., 1998; Calvert et al., 1999; Vanderhaeghe and Teyssier, 2001; Teyssier and Whitney, 2002). Diapirism has been commonly discounted as a possible mechanism for the generation of gneiss domes because, in many cases, the structural doming event is believed to post-date partial melting/magmatism (e.g., Lee et al., 2000; Rolland et al., 2001; Fisher and Olsen, this volume). More commonly in the last two decades, gneiss domes have been viewed as either contractional or extensional. For example, gneiss dome genesis has been ascribed to crustal shortening (fold interference: Ramsay, 1967; Burg et al., 1984; Rolland et al., 2001; buckling: Stipska et al., 2000), extension (e.g., related to metamorphic core complex formation: Chen et al., 1990; Brun and Van Den Driessche, 1994; Escuder Viruete et al., 2000; Yan et al., 2003), and extension-assisted diapiric ascent of partially molten crust during orogenic collapse (Vanderhaeghe et al.,
7
Gneiss Domes and Crustal Flow 1999). Recent models invoke the role of erosion coupled with crustal folding (Burg et al., this volume), as well as multi-stage models involving both contraction and extension (Lee et al., 2000) and extrusion of orogenic crust (“leaky channel,” Beaumont et al., 2001). Relating Gneiss Dome Characteristics to Processes Structural data that provide evidence for the trajectory, conditions, and mechanisms of flow are important for evaluating dome-forming mechanisms. Interpretation of structural elements must consider the relationship of the structures to the threedimensional geometry of the dome and evaluate the kinematics of flow of material in the dome and surrounding metamorphic rocks. Maps of foliation trajectory may reveal the presence of subdomes (Fig. 4) (Lagarde et al., 1994; Vanderhaeghe, this volume) and other features important for interpreting the origin and internal dynamics of domes, but foliation and, in particular, lineation trajectories are required to test diapiric flow versus fold interference models of doming. Diapirs display a specific set of fabric orientations (e.g., radial lineations), finite strain distribution (e.g., flattening or constriction at particular localities), and kinematic relations (diapir up) that distinguish them from interference among folds in general. Fold interference is expected to produce lineation patterns predictable from a doubly folded surface and finite strain patterns that are relatively uniform or in which variations are not related to position in the dome. During �
crustal folding, kinematics in the flank of the dome should be attributable to flexural flow (dome-down relations). However, recent work on the effect of erosion on crustal folding by Burg et al. (this volume) shows that dome-up sense of shear and cascading folds may occur during fold growth as well, suggesting that shear sense in the flanks of the domes is not a general diagnostic criterion for dome formation. The pressure-temperature-time path of the rocks in the core and, if possible, the metamorphic cover are critical for evaluating dome-forming process(es), as well as for assessing the relationship of metamorphism to deformation and melting. Different P-T-t paths are predicted for diapirism versus other dome-forming mechanisms (Teyssier and Whitney, 2002). For example, diapiric flow results in decompression at near-isothermal conditions and may be associated with dehydration melting of crustal protoliths (Fig. 6). Because gneiss domes may experience decompression at high temperatures, culminating in late low-pressure–hightemperature metamorphic conditions, it may be difficult to determine the maximum pressure. Reaction textures involving Al2SiO5 polymorphs, garnet + cordierite, and garnet + plagioclase are helpful, including coronas/symplectite around garnet (e.g., Hollister, 1977; Brown, 1983) and fracture patterns around inclusions in garnet (Whitney et al., 2000). Evidence for polymetamorphism must be sifted to determine what textures, assemblages, and conditions are associated with domerelated processes.
�
���������������������������������������
�����������������
�
������������ �′ �
�′ �
�
������
�
�′ �
�
� �′ �
�′ �′ �
�′
�
��� � ���� ������� ����� �
��
��′ �
�
������������� �
� �′′
�′′
�′′
�′′
�
� �′′
�
� �′′ �
�′′ �′′
��� � ���� ������� ����� �
����������
�
�������������
Figure 6. Flow of lower crust beneath a rigid lid (A, B), and during crustal extension (C, D), in a channel and a diapir. Expected relative P-T-t paths are shown for lateral (channel) flow versus vertical (diapiric) flow. A-Aʹ-Aʺ are snapshots of channel flow and do not illustrate the entire history.
8
D.L. Whitney, C. Teyssier, and O. Vanderhaeghe
Obtaining the timing of metamorphism and deformation of the core gneisses and granitoids is essential for linking P-Tdeformation information to a dome-forming mechanism. Many gneiss domes contain core rocks that have experienced multiple metamorphic events, and zircons and other high-temperature minerals in the core may give old ages (unrelated to doming) even if the latest thermal event to affect the dome was much younger. For example, magmatism and crustal melting in the Naxos dome occurred in the Miocene, but zircons in granitic rocks and leucosomes also record a Paleozoic (Hercynian) event (Keay et al., 2001). Similarly, migmatite leucosomes in the Thor-Odin dome, British Columbia, Canada, contain Precambrian zircons with thin Eocene rims and overgrowths (Vanderhaeghe et al., 1999), and migmatites in the Nig˘de dome, central Anatolia, contain zircons with Paleozoic to Archean cores surrounded by thin, discontinuous, Late Cretaceous rims (Whitney et al., 2003) (Fig. 7). In general, U-Pb spot analyses of zircon rims and monazite in leucosomes, leucogranites, and host rocks provide the best indication of the timing of latest metamorphic and magmatic events. High-temperature geochronology can be linked to results from lower temperature thermochronometers to interpret cooling rates and mechanisms in the context of the overall P-T-t path. It is important to know whether the post-metamorphic peak cooling history is slow to moderate (50 °C/m.y.). Slow to moderate rates suggest that the dome-forming mechanism did not involve vertical transport of hot crust beyond the middle orogenic crust and exhumation occurred at a slow rate. Rapid cooling may indicate the shallow emplacement of hot, isothermally decompressed crust and subsequent rapid cooling and/or denudation.
layer. A positive feedback between decompression due to diapiric flow and partial melting may be responsible for sustained melting and buoyant flow. It is important to distinguish between migmatitic and magmatic diapirs, although there is a continuous spectrum between the two, depending on melt fraction/degree of crystallization and P-T-X conditions of the surrounding rock. Migmatite diapirs do not intrude their country rock, except locally: they deform along with their country rock during diapiric flow. This indicates that the viscosity contrast between a migmatite diapir and the mantling rocks, which are also typically migmatitic, is not large. Such domes can only form in cases where the overlying rocks are also weak. At the other end of the spectrum, magmatic bodies with significant viscosity contrast with their country rocks ascend by intrusive processes that may involve fracturing (Weinberg, 1996). Domes may form by the combined effect of various processes. For example, diapiric bodies may nucleate on fold instabilities, and inversely, creating domes that form by a combination of folding and buoyant ascent, with the potential of a positive feedback between these two mechanisms. Similarly, diapirs may be associated with extensional instabilities, such as crustal boudinage, resulting in domes nested within metamorphic core complexes. Diapiric doming may also be influenced by plate tectonic or orogen-scale lateral forces that affect the balance between vertical and lateral crustal flow. For example, a layer of partially molten crust may form after crustal thickening; this layer may flow in a channel and allow the orogen to grow laterally (Royden, 1996; Vanderhaeghe and Teyssier, 2001). Flow within this channel can be lateral or vertical.
Diapirism Revisited
DOMING AND CRUSTAL FLOW: RATES OF VERTICAL AND LATERAL CRUSTAL FLOW
Because many gneiss domes are cored by migmatite and granite and are comprised of lower density rocks (felsic gneiss) structurally overlain by higher density (metasedimentary/ metavolcanic) rocks, the role of diapirism in the formation of gneiss domes requires further discussion. Some field studies support diapirism (e.g., Brun et al., 1981; Hippertt, 1994; Bouhallier et al., 1995), and there is a sound mathematical and physical experimental base for understanding diapiric flow (Fletcher, 1972; Dixon, 1975; Schmeling et al., 1988; Jackson and Talbot, 1989). Fletcher (1972, after Bean, 1953) reported density contrasts for some New England gneiss domes of ~0.1 g cm−3 between dome and surrounding metapelitic rocks. More recently, Soula et al. (2001) suggested that metasedimentary rocks (2.7– 2.9 g cm−3) are likely to be denser than felsic basement (2.5–2.7 g cm−3), although the relevant parameters are the densities at high pressures (crustal depths). Density inversion can drive diapirism efficiently if the viscosities of the two layers are close, which is the case if both layers are partially molten. In addition, if melt fraction controls viscosity, the lower viscosity (diatexite) layer may move easily upward and deform the overlying metatexite
Crustal flow is a significant process in orogeny because it controls the (re)distribution of heat and mass, the thickness of the crust through time, and the evolution of la ndscapes (e.g., formation and collapse of orogenic plateaux). The relative amount and rate of vertical (diapiric) versus lateral (channel) flow are a function of the rheology of the crust and the balance between gravitational forces and other driving mechanisms for crustal flow (e.g., lateral pressure gradients). The major differences between channel flow and diapiric flow can be seen in P-t, T-t, and P-T diagrams (Fig. 6): channel flow by itself does not result in a change in temperature or pressure (horizontal paths on the P-t and T-t diagrams that accompany the rigid lid example), but may be accompanied by some cooling and decompression if the upper crust is extending and the ductile crust is thinning/cooling (moderately sloping paths to Aʺ in P-t and T-t diagrams that accompany the extension example). Vertical crustal flow results in decompression in both cases and may be dramatic (steeper path, greater magnitude) in the extension case. The cooling paths will differ between the two cases once the region of vertically flowing crust (which may be partially
9
Gneiss Domes and Crustal Flow
����µ�
�
�
����µ�
�
Figure 7. Cathodoluminescence image of zircon from the Nig˘de gneiss dome, Turkey, showing how U-Pb sensitive high-resolution ion microprobe (SHRIMP) analysis detected thin Late Cretaceous rims on older cores. Ages are in millions of years. A. Zircon from a crustally derived granite; B. Zircons from a migmatitic sillimanite gneiss. Both samples are from the core of the dome, and the Late Cretaceous ages (88–93 Ma) represent the timing of high-grade metamorphism, crustal melting, and doming. Data are reported in Whitney et al. (2003).
SYNTHESIS The origin of domes and their significance in orogeny can be evaluated for individual gneiss domes or dome systems by determining the thermal-mechanical relationship between doming, P-T paths, and melting, and by documenting the structural relationship among domes, subdomes, and mantling rocks. The magnitude, rate, and thermal consequences of the flow can be determined from the decompression record of dome rocks, the timing of melting in the P-T path, and the nature and kinematics of structures within the dome. The domal geometry and the association of domes with migmatites imply fundamental ties to crustal melting and flow, and domes can be studied for information about the relative rates of vertical versus lateral crustal flow. The ubiquitous occurrence of gneiss domes in orogens strongly suggests that vertical flow of partially molten crust is a common mode of material and heat transfer in orogeny. Ultimately, domes need to be assessed in terms of their role in stabilizing the continental crust by advection of heat and material in orogens. We propose that gneiss dome genesis is a first-order process that fundamentally affects the thermal budget of orogens by the emplacement of hot, formerly deep crust at shallower crustal levels, influencing the vertical and lateral distribution
vertical flow dominates
�������������������������������������������������������
molten) interacts with the cooler upper crust (steepening of T-t diagram and cooling segment of P-T path to Bʺ). Diapirs represent end-member vertical crustal flow. Endmember lateral flow produces regions of flat foliation, which is typically migmatitic. Crustal flow with both lateral and vertical components may result in extrusion of material (Ramsay and Huber, 1987; Beaumont et al., 2001; Rey, 2001), and the structural features exhibited by exhumed orogenic crust vary as a function of the relative rates of vertical and lateral crustal flow (Figs. 7 and 8). These rates, along with other variables such as viscosity contrast and effects of surface processes, also influence the magnitude of vertical flow. Figure 8 illustrates a possible interaction between channel flow and vertical (diapiric) flow. If the rate of lateral flow is high relative to diapiric flow, domes may not form. If the diapiric instability is fast relative to lateral flow, however, domes may form readily and reach shallow levels where they cool rapidly and preserve their diapiric structure. Intermediate scenarios include the formation of domes that are drawn into a recumbent shape by lateral flow. Large-scale (~10 km) recumbent and sheath folds recognized in some high-grade terrains (Dirks et al., 1997) may be smeared out structures that initiated as domes. According to this conceptual model, gneiss domes survive as domes because diapiric ascent is sufficiently rapid to create and preserve structures related to vertical flow and because the domes are emplaced at sufficiently shallow levels in the crust to crystallize and preserve their domal structure. Studies of exhumed migmatitic complexes confirm that partially molten crust may cool very rapidly following high-temperature decompression (Brown and Dallmeyer, 1996). The precise interaction between the partially molten diapir and the cool, brittle upper crust needs further investigation. We propose that rapid vertical flow of partially molten crust, followed by rapid cooling upon emplacement of migmatitic crust at shallower levels, are common phenomena in orogens, accounting for the ubiquity of gneiss domes.
���������������
��������������
lateral (channel) flow dominates
Figure 8. Schematic diagram showing the range of structures produced as a function of vertical versus lateral crustal flow. Domal structures (e.g., diapirs) characterize vertical flow-dominated systems. Migmatite nappes represent an intermediate case with significant components of vertical and lateral flow. Horizontal fabrics characterize systems dominated by lateral flow.
10
D.L. Whitney, C. Teyssier, and O. Vanderhaeghe
of heat-producing elements and the stabilization of continents. Regarding heat and mass transfer, gneiss dome genesis is analogous to magmatic, erosional/sedimentary, and other surface and tectonic processes (Sandiford and McLaren, 2002). Once the pressure-temperature-time–deformation histories of domes are better characterized in the context of orogenic evolution, the next step is to evaluate questions of magnitude and rates of heat and mass transport and the dimensions and scales of crustal flow, to document the rates and durations of melting and transport of melt (segregation), or partial melt (in mass motion), and to determine if there are characteristic rates and magnitudes for decompression and cooling of partially molten domes. These data and analyses can renew our understanding of the significance of dome formation for compositional and rheological structuring of continental crust. ACKNOWLEDGMENTS This work has been supported by National Science Foundation grants EAR- 9814669, EAR-0106953, and EAR-0106667 to CT and DLW. This paper was greatly improved by constructive reviews and editorial comments by Mike Brown, Mike Sandiford, Carlo Alberto Ricci, and Christine Siddoway. We are also grateful for the assistance of all those who gave us information and advice for the Gneiss Dome Atlas (Appendix), including W. Schwerdtner, Cynthia Dusel-Bacon, and John Percival. REFERENCES CITED Arnold, J., Sandiford, M., and Wetherley, S., 1995, Metamorphic events in the eastern Arunta Inlier; Part 1, Metamorphic petrology: Precambrian Research, v. 71, p. 183–205, doi: 10.1016/0301-9268(94)00061-U. Bean, R.J., 1953, Relation of gravity anomalies to the geology of central Vermont and New Hampshire: Geological Society of America Bulletin, v. 64, p. 509–537. Beaumont, C., Jamieson, R.A., Nguyen, M.H., and Lee, B., 2001, Himalayan tectonics explained by extrusion of a low-viscosity crustal channel coupled to focused surface denudation: Nature, v. 414, p. 738–742, doi: 10.1038/414738A. Berner, H., Ramberg, H., and Stephansson, O., 1972, Diapirism in theory and experiment: Tectonophysics, v. 15, p. 197–218, doi: 10.1016/00401951(72)90085-6. Bouchez, J.L., 1997, Granite is never isotropic: an introduction to AMS studies of granitic rocks, in Bouchez, J.L., Hutton, D.H.W., and Stephens, W.E., eds., Granite: From Segregation of Melt to Emplacement Fabrics: Dordrecht, Kluwer Publishing, p. 95–112. Bouhallier, H., Chardon, D., and Choukroune, P., 1995, Strain patterns in Archaean dome-and-basin structures: The Dharwar craton (Karnataka, South India): Earth and Planetary Science Letters, v. 135, p. 57–75, doi: 10.1016/0012-821X(95)00144-2. Brown, M., 1983, The petrogenesis of some migmatites from the Presqu’ile de Rhuys, Southern Brittany, France, in Atherton, M.P., and Gribble, C.D., eds., Migmatites, Melting, and Metamorphism: Nantwich, Shiva, p. 174–200. Brown, M., and Dallmeyer, R.D., 1996, Rapid Variscan exhumation and the role of magma in core complex formation; southern Brittany metamorphic belt, France: Journal of Metamorphic Geology, v. 14, p. 361–379, doi: 10.1046/J.1525-1314.1996.06024.X. Brown, D., Van Gool, J., Calon, T., and Rivers, T., 1991, The geometric and kinematic development of the Emma Lake Thrust stack, Grenville front, southwestern Labrador: Canadian Journal of Earth Sciences, v. 28, p. 136–144.
Brun, J.-P., 1980, The cluster-ridge pattern of mantled gneiss domes in eastern Finland: Evidence for large-scale gravitational instability in the Proterozoic crust: Earth and Planetary Science Letters, v. 47, p. 441–449, doi: 10.1016/0012-821X(80)90032-1. Brun, J.-P., and Van Den Driessche, J., 1994, Extensional gneiss domes and detachment faults—structure and kinematics: Bulletin de la Société Géologique de France, v. 165, p. 519–530. Brun, J.-P., Gapais, D., and Le Theoff, B., 1981, The mantled gneiss domes of Kuopuo (Finland): Interfering diapirs: Tectonophysics, v. 74, p. 283–304, doi: 10.1016/0040-1951(81)90194-3. Buick, I.S., and Holland, T.J.B., 1989, The P-T-t path associated with crustal extension, Naxos, Cyclades, Greece, in Daly, J.S., Cliff, R.A., and Yardley, B.W.D., eds., Evolution of Metamorphic Belts: London, Geological Society Special Publication 43, p. 365–369. Burg, J.-P., Guiraud, M., Chen, G.M., and Li, G.C., 1984, Himalayan metamorphism and deformations in the north Himalayan Belt (southern Tibet, China): Earth and Planetary Science Letters, v. 69, p. 391–400, doi: 10.1016/0012-821X(84)90197-3. Burg, J.-P., and Vanderhaeghe, O., 1993, Structures and way-up criteria in migmatites with application to the Velay dome (French Massif Central): Journal of Structural Geology, v. 15, p. 1293–1301, doi: 10.1016/01918141(93)90103-H. Calvert, A., Gans, P.B., and Amato, J.M., 1999, Diapiric ascent and cooling of a sillimanite gneiss dome revealed by 40Ar/39Ar thermochronology: the Kigluaik Mountains, Seward Peninsula, Alaska, in Ring, U., et al., eds., Exhumation Processes: Normal Faulting, Ductile Flow, and Erosion: London, Geological Society Special Publication 154, p. 205–232. Chardon, D., Choukroune, P., and Jayananda, M., 1998, Sinking of the Dharwar Basin (south India): Implications for Archaean tectonics: Precambrian Research, v. 91, p. 15–39, doi: 10.1016/S0301-9268(98)00037-0. Chen, Z., Liu, Y., Hodges, K.V., Burchfiel, B.C., Royden, L.H., and Deng, C., 1990, The Kangmar Dome: A metamorphic core complex in southern Xizang (Tibet): Science, v. 250, p. 1552–1556. Collins, W.J., Van Kranendonk, M.J., and Teyssier, C., 1998, Partial convective overturn of Archean crust in the east Pilbara Craton, western Australia: Driving mechanisms and tectonic implications: Journal of Structural Geology, v. 20, p. 1405–1424, doi: 10.1016/S0191-8141(98)00073-X. Coward, M.P., 1981, Diapirism and gravity tectonics: report of a Tectonic Studies Group conference held at Leeds University, 25–26 March 1980: Journal of Structural Geology, v. 3, p. 89–95, doi: 10.1016/01918141(81)90059-6. Cruden, A.R., 1988, Deformation around a rising diapir modeled by creeping flow past a sphere: Tectonics, v. 7, p. 1091–1110. Cruden, A.R., 1990, Flow and fabric development during the diapiric rise of magma: Journal of Geology, v. 98, p. 681–698. Davis, G.A., and Lister, G.S., 1988, Detachment faulting in continental extension: perspectives from the southwestern U.S. Cordillera: Boulder, Colorado, Geological Society of America Special Paper 218, p. 133–159. de Sigoyer, J., Guillot, S., Lardeaux, J.-M., and Mascle, G., 1997, Glaucophane bearing eclogites in the Tso Morari dome (eastern Ladakh, NW Himalaya): European Journal of Mineralogy, v. 9, p. 1073–1083. Dirks, P.H.G.M., Zhang, J.S., and Passchier, C.W., 1997, Exhumation of highpressure granulites and the role of lower crustal advection in the North China craton near Datong: Journal of Structural Geology, v. 19, p. 1343– 1358, doi: 10.1016/S0191-8141(97)00044-8. Dixon, J.M., 1975, Finite strain and progressive deformation in models of diapiric structures: Tectonophysics, v. 28, p. 89–124, doi: 10.1016/00401951(75)90060-8. Escuder Viruete, J., Indares, A., and Arenas, R., 2000, P-T paths derived from garnet growth zoning in an extensional setting: an example from the Tormes Gneiss Dome (Iberian Massif, Spain): Journal of Petrology, v. 41, p. 1489–1515. Eskola, P.E., 1949, The problem of mantled gneiss domes: Quarterly Journal of the Geological Society of London, v. 104, p. 461–476. Faure, M., Lalevee, F., Gusokujima, Y., Iiyama, J.-T., and Cadet, J.-P., 1986, The pre-Cretaceous deep-seated tectonics of the Abukuma Massif and its place in the structural framework of Japan: Earth and Planetary Science Letters, v. 77, p. 384–398, doi: 10.1016/0012-821X(86)90148-2. Faure, M., Lin, W., Monie, P., Le Breton, N., Poussineau, S., Panis, D., and Deloule, E., 2003, Exhumation tectonics of the ultrahigh-pressure metamorphic rocks in the Qinling orogen in east China: New petrologicalstructural-radiometric insights from the Shandong Peninsula: Tectonics, v. 22, 3-1018, doi: 10.1029/2002TC001450.
Gneiss Domes and Crustal Flow Ferré, E.C., Teyssier, C., Jackson, M., Thill, J.W., and Rainey, E.S.G., 2003, Magnetic Susceptibility Anisotropy: a new petrofabric tool in migmatites: Journal of Geophysical Research, v. 107, doi: 10.1029/2002JB001790. Fletcher, R.C., 1972, Application of a mathematical model to the emplacement of mantled gneiss domes: American Journal of Science, v. 272, p. 197–216. Fowler, T.J., and Osman, A.F., 2001, Gneiss-cored interference dome associated with two phases of late Pan-African thrusting in the central Eastern Desert, Egypt: Precambrian Research, v. 108, p. 17–43, doi: 10.1016/S03019268(00)00146-7. Hippertt, J.F., 1994, Structures indicative of helicoidal flow in a migmatitic diapir (Bacao Complex, southeastern Brazil): Tectonophysics, v. 234, p. 169–196, doi: 10.1016/0040-1951(94)90210-0. Holdsworth, R.E., 1989, The geology and evolution of a Caledonian fold and ductile thrust zone, Kyle of Tongue region, Sutherland, Northern Scotland: Journal of the Geological Society of London, v. 146, p. 809–823. Hollister, L.S., 1977, The reaction forming cordierite from garnet, the Khtada Lake metamorphic complex, British Columbia: Canadian Mineralogist, v. 15, p. 217–229. Jackson, M.P.A., and Talbot, C.J., 1989, Anatomy of mushroom-shaped diapirs: Journal of Structural Geology, v. 11, p. 211–230, doi: 10.1016/01918141(89)90044-8. Keay, S., Lister, G., and Buick, I., 2001, The timing of partial melting, Barrovian metamorphism and granite intrusion in the Naxos metamorphic core complex, Cyclades, Aegean Sea, Greece: Tectonophysics, v. 342, p. 275–312, doi: 10.1016/S0040-1951(01)00168-8. Lagarde, J.-L., Dallain, C., Ledru, P., and Courrioux, G., 1994, Strain pattern within the Variscan granite dome of Velay, French Massif Central: Journal of Structural Geology, v. 16, p. 839–852, doi: 10.1016/01918141(94)90149-X. Lee, J., Hacker, B.R., Dinklage, W.S., Wang, Y., Gans, P., Calvert, A., Wan, J., Chen, W., Blythe, A.E., and McLelland, W., 2000, Evolution of the Kangmar Dome, southern Tibet: Structural, petrologic, and thermochronologic constraints: Tectonics, v. 19, p. 872–895, doi: 10.1029/1999TC001147. Leloup, P.H., Arnaud, N., Lacassin, R., Kienast, J.R., Harrison, T.M., Pan Trong, T.T., Replumaz, A., and Tapponnier, P., 2001, New constraints on the structure, thermochronology, and timing of the Ailao Shan-Red River shear zone, SE Asia: Journal of Geophysical Research, v. 106, p. 6683–6732, doi: 10.1029/2000JB900322. Miller, D.M., 1980, Structural study of the northern Albion Mountains, southcentral Idaho: Geological Society of America Memoir 153, p. 399–423. Nelson, K.D., Zhao, W., Brown, L.D., Kuo, J., Jinkai, C., Xianwen, L., Klemperer, S.L., Makovsky, Y., Meissner, R., Mechie, J., Kind, R., Wenzel, F., Ni, J., Nabelek, J., Chen, L., Handong, T., Wenbo, W., Jones, A.G., Booker, J., Unsworth, M., Kidd, W.S.F., Hauck, M., Alsdorf, D., Ross, A., Cogan, M., Changde, W., Sandvol, E., and Edwards, M., 1996, Partially molten middle crust beneath southern Tibet: Synthesis of Project INDEPTH results: Science, v. 274, p. 1684–1688, doi: 10.1126/ SCIENCE.274.5293.1684. Norlander, B.N., Whitney, D.L., Teyssier, C., and Vanderhaeghe, O., 2002, Partial melting and decompression of the Thor-Odin dome, Shuswap metamorphic core complex, Canadian Cordillera: Lithos, v. 61, p. 103–125, doi: 10.1016/S0024-4937(02)00075-0. Ramberg, H., 1980, Diapirism and gravity collapse in the Scandinavian Caledonides: Journal of the Geological Society of London, v. 137, p. 261–270. Ramsay, J.G., 1967, Folding and Fracturing of Rocks: New York, McGraw-Hill, 568 p. Ramsay, J.G., and Huber, M.I., 1987, The techniques of modern structural geology, Volume 2: Folds and fractures: London, Academic Press, 700 p. Rey, P., 2001, From continental thickening and divergent collapse to active continental rifting, in Miller, J.A., et al., eds., Continental Reactivation and Reworking: London, Geological Society Special Publication 184, p. 77–88. Rey, P., Vanderhaeghe, O., and Teyssier, C., 2001, Gravitational collapse of the continental crust: Definition, regimes and modes: Tectonophysics, v. 342, p. 435–449, doi: 10.1016/S0040-1951(01)00174-3. Rolland, Y., Mahéo, G., Guillot, S., and Pêcher, A., 2001, Tectono-metamorphic evolution of the Karakorum Metamorphic complex (Dassu-Askole area, NE Pakistan): Exhumation of mid-crustal HT-MP gneisses in a convergent context: Journal of Metamorphic Geology, v. 19, p. 717–737. Royden, L.H., 1996, Coupling and decoupling of crust and mantle in convergent orogens: Implications for strain partitioning in the crust: Journal of Geophysical Research, v. 101, p. 17679–17705, doi: 10.1029/96JB00951.
11
Saltzer, S.D., and Hodges, K.V., 1988, The Middle Mountain shear zone, southern Idaho: kinematic analysis of an early Tertiary high-temperature detachment: Geological Society of America Bulletin, v. 100, p. 96–103, doi: 10.1130/0016-7606(1988)1002.3.CO;2. Sandiford, M., and McLaren, S., 2002, Tectonic feedback and the ordering of heat producing elements within continental lithosphere: Earth and Planetary Science Letters, v. 204, p. 133–150, doi: 10.1016/S0012821X(02)00958-5. Schmeling, H., Cruden, A.R., and Marquart, G., 1988, Finite deformation in and around a fluid sphere moving through a viscous medium; implications for diapiric ascent: Tectonophysics, v. 149, p. 17–34, doi: 10.1016/00401951(88)90116-3. Skehan, J.W., 1961, The Green Mountain anticlinorium in the vicinity of Wilmington and Woodford, Vermont: Bulletin of the Vermont Geological Survey, p. 159. Soula, J.C., 1982, Characteristics and mode of emplacement of gneiss domes and plutonic domes in central-eastern Pyrenees: Journal of Structural Geology, v. 4, p. 313–342, doi: 10.1016/0191-8141(82)90017-7. Soula, J.-C., Debat, P., Brusset, S., Bessier, G., Christophoul, F., and Deramond, J., 2001, Thrust-related, diapiric, and extensional doming in a frontal orogenic wedge: Example of the Montagne Noire, Southern French Hercynian Belt: Journal of Structural Geology, v. 23, p. 1677–1699, doi: 10.1016/S0191-8141(01)00021-9. Spear, F.S., Kohn, M.J., Cheney, J.T., and Florence, F., 2002, Metamorphic, thermal, and tectonic evolution of central New England: Journal of Petrology, v. 43, p. 2097–2120, doi: 10.1093/PETROLOGY/43.11.2097. Steck, A., Epard, J.L., Vannay, J.C., Hunziker, J., Girard, M., Morard, A., and Robyr, M., 1998, Geological transect across the Tso Morari and Spiti areas: The nappe structures of the Tethys Himalaya: Eclogae Geologicae Helvetiae, v. 91, p. 103–122. Stipska, P., Schulmann, K., and Hock, V., 2000, Complex metamorphic zonation of the Thaya dome: result of buckling and gravitational collapse of an imbricated nappe sequence, in Cosgrove, J.W., and Ameen, M.S., eds., Forced Folds and Fractures: London, Geological Society Special Publication 169, p. 197–211. Talbot, C.J., 1974, Fold nappes as asymmetric mantled gneiss domes and ensialic orogeny: Tectonophysics, v. 24, p. 259–276, doi: 10.1016/00401951(74)90011-0. Thompson, J.B., Robinson, P., Clifford, T.N., and Trask, N.J., 1968, Nappes and gneiss domes in west-central New England, in Zen, E-An et al., eds., Studies of Appalachian Geology: Northern and maritime: New York, Interscience Publishers, p. 203–218. Teyssier, C., and Whitney, D., 2002, Gneiss domes and orogeny: Geology, v. 30, p. 1139–1142, doi: 10.1130/0091-7613(2002)0302.0.CO;2. Vanderhaeghe, O., Teyssier, C., and Wysoczanski, R., 1999, Structural and geochronological constraints on the role of partial melting during the formation of the Shuswap metamorphic core complex at the latitude of the Thor-Odin Dome, British Columbia: Canadian Journal of Earth Sciences, v. 36, p. 917–943, doi: 10.1139/CJES-36-6-917. Vanderhaeghe, O., and Teyssier, C., 2001, Partial melting and flow of orogens: Tectonophysics, v. 342, p. 451–472, doi: 10.1016/S0040-1951(01)001755. van Staal, C.R., and Williams, P.F., 1983, Evolution of a Svecofennian mantled gneiss dome in SW Finland, with evidence for thrusting: Precambrian Research, v. 21, p. 101–128, doi: 10.1016/0301-9268(83)90007-4. Weinberg, R.F., 1996, Ascent mechanism of felsic magmas: news and views: Transactions of the Royal Society of Edinburgh, v. 87, p. 95–103. Weinberg, R.F., and Podladchikov, Y.Y., 1994, Diapiric ascent of magmas through power-law crust and mantle: Journal of Geophysical Research, v. 99, p. 9543–9559, doi: 10.1029/93JB03461. Whitney, D.L., Cooke, M.L., and Du Frane, S.A., 2000, Modeling of radial fracturing at corners of inclusions in garnet using fracture mechanics: Journal of Geophysical Research, v. 105, p. 2843–2853, doi: 10.1029/ 1999JB900375. Whitney, D.L., Teyssier, C., and Fayon, A.K., 2004, Isothermal decompression, partial melting, and the exhumation of deep continental crust, in Grocott, J., et al., eds., Vertical Coupling and Decoupling in the Lithosphere: London, Geological Society Special Publication 227, p. 313–326. Whitney, D.L., Teyssier, C., Fayon, A.K., Hamilton, M.A., and Heizler, M., 2003, Tectonic controls on metamorphism, partial melting, and intrusion: timing and duration of regional metamorphism and magmatism in the Nig˘de Massif, Turkey: Tectonophysics, v. 376, p. 37–60, doi: 10.1016/ J.TECTO.2003.08.009.
12
D.L. Whitney, C. Teyssier, and O. Vanderhaeghe
Yan, D.-P., Zhou, M.-F., Song, H., and Fu, Z., 2003, Structural style and tectonic significance of the Jianglang Dome in the eastern margin of the Tibetan Plateau, China: Journal of Structural Geology, v. 25, p. 765–779, doi: 10.1016/S0191-8141(02)00059-7. Yin, A., 1991, Mechanisms for the formation of domal and basinal detachment faults—a 3-dimensional analysis: Journal of Geophysical Research, v. 96, p. 14577–14594.
APPENDIX: ATLAS OF GNEISS DOMES Table A1 and Figure 1 contain information about gneiss dome locations based on maps and descriptions in the literature (see Appendix reference list). Domal structures listed were not necessarily described as “gneiss domes” by the authors of the original articles.
Although the main body of this paper is focused on gneissic domes containing anatectic migmatites, the Atlas is more inclusive. The Atlas (Table A1) contains data for domes consisting of high-grade metamorphic rocks, whether or not there is evidence for partial melting and/or magmatism at any stage of formation of the dome. For Precambrian examples, some domes are better described as granitic domes, but we have included these if they are described as structural domes (i.e., having a planar fabric radial to the core of the dome). Numbers on the map (Fig. 1) are keyed to numbers in Table A1, and the references are annotated with the gneiss domes described in each article. Gneiss domes whose dimensions were used in the generation of Figure 5 are listed in bold letters in Table A1.
13
Appendix TABLE A1. WORLD ATLAS OF GNEISS DOMES
TABLE A1. WORLD ATLAS OF GNEISS DOMES (continued) Russia
Precambrian Domes India [1]*
Dharwar: Chitradurga (3 domes), Guddetalakoppalu, Hanchihalli, Holenarsipur, Honnali, Honnavalli, Javanahalli, Malali, Ramanahalli, Rampura, Shimoga, Tarikere (Chadwick et al., 1991; Bouhallier et al., 1993, 1995; Chardon et al., 1996)
[2]
Kodarma (Bihar)
[3]
Seliberi (Rajasthan)
Australia [4]
Yilgarn: Agnew, Bullfinch-Parker area (Parker, Ghooli, Rankin), Dalgety, Marymia, Pioneer, Widgimooltha (Bickle and Archibald, 1984; Bloem et al., 1997)
[5]
Pilbara: Corunna Downs, Goodin, McPhee, Mt Edgar, North Pole, Shaw, Sylvania, Tambourah, Wyloo (Collins et al., 1998; van Kranendonk and Collins, 1998)
[6]
Northern Territories: Browns Range, Golden Dyke
[8]
Latvasiurian (Fennoscandia/Lake Ladoga)
[22]
Lipnyazh (Ukraine shield), [23] Sunnagin (Aldan Shield);
[24]
Tatar (Kola Peninsula); [25] Teya (Siberia), [26] Tonod (Baikal/ S margin Siberian craton)
United States [27]
Michigan: Watersmeet, Carney Lake, Republic, Peavy (Barovich et al., 1991; Schneider et al., 2004)
[27]
Wisconsin: Dunbar (+ unnamed domes) (Sims and Schulz, 1982)
[27]
Minnesota: McGrath (Holm and Lux, 1996) South Dakota: Bear Mountain, Park New York (Adirondacks): Snowy Mountain, Thirteenth Lake, Westport Vermont: Rayponda, Sadawga (Ratcliffe, 1990)
Africa [28]
Egypt: Maetiq/Meatiq, Wadi Hafafit (5 domes) (Sturchio et al., 1983)
E Finland: Kuopio-Joensuu area (18 domes) (Brun, 1980)
[29]
Morocco: Sirwa
[8]
Pitkaranta (several domes; Kareledic) (Eskola, 1949)
[30]
[9]
SW Finland (Svecofennides): Bergon, Mustio, Vaermalae (van Staal and Williams, 1983; Bleeker and Westra, 1987)
Ivory Coast: Reguibat (Caen-Vachette et al., 1984) Ghana: Navrongo (Attoh and Ekwueme, 1997)
[31]
Nigeria: Kushaka (Turner, 1983)
[32]
Gabon: Abamie, Diany Miyole (Ledru et al., 1989)
[33]
Malawi: Champira, Pinite Champina (Haslam et al., 1980) Zambia: Mpande, Luswisi (Mallick, 1967) Zimbabwe: unnamed (Ridley et al., 1997) Namibia: Geelkop
Scandinavia (Finland) [7]
Canada [10]
Alberta: Tulip Lake, Wylie Lake (Langenberg and Ramsden, 1980; McDonough et al., 2000)
[34]
[11]
Northwest Territories: Hackett River (Percival, 1979)
[35]
[12]
Manitoba: Herblet Lake, Defender Lake, Wapisu (Symons and Harris, 2000)
[13]
Saskatchewan: Carswell, Porter Lake
[14]
Yukon: King Solomon, Coal Creek
[15]
Quebec: Renia, Jalobert, Watshishou (Gelinas and Perreault, 1985; Gervais et al., 2004)
[16]
N and W Ontario: Ash Bay, Aulneau, Cedar Lake, Dryberry, Law Lake, Mystery, Northern, Rainy Lake/Rice Bay, Shaw, Twilight (Schwerdtner et al., 1978; Muir, 1979; Shanks, 1986; Evins, 2000)
[17]
SE Ontario: Anstruther, Burleigh, Preston East
[18]
Labrador (2 unnamed domes)
[20]
[21]
Madagascar: unnamed domes Saudi Arabia: Arabian-Nubian, Qadda [36]
China: Qianax/Qianan (N China) (Wang et al., 1988)
[37]
Greenland: Amitsoq, Kap Farvel (Moorbath et al., 1975) Sri Lanka: unnamed domes
North American Cordillera, United States and Canada [38]
British Columbia, Canada: Malton, Frenchman(s) Cap, Thor-Odin, Pinnacles, Valhalla/Passmore (Reesor and Moore, 1971; Chamberlain et al., 1980; Armstrong et al., 1991; Simony and Carr, 1997)
Brazil [19]
South Africa: Johannesberg-Pretoria
Sao Francisco craton: Bacao (Quadrilatero Ferrifero), Boa Vista, Mata Verde, Serra dos Meiras,Sete Voltas, St. Peters (Hippertt, 1994; Marshak et al., 1997) Borborema province: Alto do Brejo, Itabaiana, Jirau do Ponciano, Santana, Rio Itapicuru greenstone belt: Ambrosio, Araci, Barrocas, Nordestina, Pedra Alta, Salgadalia, Santa Luz (Baars, 1997) Simao Dias (Silva, 1995) Goias: Cristalina, Hidrolina others: Lajes
(continued)
[38]
Washington, USA: Lincoln, Okanogan/Kettle, Spokane/Priest River (WA-ID), (Rhodes and Cheney, 1981; Hansen and Goodge, 1988; Doughty and Price, 1999)
[38]
Montana, USA: Bitterroot (Hodges and Applegate, 1993)
[39]
Idaho, Utah, USA: Albion, Big Bertha, Grouse Ck, Raft River (Armstrong, 1968; Miller, 1980)
[40]
Alaska, USA: Amy, Big Delta, Ester, Gilmore, Ketchum, Mastodon, Salcha River, Slaven (Dusel-Bacon and Foster, 1983; Dusel-Bacon et al., 2003)
[41]
Seward Peninsula/Bering Strait: Kigluaik, Bendeleben, unnamed (Amato et al., 1994)
(continued)
14
Appendix TABLE A1. WORLD ATLAS OF GNEISS DOMES (continued)
Northeast Russia [42]
TABLE A1. WORLD ATLAS OF GNEISS DOMES (continued) [53]
Poland/Czech Repuplic (Bohemian Massif/W Sudetes): Desenska, Desna, Dyje, Keprnik, Orlica, Orlice-Klodzko, Snieznik (2), Svratka, Thaya (Frank, 1993; Maluski et al., 1995; Floyd et al., 1996; Konopasek et al., 2002)
[54]
Austria: Gleinalm, Sonnblick (Reddy et al., 1993; Neebauer et al., 1995)
[55]
Hungary: Szeghlom (Toth et al., 2000)
[56]
former Yugoslavia: Crnook (Bonev, 1999)
[56]
Bulgaria: Bjala-reka, Kasebir/Kesebir, Lyaskovo, Madan (Dimitrijevic, 1978; Kozhoukhharov, 1980)
[57]
Greece (Aegean): Ios, Naxos (Jansen and Schuiling, 1976; van der Maar and Jansen, 1983)
[58]
Turkey (Anatolia): Kazdag˘, Menderes, Nig˘de, Uludag˘ (Ketin, 1947; de Graciansky, 1966; Whitney and Dilek, 1998; Okay and Satır, 2000)
[59]
Algeria: Edough (Caby et al., 2001)
Bering Strait (Chukotka Peninsula): Alarmaut, Kool’en, Neshkan, Senyavin (Natal’in et al., 1999)
North American Appalachians, United States and Canada (including Grenville) [43]
Quebec, Canada: Lemieux, Notre Dame (McNeice et al., 1991)
[44]
Maine, USA: Tumbledown (Solar and Brown, 1995)
[44]
New Hampshire, USA: Alstead, Carter, Chandler Ridge, Croyden, Jefferson, Keene, Mascoma,
[44]
Smarts Mt, Swanzey (NH/VT), Unity (Thompson et al., 1968; Spear et al., 2002)
[44]
Vermont, USA: Athens, Chester, Guilford, Lebanon, Pomfret, Strafford, Vernon (VT/MA) (Thompson et al., 1968; Hepburn, 1986; Menard and Spear, 1994; Vyhnal and Chamberlain, 1996)
[44]
Massachusetts, USA: Goshen, Monson, Pelham, Shelburne Falls, Tully, Vernon (VT/MA), Warwick (Balk, 1946; Thompson et al., 1968; Moecher, 1999)
[44]
Connecticut, USA: Bristol, Collinsville, Glastonbury, Granby, Granville (MA/CT), Killingworth, Lyme, Maromas, Montville, Stony Creek, Waterbury, Willimantic (Wintsch, 1979; Dietsch, 1989; Moecher, 1999)
[45]
Maryland, USA (Baltimore Gneiss): Chattolanee, Clarksville, Mayfield, Phoenix, Texas, Towson, Woodstock (Kodama and Chapin, 1984; Olsen, 1999)
[45]
Pennsylvania, USA: Woodville (Eskola, 1949)
[45]
Delaware, USA: Mill Creek (Higgins et al., 1973)
[46]
Virginia, USA: Sherwill, State Farm (Hatcher, 1984; Farrar, 1984)
[46]
South and North Carolina, USA: Toxaway
European Caledonides [47]
Greenland: Niggli Spids (Leslie and Nutman, 2000)
[48]
Ireland: Carna, Galway-Kilkieran, Delaney (?) (Leake, 1978)
[49]
Norway: Akland, Dalsvatn, Einunnfjellet, Levang, Lonset, Moerheheia (Abrahamsen, 1985; McClellan, 1994)
Variscan/Hercynian and Alpine Orogens [50]
Spain: Lugo, Mina Afortunada, Betic elongate domes (Abalos and Eguiluz, 1992; Reche et al., 1998)
[51]
France/Spain (Pyrenees): Agly, Aston, Bossòst, Garonne/ Garona, Canegou, Chiroulet, Gavarnie-Heas-Barroude, Hospitalet, Lesponne, Pallaresa, Tormes/Zamora, unnamed domes (Zwart, 1978; Soula, 1982; Verhoef et al., 1984; Soula et al., 1986; Pouget, 1987, 1991; Carreras and Capella, 1994; Escuder Viruete et al., 2000; Mezger and Passchier, 2003)
[52]
France (Massif Central & Brittany): Levezou, Montagne Noire (Agout, Caroux), Plougenast, Tulle, Velay (Schuiling, 1960; Den Tex, 1975; Saunier, 1986; Burg, 1987; van den Driessche and Brun, 1992; Roig and Faure, 1998; Lagarde et al., 1994; Ledru et al., 2001)
[53]
Germany (Bohemian Massif): Annaberg/Aunaberg, Boellstein, Catherine, Freiberg, Saxonian (Chatterjee, 1960; Mlcoch and Schulmann, 1992; Stipska et al., 2000)
(continued)
Himalaya and Southeast Asia [60]
Pamir: Kongur Shan, Mustaghata (Brunel et al., 1994)
[60]
Karakoram (Pakistan): Askole-Panhma, Dassu, Hemasil, Ho Lungma, Mangol (Rolland et al., 2001)
[60]
Karakoram (SW Tibet): Gurla Mandhata (Murphy et al., 2002)
[60]
other, India/Pakistan: Gianbul, Loe Sar/Nanga Parbat; Tso Morari, Sanko (Ladakh); Bhazun (Zanskar); Sikkim, Suru (India) (Kündig, 1989; Gapais et al., 1992; Kumar et al., 1994; Steck et al., 1998; DiPietro et al., 1999; Robyr and Steck, 2001)
[60]
South Tibet: Kangmar/Kangma, Mabja, unnamed (Chen et al., 1990)
[61]
East Tibetan Plateau: Changqiang, Gongcai, Jianglang, Motianling, Qiaoziding, Qiasi, Xunlongbao, Yaside (Yan et al., 2003)
[62]
Vietnam: Bu Khang, Dai Nui Con Voi, Song Chay (Jolivet et al., 1999; Leloup et al., 2001; Maluski et al., 2001)
[63]
Thailand: Doi Inthanon (Dunning et al., 1995)
Miscellaneous [64]
Australia: Entia (Paleozoic) (Arnold et al., 1995)
[65]
Ultrahigh-P terrains: Toiman (Kazakhstan); Qinling-Dabieshan (China): Foping, Hannan, Jiulingshan, Lushan, Luotian, Shennongjia, Wugongshan, Yuexi (Faure et al., 1998) China: Danba domes (Songpan-Garze orogenic belt, W Sichuan)
[66]
Japan: Oshirabetsu, Otagiri, Samegawa (Faure et al., 1986) Antarctica: [67] Marie Byrd Land (Fosdick Mountains); [68] Shackleton Range (Fuchs); [69] Read Mountains; [70] Grove Mts (Marsh, 1984; Smith and Richard, 1991; Talarico and Kroner, 1999)
*Numbers in brackets [ ] are keyed to map in Figure 1.
Gneiss Domes and Crustal Flow REFERENCES CITED FOR WORLD ATLAS OF GNEISS DOMES Abalos, B., and Eguiluz, L., 1992, Structural geology of the Mina Afortunada gneiss dome (Badajoz-Cordoba shear zone, SW Spain): Annales Tectonicae, v. 6, p. 95–110. (Mina Afortunada dome, Spain) Abrahamsen, N., 1985, Possible types of rotations and translations in the Scandinavian Caledonides: Journal of Geodynamics, v. 2, p. 245–263, doi: 10.1016/0264-3707(85)90013-4. (Lonset dome, Norway) Amato, J.M., Wright, J.E., Gans, P.B., and Miller, E.L., 1994, Magmatically induced metamorphism and deformation in the Kigluaik gneiss dome, Seward Peninsula, Alaska: Tectonics, v. 13, p. 515–527, doi: 10.1029/ 93TC03320. (Kigluaik/Bering Strait domes, Alaska, United States) Armstrong, R.L., 1968, Mantled gneiss domes in the Albion Mountains, southern Idaho: Geological Society of America Bulletin, v. 79, p. 1295–1314. (Albion-Raft River-Grouse Creek, Idaho and Utah, United States) Armstrong, R.L., Parrish, R.R., van der Heyden, P., Scott, K., Runkle, D., and Brown, R.L., 1991, Early Proterozoic basement exposures in the southern Canadian Cordillera; core gneiss of Frenchman Cap, Unit I of the Grand Forks Gneiss, and the Vaseaux Formation: Canadian Journal of Earth Sciences, v. 28, p. 1169–1201. (Frenchman[s] Cap dome, Shuswap Complex, British Columbia, Canada) Arnold, J., Sandiford, M., and Wetherley, S., 1995, Metamorphic events in the eastern Arunta Inlier; Part 1, Metamorphic petrology: Precambrian Research, v. 71, p. 183–205, doi: 10.1016/0301-9268(94)00061-U. (Entia dome, Arunta block, Australia) Attoh, K., and Ekwueme, B.N., 1997. The West African Shield, in De Wit, M., and Ashwal, L.D., eds., Greenstone Belts: Oxford Monographs on Geology and Geophysics, v. 35, p. 517–528. (Ghana, Africa) Baars, F.J., 1997, The Sao Francisco Craton, in De Wit, M., and Ashwal, L.D., eds., Greenstone Belts: Oxford Monographs on Geology and Geophysics, v. 35, p. 529–557. (Rio Itapicuru Greenstone Belt: Ambrosio, Araci, Barrocas, Salgadalia domes, Brazil) Balk, R., 1946, Gneiss dome at Shelburne Falls, Massachusetts: Geological Society of America Bulletin, v. 57, p. 125–160. (Shelburne Falls, Goshen domes, Massachusetts, United States) Barovich, K.M., Patchett, P.J., Peterman, Z.E., Sims, P.K., and Day, W.C., 1991, Neodymium isotopic evidence for early Proterozoic units in the Watersmeet gneiss dome, northern Michigan. Temperature-pressure estimates of dynamically recrystallized rocks in the early Proterozoic Mountain shear zone, northeastern Wisconsin: U. S. Geological Survey Bulletin B 1904G,H, 23 p. (Watersmeet dome, Michigan, United States) Bickle, M.J., and Archibald, N.J., 1984, Chloritoid and staurolite stability; implications for metamorphism in the Archaean Yilgarn Block, Western Australia: Journal of Metamorphic Geology, v. 2, p. 179–203. (Pioneer, Widgiemooltha domes, Yilgarn block, W Australia) Bleeker, W., and Westra, L., 1987, The evolution of the Mustio gneiss dome, Svecofennides of SW Finland: Precambrian Research, v. 36, p. 227–240, doi: 10.1016/0301-9268(87)90022-2. (Mustio dome, Finland) Bloem, E.J.M., Dalstra, H.J., Ridley, J.R., and Groves, D.I., 1997, Granitoid diapirism during protracted tectonism in an Archaean granitoid-greenstone belt, Yilgarn Block, Western Australia: Precambrian Research, v. 85, p. 147–171, doi: 10.1016/S0301-9268(97)00034-X. (Bullfinch-Parker dome area: Parker, Ghooli, Rankin domes, W. Australia) Bonev, N., 1999, Extensional exhumation of metamorphic complexes in Kesebir gneiss dome (eastern Rhodope, South Bulgaria): Eos (Transactions, American Geophysical Union), v. 80, no. 46, p. 1066. (Kesebir/Kasebirdome, Rhodope Massif, Bulgaria) Bouhallier, H., Chardon, D., and Choukroune, P., 1995, Strain patterns in Archaean dome-and-basin structures: The Dharwar craton (Karnataka, South India): Earth and Planetary Science Letters, v. 135, p. 57–75, doi: 10.1016/0012-821X(95)00144-2. (Shimoga and Chitradurga domes, Dharwar, India) Bouhallier, H., Choukroune, P., and Ballevre, M., 1993, Diapirism, bulk homogeneous shortening and transcurrent shearing in the Archaean Dharwar Craton; the Holenarsipur area, southern India: Precambrian Research, v. 63, p. 43–58, doi: 10.1016/0301-9268(93)90004-L. (Honnali, Tarikere, Malali, Hanchihalli, Guddetalakoppalu, Javanahalli, Holenarsipur domes, Dharwar, India) Brun, J.-P., 1980, The cluster-ridge pattern of mantled gneiss domes in eastern Finland: Evidence for large-scale gravitational instability in the Protero-
15
zoic crust: Earth and Planetary Science Letters, v. 47, p. 441–449, doi: 10.1016/0012-821X(80)90032-1. (Kuopio-Joensuu area domes, Finland) Brunel, M., Arnaud, N., Tapponnier, P., Pan, Y., and Wang, Y., 1994, Kongur Shan normal fault; type example of mountain building assisted by extension (Karakoram Fault, eastern Pamir): Geology, v. 22, p. 707–710, doi: 10.1130/0091-7613(1994)0222.3.CO;2. (Kongur Shan, Mustaghata domes, Pamir) Burg, J.-P., 1987, Regional shear variation in relation to diapirism and folding: Journal of Structural Geology, v. 9, p. 925–934, doi: 10.1016/01918141(87)90002-2. (Levezou dome, Massif Central, France) Caby, R., Hammor, D., and Delor, C., 2001, Metamorphic evolution, partial melting and Miocene exhumation of lower crust in the Edough metamorphic core complex, West Mediterranean orogen, eastern Algeria: Tectonophysics, v. 342, p. 239–273, doi: 10.1016/S0040-1951(01)00166-4. (Edough massif, Algeria) Caen-Vachette, M., Tempier, P., and Camil, J., 1984, Rb/Sr age of 1670 Ma for mylonites of the Sassandra Event (Ivory Coast); effects on dating the final-Eburnean movements of the West African Craton: Journal of African Earth Sciences, v. 2, p. 359–363, doi: 10.1016/0899-5362(84)90009-5. (Reguibat dome, Ivory Coast) Carreras, J., and Capella, I.J., 1994, Tectonic levels in the Palaeozoic basement of the Pyrenees; a review and a new interpretation: Journal of Structural Geology, v. 16, p. 1509–1524, doi: 10.1016/0191-8141(94)90029-9. (Garonne/Garona dome, Pyrenees, France/Spain) Chadwick, B., Vasudev, V.N., Krishna, R.B., and Hedge, G.V., 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: Journal of the Geological Society of India, v. 38, p. 457–484. (Honnali dome, Dharwar, India) Chamberlain, V.E., Lambert, R.St.J., and Holland, J.G., 1980, The Malton Gneiss; Archean gneisses in British Columbia: Precambrian Research, v. 11, p. 297–306, doi: 10.1016/0301-9268(80)90069-8. (Malton, Shuswap Complex, British Columbia, Canada) Chardon, D., Choukroune, P., and Jayananda, M., 1996, Strain patterns, décollement and incipient sagducted greenstone terrains in the Archaean Dharwar Craton (South India): Journal of Structural Geology, v. 18, p. 991–1004, doi: 10.1016/0191-8141(96)00031-4. (Rampura, Honnavalli, Ramanahalli domes, Dharwar, India) Chatterjee, N.D., 1960, Geologische Untersuchungen im Kristallin des Boellsteiner Odenwaldes: Neues Jahrbuch fuer Geologie und Palaeontologie, Abhandlungen, v. 111, p. 137–180. (Boellstein dome, Germany) Chen, Z., Liu, Y., Hodges, K.V., Burchfiel, B.C., Royden, L.H., and Deng, C., 1990, The Kangmar Dome; a metamorphic core complex in southern Xizang (Tibet): Science, v. 250, no. 4987, p. 1552–1556. (Kangmar dome, Tibet) Collins, W.J., Van Kranendonk, M.J., and Teyssier, C., 1998, Partial convective overturn of Archean crust in the east Pilbara Craton, western Australia: Driving mechanisms and tectonic implications: Journal of Structural Geology, v. 20, p. 1405–1424, doi: 10.1016/S0191-8141(98)00073-X. (Mt Edgar, Corunna Downs domes, Australia) de Graciansky, P.C., 1966, Le massif cristallin du Menderes (Taurus occidental, Asie Mineure); un exemple possible de vieux socle granitique remobilize: Revue de Geographie Physique et de Geologie Dynamique, v. 8, p. 289–306. (Menderes Massif, Turkey) Den Tex, E., 1975, Thermally mantled gneiss domes; the case for convective heat flow in more or less solid orogenic basement, in Borradaile, G.J., Ritsema, A.R., Rondeel, H.E., and Simon, O.J., eds., Progress in geodynamics: Geodynamics Project, scientific report 13, p. 62–79. (Agout/Montagne Noire dome, Massif Central, France; Lepontine dome, Alps, Switzerland) Dietsch, C., 1989, The Waterbury Dome, west-central Connecticut: a triple window exposing deeply deformed, multiple tectonic units: American Journal of Science, v. 289, p. 1070–1097. (Waterbury, Collinsville, Granby domes, Connecticut, United States) Dimitrijevic, M.D., 1978, On the metamorphism in Yugoslavia, in Zwart, H.J., Sobolev, V.S., and Niggli, E. eds., Metamorphic map of Europe: Leiden and Paris, Subcommission for the Cartography of the Metamorphic Belts of the World and the United Nations Educational, Scientific and Cultural Organization, scale 1:25,000; explanatory text, p. 162–166. (Crnook dome, Yugoslavia) DiPietro, J.A., Pogue, K.R., Hussain, A., and Ahmad, I., 1999, Geologic map of the Indus syntaxis and surrounding area, Northwest Himalaya, Pakistan, in Macfarlane, A., Sorkhabi, R.B., and Quade, J., eds., Himalaya and Tibet;
16
D.L. Whitney, C. Teyssier, and O. Vanderhaeghe
mountain roots to mountain tops: Boulder, Colorado, Geological Society of America Special Paper 328, p. 159–178. (Loe Sar dome, Nanga Parbat/ Himalaya, Pakistan) Doughty, P., and Price, R.A., 1999, Tectonic evolution of the Priest River Complex, northern Idaho and Washington: a reappraisal of the Newport Fault with new insights on metamorphic core complex formation: Tectonics, v. 18, p. 375–393, doi: 10.1029/1998TC900029. (Spokane dome/Priest River Complex, Washington, United States) Dunning, G.R., Macdonald, A.S., and Barr, S.M., 1995, Zircon and monazite UPb dating of the Doi Inthanon core complex, northern Thailand; implications for extension within the Indosinian Orogen: Tectonophysics, v. 251, p. 197–213, doi: 10.1016/0040-1951(95)00037-2. (Doi Inthanon dome, Thailand) Dusel-Bacon, C., and Foster, H.L., 1983, A sillimanite gneiss dome in the Yukon crystalline terrane, east-central Alaska; petrography and garnet-biotite geothermometry: Reston, Virginia, U.S. Geological Survey Professional Paper P 1170-E, p. E1–E25. (Big Delta dome, Alaska, United States) Dusel-Bacon, C., Wooden, J.L., and Layer, P.W., 2003, A Cretaceous SHRIMP U-Pb zircon age for the West Point orthogneiss: Evidence for another gneiss dome in the Yukon-Tanana Upland, in Galloway, J.P., ed., Studies in Alaska by the U.S. Geological Survey during 2001: Reston, Virginia, U.S. Geological Survey Professional Paper 1678, p. 41–60. (Salcha River dome, Alaska, United States) Escuder Viruete, J., Indares, A., and Arenas, R., 2000, P-T paths derived from garnet growth zoning in an extensional setting: an example from the Tormes Gneiss Dome (Iberian Massif, Spain): Journal of Petrology, v. 41, p. 1489–1515. (Tormes dome, Spain) Eskola, P.E., 1949, The problem of mantled gneiss domes: Quarterly Journal of the Geological Society of London, v. 104, p. 461–476. (Pitkaranta/Lake Ladoga, Kuopio, Joensuu domes, Finland; Woodstock dome/Baltimore Gneiss, Maryland, United States; Alpine massifs: St. Gotthard, Aar, Mont Blanc, Gran Paradiso, Tauern; Variscan and Caledonian massifs) Evins, P.M., 2000, Structure of the Twilight—Mystery Lakes Gneiss Domes: Implications for the tectonic history of the Archean Winnipeg River Subprovince, NW Ontario, [M.Sc. thesis]: Toronto, Canada, University of Toronto, 151 p. (Twilight, Mystery, Cedar Lakes domes, western Superior Province, Ontario, Canada) Farrar, S.S., 1984, The Goochland granulite terrane: Remobilized Grenville basement in the eastern Virginia Piedmont, in Bartholomew, M.J., ed., The Grenville event in the Appalachians and related topics: Boulder, Colorado, Geological Society of America Special Paper 194, p. 215–229. (State Farm Gneiss, Virginia, United States) Faure, M., Lalevee, F., Gusokujima, Y., Iiyama, J.-T., and Cadet, J.-P., 1986, The pre-Cretaceous deep-seated tectonics of the Abukuma Massif and its place in the structural framework of Japan: Earth and Planetary Science Letters, v. 77, p. 384–398, doi: 10.1016/0012-821X(86)90148-2. (Samegawa dome, Japan) Faure, M., Lin, W., and Sun, Y., 1998, Doming in the southern foreland of the Dabieshan (Yangtse Block, China): Terra Nova, v. 10, p. 307–311, doi: 10.1046/J.1365-3121.1998.00207.X. (Wugongshan dome, Dabie Shan, China) Floyd, P.A., Winchester, J.A., Ciesielczuk, J., Lewandowska, A., Szczepanski, J., and Turniak, K., 1996, Geochemistry of early Palaeozoic amphibolites from the Orlica-Snieznik Dome, Bohemian Massif; petrogenesis and palaeotectonic aspects: Geologische Rundschau, v. 85, p. 225–238, doi: 10.1007/S005310050070. (Orlica-Snieznik domes/Bohemian Massif, Poland/Czech Republic) Frank, W., 1993, The Saxonian granulites—a metamorphic core complex: Geologische Rundschau, v. 82, p. 505–515. (Bohemian Massif). Gapais, D., Pêcher, A., Gilbert, E., and Ballever, M., 1992, Synconvergence spreading of the Higher Himalaya Crystalline in Ladakh: Tectonics, v. 11, p. 1045–1056. (Sanko dome, Ladakh, Pakistan) Gelinas, L., and Perreault, S., 1985, The metamorphic evolution of the Renia gneissic dome on the east flank of the Labrador Trough, Ungava Bay: Geological Association of Canada, Mineralogical Association of Canada, Canadian Geophysical Union, Joint Annual Meeting, v. 10, p. A20. (Renia dome, Canada) Gervais, F., Nadeau, L., and Malo, M., 2004, Migmatitic structures and solidstate diapirism in orthogneiss domes, Eastern Grenville Province, Canada, in Whitney, D.L., Teyssier, C., Hodges, K.V., and Siddoway, C.S., eds., Gneiss Domes and Orogeny: Geological Society of America Special Paper 380, this volume. (Jalobert, Watshishou domes, Quebec, Canada)
Hansen, V.L., and Goodge, J.W., 1988, Metamorphism, structural petrology, and regional evolution of the Okanogan Complex, northeastern Washington, in Ernst, W.G., ed., Metamorphism and crustal evolution of the Western United States: Englewood Cliffs, New Jersey, Prentice-Hall, Rubey volumes, v. 7, p. 233–270. (Okanogan dome, Washington, United States) Haslam, H.W., Brewer, M.S., Davis, A.E., and Darbyshire, D.P.F., 1980, Anatexis and high-grade metamorphism in the Champira Dome, Malawi; petrological and Rb-Sr studies: Mineralogical Magazine, v. 43, p. 701–714. (Champira dome, Malawi) Hatcher, R.D., 1984, Southern and Central Appalachian basement massifs, in Bartholomew, M.J., ed., The Grenville Event in the Appalachians and related topics: Boulder, Colorado, Geological Society of America Special Paper 94, p. 149–153. (Toxaway dome/South Carolina, North Carolina, United States) Hepburn, J.C., 1986, Geology of the Guilford Dome area, Brattleboro Quadrangle, southeastern Vermont: Vermont Geology, v. 4, p. B01–B09. (Guilford dome, Vermont, United States) Higgins, M.W., Fisher, G.W., and Zietz, I., 1973, Aeromagnetic discovery of a Baltimore Gneiss Dome in the Piedmont of Northwestern Delaware and Southeastern Pennsylvania: Geology, v. 1, p. 41–43. (Mill Creek dome, Delaware, United States) Hippertt, J.F., 1994, Structures indicative of helicoidal flow in a migmatitic diapir (Bacao Complex, southeastern Brazil): Tectonophysics, v. 234, p. 169–196, doi: 10.1016/0040-1951(94)90210-0. (Bacao dome, Brazil) Hodges, K.V., and Applegate, J.D., 1993, Age of Tertiary extension in the Bitterroot metamorphic core complex, Montana and Idaho: Geology, v. 21, p. 161–164, doi: 10.1130/0091-7613(1993)0212.3.CO;2. (Bitterroot dome, Montana, United States) Holm, D.K., and Lux, D.R., 1996, Core complex model proposed for gneiss dome development during collapse of the Paleoproterozoic Penokean orogen, Minnesota: Geology, v. 24, p. 343–346, doi: 10.1130/00917613(1996)0242.3.CO;2. (McGrath dome, Minnesota, United States) Jansen, J.B.H., and Schuiling, R.D., 1976, Metamorphism on Naxos; petrology and geothermal gradients: American Journal of Science, v. 276, p. 1225– 1253. (Naxos dome, Aegean, Greece) Jolivet, L., Maluski, H., Beyssac, O., Goffe, B., Lepvrier, C., Phan, T.T., and Nguyen, V.V., 1999, Oligocene-Miocene Bu Khang extensional gneiss dome in Vietnam; geodynamic implications: Geology, v. 27, p. 67–70, doi: 10.1130/0091-7613(1999)0272.3.CO;2. (Bu Khang, Dai Nui Con Voi domes, Vietnam) Ketin, I., 1947, Uludag masifinin tektonigi hakkinda; ueber die Tektonik des Uludag-Massivs: Turkiye Jeoloji Kurumu Bulteni, v. 1, p. 60–88. (Uludag Massif, Turkey) Konopasek, J., Schulmann, K., and Johan, V., 2002, Eclogite-facies metamorphism at the eastern margin of the Bohemian Massif; subduction prior to continental underthrusting?: European Journal of Mineralogy, v. 14, p. 701–713, doi: 10.1127/0935-1221/2002/0014-0701. (Svratka dome/ Bohemian Massif, Czech Republic) Kodama, K.P., and Chapin, D.A., 1984, A detailed gravity study of the Chattolanee Baltimore Gneiss Dome, Maryland, USA: Earth and Planetary Science Letters, v. 68, p. 286–296, doi: 10.1016/0012-821X(84)90160-2. (Chattolanee dome/Baltimore Gneiss, Maryland, United States) Kozhoukharov, D., 1980, The Prerhodopian and Rhodopian complex in the Shiroka Lake Anticline and the Lyaskovo Dome, in Kozhoukharov, D., and Dabovski, Ch., eds., The Precambrian in South Bulgaria: International Geological Correlation Programme Project No. 022, p. 50–62. (Lyaskovo dome, Bulgaria) Kumar, A., Mehta, Y.P., Lal, N., and Jain, A.K., 1994, Thermo-tectonic history of Himalayan domes and windows using fission track technique: Journal of Nepal Geological Society, v. 10, p. 77–78. (Suru dome, Himalaya, India) Kündig, R., 1989, Domal structures and high-grade metamorphism in the Higher Himalayan Crystalline, Zanskar region, Northwest Himalaya, India: Journal of Metamorphic Geology, v. 7, p. 43–55. (Zanskar domes, India) Lagarde, J.-L., Dallain, C., Ledru, P., and Courrioux, G., 1994, Strain pattern within the Variscan granite dome of Velay, French Massif Central: Journal of Structural Geology, v. 16, p. 839–852, doi: 10.1016/01918141(94)90149-X. (Velay dome, France) Langenberg, C.W., and Ramsden, J., 1980, The geometry of folds in granitoid rocks of northeastern Alberta: Tectonophysics, v. 66, p. 269–285, doi: 10.1016/0040-1951(80)90050-5. (Tulip Lake dome, Alberta, Canada)
Gneiss Domes and Crustal Flow Leake, B.E., 1978, Granite emplacement; the granites of Ireland and their origin, in Bowes, D.R., and Leake, E., eds., Crustal evolution in northwestern Britain and adjacent regions: Geological Journal, v. 10, p. 221–248. (Carna, Galway-Kilkieran dome) Ledru, P., Courrioux, G., Dallain, C., Lardeaux, J.M., Montel, J.M., Vanderhaeghe, O., and Vitel, G., 2001, The Velay Dome (French Massif Central); melt generation and granite emplacement during orogenic evolution: Tectonophysics, v. 342, p. 207–237, doi: 10.1016/S0040-1951(01)00165-2. (Velay dome, Massif Central, France) Ledru, P., N’Dong, J.E., Johan, V., Prian, J.-P., Coste, B., and Haccard, D., 1989, Structural and metamorphic evolution of the Gabon orogenic belt; collision tectonics in the lower Proterozoic?: Precambrian Research, v. 44, p. 227–241. (Abamie, Diany Miyole domes, Gabon) Leloup, P.H., Arnaud, N., Lacassin, R., Kienast, J.R., Harrison, T.M., Pan Trong, T.T., Replumaz, A., and Tapponnier, P., 2001, New constraints on the structure, thermochronology, and timing of the Ailao Shan-Red River shear zone, SE Asia: Journal of Geophysical Research, v. 106, p. 6683–6732, doi: 10.1029/2000JB900322. (Dai Nui Con Voi and domes of the Red River Shear Zone, SE Asia) Leslie, A.G., and Nutman, A.P., 2000, Episodic tectono-thermal activity in the southern part of the East Greenland Caledonides: Geology of Greenland Survey Bulletin, v. 186, p. 42–49. (Niggli Spids dome, Greenland) Mallick, D.I.J., 1967, The metamorphic development of the Mpande dome in Zambia: Geologische Rundschau, v. 56, p. 670–691. (Mpande dome, Zambia) Maluski, H., Lepvrier, C., Jolivet, L., Carter, A., Roques, D., Beyssac, O., Tang, T.T., Thang, N.D., and Avigad, D., 2001, Ar-Ar and fission-track ages in the Song Chay Massif; Early Triassic and Cenozoic tectonics in northern Vietnam: Journal of Asian Earth Sciences, v. 19, p. 233–248, doi: 10.1016/S1367-9120(00)00038-9. (Song Chay dome, Vietnam) Maluski, H., Rajlich, P., and Soucek, J., 1995, Pre-Variscan, Variscan and early Alpine thermo-tectonic history of the north-eastern Bohemian Massif; an 40Ar/39Ar study: Geologische Rundschau, v. 84, p. 345–358, doi: 10.1007/S005310050010. (Desna, Keprnik, Snieznik domes/Bohemian Massif, Poland) Marsh, P.D., 1984, The stratigraphy and structure of the Lagrange Nunataks, northern Fuchs Dome and Herbert Mountains of the Shackleton Range: British Antarctic Survey Bulletin, v. 63, p. 19–40. (Fuchs dome, Antarctica) Marshak, S., Tinkham, D., Alkmim, F., Brueckner, H., and Bornhorst, T.J., 1997, Dome-and-keel provinces formed during Paleoproterozoic orogenic collapse-core complexes, diapirs, or neither? Examples from the Quadrilatero Ferrifero and the Penokean Orogen: Geology, v. 25, p. 415–418, doi: 10.1130/0091-7613(1997)0252.3.CO;2. McDonough, M.R., Grover, T.W., McNicoll, V.J., Schetselaar, E.M., Cooley, M.A., Robinson, N.M., Van Ham, J.A., and Bednarski, J., 2000, Geology, Wylie Lake, Alberta-Saskatchewan: Ottawa, Ontario, Geological Survey of Canada Map 1951A, scale 1:50,000, 1 sheet. (Wylie Lake dome, Canada) McClellan, E.A., 1994, Contact relationships in the southeastern Trondheim nappe complex, central-southern Norway; implications for early Paleozoic tectonism in the Scandinavian Caledonides: Tectonophysics, v. 231, p. 85–111, doi: 10.1016/0040-1951(94)90123-6. (Einunnfjellet dome, Norway) McNeice, G.W., Boerner, D.E., Kurtz, R.D., and Jones, A.G., 1991, Gravity studies at the Lemieux Dome, Gaspe, Quebec: Calgary, Alberta, Geological Survey of Canada Open File Report 2291, 35 p. (Lemieux dome, Quebec, Canada) Menard, T., and Spear, F.S., 1994, Metamorphic P-T paths from calcic pelitic schists from the Strafford Dome, Vermont, United States: Journal of Metamorphic Geology, v. 12, p. 811–826. (Strafford dome, Vermont, United States) Mezger, J.E., and Passchier, C.W., 2003, Polymetamorphism and ductile deformation of staurolite-cordierite schist of the Bossòst dome: Indication for Variscan extension in the Axial Zone of the central Pyrenees: Geological Magazine, v. 140, p. 595–612, doi: 10.1017/S0016756803008112. (Bossòst dome, Spain/France) Miller, D.M., 1980, Structural study of the northern Albion Mountains, south-central Idaho: Geological Society of America, v. 153, p. 399–423. (Albion-Raft River-Grouse Creek complex, Idaho and Utah, United States)
17
Mlcoch, B., and Schulmann, K., 1992, Superposition of Variscan ductile shear deformation on pre-Variscan mantled gneiss structure (Catherine Dome, Erzgebirge, Bohemian Massif): Geologische Rundschau, v. 81, p. 501– 513. (St. Catherine Dome/Bohemian Massif, Czech Republic) Moecher, D.P., 1999, The distribution, style, and intensity of Alleghanian metamorphism in south-central New England; petrologic evidence from the Pelham and Willimantic domes: Journal of Geology, v. 107, p. 449–471, doi: 10.1086/314359. (Pelham dome, Massachusetts; Willimantic dome, Connecticut, United States) Moorbath, S., O’Nions, R.K., and Pankhurst, R.J., 1975, The evolution of early Precambrian crustal rocks at Isua, West Greenland; geochemical and isotopic evidence: Earth and Planetary Science Letters, v. 27, p. 229–239, doi: 10.1016/0012-821X(75)90034-5. (Amitsoq gneiss, Greenland) Muir, T.L., 1979, Discrimination between extrusive and intrusive Archean ultramafic rocks in the Shaw Dome area using selected major and trace elements: Canadian Journal of Earth Sciences, v. 16, p. 80–90. (Shaw dome, Ontario, Canada) Murphy, M.A., Yin, A., Kapp, P., Harrison, T.M., Manning, C.E., Ryerson, F.J., Ding, L., and Guo, J., 2002, Structural evolution of the Gurla Mandhata detachment system, Southwest Tibet; implications for the eastward extent of the Karakoram fault system: Geological Society of America Bulletin, v. 114, p. 428–447, doi: 10.1130/0016-7606(2002)1142.0.CO;2. (Karakoram, Tibet) Natal’in, B.A., Amato, J.M., Toro, J., and Wright, J.E., 1999, Paleozoic rocks of northern Chukotka Peninsula, Russian Far East; implications for the tectonics of the Arctic region: Tectonics, v. 18, p. 977–1003, doi: 10.1029/ 1999TC900044. (Koolen dome/Bering Strait region, Russia) Neubauer, F., Dallmeyer, R.D., Dunkl, I., and Schirnik, D., 1995, Late Cretaceous exhumation of the metamorphic Gleinalm Dome, Eastern Alps; kinematics, cooling history and sedimentary response in a sinistral wrench corridor: Tectonophysics, v. 242, p. 79–98, doi: 10.1016/00401951(94)00154-2. (Gleinalm dome, Austria) Okay, A.I., and Satır, M., 2000, Coeval plutonism and metamorphism in a latest Oligocene metamorphic core complex in Northwest Turkey: Geological Magazine, v. 137, p. 495–516, doi: 10.1017/S0016756800004532. (Kazdag dome, Turkey) Olsen, S.N., 1999, Petrology of the Baltimore Gneiss in the Northeast Towson Dome, Maryland Piedmont, in Valentino, D.W., and Gates, A.E., eds., The Mid-Atlantic Piedmont; tectonic missing link of the Appalachians: Boulder, Colorado, Geological Society of America Special Paper 330, p. 113–126. (Towson dome/Baltimore Gneiss, Maryland, United States) Percival, J.A., 1979, Kyanite-bearing rocks from the Hackett River area, N.W.T.; implications for Archean geothermal gradients: Contributions to Mineralogy and Petrology, v. 69, p. 177–184. (Hackett River dome, Northwest Territories, Canada) Pouget, P., 1987, Le massif granitique de Lesponne (Haut Pyrenees); un exemple de massif plutonique hercynien a mise en place diapirique syncinematique: Geologische Rundschau, v. 76, p. 187–199. (Lesponne dome, Pyrenees, France) Pouget, P., 1991, Hercynian tectonometamorphic evolution of the Bosost Dome (French-Spanish central Pyrenees): Journal of the Geological Society of London, v. 148, p. 299–314. (Bosost dome, associated with Garona/ Garonne dome, Pyrenees, France/Spain) Ratcliffe, N.M., 1990, Comparative tectonics of basement massifs in the Northern Appalachians with special reference to the Green Mountain Massif of Vermont: Vermont Geology, v. 6, p. 44. (Rayponda and Sadawga domes, Vermont, United States) Reche, J., Martinez, F.J., Arboleya, M.L., Dietsch, C., and Briggs, W.D., 1998, Evolution of a kyanite-bearing belt within a HT-LP orogen; the case of NW Variscan Iberia: Journal of Metamorphic Geology, v. 16, p. 379–394. (Lugo dome, Spain) Reddy, S.M., Cliff, R.A., and East, R., 1993, Thermal history of the Sonnblick Dome, Southeast Tauern Window, Austria; implications for heterogeneous uplift within the Pennine basement: Geologische Rundschau, v. 82, p. 667–675. (Sonnblick, Hochalm domes, Austria) Reesor, J.E., and Moore, J.M., 1971, Petrology and structure of Thor-Odin gneiss dome, Shuswap metamorphic complex, British Columbia: Bulletin of the Geological Survey of Canada, v. 195, 120 p. (Thor-Odin dome/ Shuswap Complex, British Columbia, Canada) Rhodes, B.P., and Cheney, E.S., 1981, Low-angle faulting and the origin of Kettle Dome, a metamorphic core complex in northeastern Washington: Geology, v. 9, p. 366–369. (Kettle dome, Washington, United States)
18
D.L. Whitney, C. Teyssier, and O. Vanderhaeghe
Ridley, J.R., Vearncombe, J.R., and Jelsma, H.A., 1997, Relations between greenstone belts and associated granitoids, in De Wit, M., and Ashwal, L.D., eds., Greenstone Belts: Oxford, UK, Oxford University Press, Monographs on Geology and Geophysics, v. 35, p. 376–397. (Zimbabwe, Africa) Robyr, M., and Steck, A., 2001, Thrusting, extension and doming during the polyphase tectono-metamorphic evolution of the High Himalayan Crystalline Sequence in the SE Zanskar and NW Lahul: Journal of Asian Earth Sciences, v. 19, no. 3A, p. 52–54. (Gianbul dome, Himalaya, Pakistan) Roig, J.-Y., and Faure, M., 1998, Folding and granite emplacement inferred from structural, strain, TEM and gravimetric analyses; the case study of the Tulle Antiform, SW French Massif Central: Journal of Structural Geology, v. 20, p. 1169–1189, doi: 10.1016/S0191-8141(98)00061-3. (Tulle antiform/dome, Massif Central, France) Rolland, Y., Mahéo, G., Guillot, S., and Pêcher, A., 2001, Tectono-metamorphic evolution of the Karakorum metamorphic complex (Dassu-Askole area, NE Pakistan); exhumation of mid-crustal HT-MP gneisses in a convergent context: Journal of Metamorphic Geology, v. 19, p. 717–737. (Dassu, Askole, Panhma, + unnamed domes, Himalaya, Pakistan) Saunier, J.F., 1986, Un domaine cristallophyllien dans le Proterozoique superieur (Brioverien) de Bretagne centrale; Le dome de Plouguenast (Cotes du Nord): Documents—Bureau de Recherches Géologiques et Minieres, v. 109, p. 139. (Plougenast dome, Brittany, France) Schuiling, R.D., 1960, Le dome gneissique de l’Agout (Tarn et Herault): Memoires de la Société Géologique de France, Nouvelle Serie, v. 39, no. 91, p. 59. (Agout dome, Massif Central, France) Schwerdtner, W.M., Sutcliffe, R.H., and Troeng, B., 1978, Patterns of total strain in the crestal region of immature diapirs: Canadian Journal of Earth Sciences, v. 15, p. 1437–1447. (Ash Bay, Law Lake domes, Ontario, Canada) Shanks, W.S., 1986, A Structural Analysis of the Cedar Lake Dome, Northwestern Ontario [M.Sc. thesis]: Toronto, Canada, University of Toronto. Silva, L.J.H. del Rey, 1995, The evolution of basement gneiss domes of the Sergipano fold belt (NE Brazil) and its importance for the analysis of Proterozoic basins: Journal of South American Earth Sciences, v. 8, p. 325–340. (Itabaiana, Simao Dias domes, Brazil) Simony, P.S., and Carr, S.D., 1997, Large lateral ramps in the Eocene Valkyr shear zone; extensional ductile faulting controlled by plutonism in southern British Columbia: Journal of Structural Geology, v. 19, p. 769–784, doi: 10.1016/S0191-8141(97)00011-4. (Valhalla/Passmore domes/ Shuswap Complex, British Columbia, Canada) Sims, P.K., and Schulz, K.J., 1982, Newly recognized gneiss dome, northeastern Wisconsin: Reston, Virginia, U.S: Geological Survey Professional Paper P1375, p. 53–54. (Dunbar dome, Wisconsin, United States) Smith, C.H., and Richard, S.M., 1991, A layered, amphibolite to granulite facies, migmatite sequence in the Fosdick metamorphic complex, West Antarctica: International Symposium on Antarctic Earth Sciences, v. 6, p. 537–538. (Fosdick Mountains, Antarctica) Solar, G.S., and Brown, M., 1995, Structural significance of contrasting styles of migmatization; evidence from the Tumbledown Dome area, westcentral Maine: Geological Society of America, Abstracts with Programs, v. 27, no. 6, p. 224. (Tumbledown dome, Maine, United States) Soula, J.-C., 1982, Characteristics and mode of emplacement of gneiss domes and plutonic domes in central-eastern Pyrenees: Journal of Structural Geology, v. 4, p. 313–342, doi: 10.1016/0191-8141(82)90017-7. (Pyrenees gneiss domes) Soula, J.-C., Debat, P., Deramond, J., Guchereau, J.-Y., Lamouroux, C., Pouget, P., and Roux, L., 1986, Evolution structurale des ensembles metamorphiques des gneiss et des granitoides dans les Pyrenees centrales: Bulletin de la Société Géologique de France, Huitieme Serie, v. 2, p. 79–93. (Chiroulet, Gavarnie-Heas-Barroude, Lesponne domes, Pyrenees, France) Spear, F.S., Kohn, M.J., Cheney, J.T., and Florence, F., 2002, Metamorphic, thermal, and tectonic evolution of central New England: Journal of Petrology, v. 43, p. 2097–2120, doi: 10.1093/PETROLOGY/43.11.2097. (Keene, Alstead domes, New Hampshire, United States) Steck, A., Epard, J.-L., Vannay, J.-C., Hunziker, J., Girard, M., Morard, A., and Robyr, M., 1998, Geological transect across the Tso Morari and Spiti areas; the nappe structures of the Tethys Himalaya: Eclogae Geologicae Helvetiae, v. 91, p. 103–122. (Tso Morari/Himalaya, Pakistan) Stipska, P., Schulmann, K., and Hoeck, V., 2000, Complex metamorphic zonation of the Thaya Dome; result of buckling and gravitational collapse of an imbricated nappe sequence, in Cosgrove, J.W., and Ameen, M.S., eds.,
Forced folds and fractures: London, Geological Society Special Publication 169, p. 197–211. (Thaya dome/Bohemian Massif, Czech Republic) Sturchio, N.C., Sultan, M., and Batiza, R., 1983, Geology and origin of Meatiq Dome, Egypt; a Precambrian metamorphic core complex?: Geology, v. 11, p. 72–76. (Meatiq dome—also spelled Maetiq—and Wadi Hafafit culmination, Egypt) Symons, D.T.A., and Harris, M.J., 2000, The approximately 1830 Ma TransHudson hairpin from paleomagnetism of the Wapisu gneiss dome, Kisseynew Domain, Manitoba: Canadian Journal of Earth Sciences, v. 37, p. 913–922, doi: 10.1139/CJES-37-6-913. (Wapisu dome, Manitoba, Canada) Talarico, F., and Kroner, U., 1999, Geology and tectono-metamorphic evolution of the Read Group, Shackleton Range; a part of the East Antarctic Craton, in Millar, I.L. and Talarico, F., eds., The EUROSHACK Project; results of the European expedition to the Shackleton Range, 1994/95: Terra Antarctica, v. 6, p. 183–202. (Read Mountains, Antarctica) Thompson, J.B., Robinson, P., Clifford, T.N., and Trask, N.J., 1968, Nappes and gneiss domes in west-central New England, in Zen, E-An, et al., eds., Studies of Appalachian Geology: Northern and maritime: New York, Interscience Publishers, p. 203–218. (Pelham, Warwick, Monson/ Tully, domes, Massachusetts; Mascoma, Croyden, Unity, Alstead, Keene domes, New Hampshire; Vernon dome, Vermont–New Hampshire, United States) Toth, T.M., Schubert, F., and Zachar, J., 2000, Neogene exhumation of the Variscan Szeghalom Dome, Pannonian Basin, E. Hungary: Geological Journal, v. 35, p. 265–284, doi: 10.1002/GJ.861. (Szeghlom dome, Hungary) Turner, D.C., 1983, Upper Proterozoic schist belts in the Nigerian sector of the Pan-African Province of West Africa: Precambrian Research, v. 21, p. 55–79, doi: 10.1016/0301-9268(83)90005-0. (Kushaka dome, Kibaran orogen, Nigeria) van den Driessche, J., and Brun, J.P., 1992, Tectonic evolution of the Montagne Noire (French Massif Central); a model of extensional gneiss dome: Geodinamica Acta, v. 5, p. 85–99. (Montagne Noire dome, Massif Central, France) van der Maar, P.A., and Jansen, J.B.H., 1983, The geology of the polymetamorphic complex of Ios, Cyclades, Greece and its significance for the Cycladic Massif: Geologische Rundschau, v. 72, p. 283–299. (Ios, Aegean, Greece) van Kranendonk, M.J., and Collins, W.J., 1998, Timing and tectonic significance of late Archaean, sinistral strike-slip deformation in the central Pilbara structural corridor, Pilbara Craton, Western Australia: Precambrian Research, v. 88, p. 207–231, doi: 10.1016/S0301-9268(97)00069-7. (Tambourah, Shaw, North Pole, McPhee domes, W Australia) van Staal, C.R., and Williams, P.F., 1983, Evolution of a Svecofennian mantled gneiss dome in SW Finland, with evidence for thrusting: Precambrian Research, v. 21, p. 101–128, doi: 10.1016/0301-9268(83)90007-4. (Bergon dome, Finland) Verhoef, P.N.W., Vissers, R.L.M., and Zwart, H.J., 1984, A new interpretation of the structural and metamorphic history of the western Aston massif (central Pyrenees, France): Geologie en Mijnbouw, v. 63, p. 399–410. (Aston dome, Pyrenees, France) Vyhnal, C.R., and Chamberlain, C.P., 1996, Preservation of early isotopic signatures during prograde metamorphism, eastern Vermont: American Journal of Science, v. 296, p. 394–419. (Athens, Pomfret domes, Vermont, United States) Wang, K., Windley, B.F., and Sills, J.D., 1988, Archaean gneiss complex of eastern Hebei Province: North China: Chemical Geology, v. 70, p. 149. (Qianax dome, N China) Whitney, D.L., and Dilek, Y., 1998, Metamorphism during Alpine crustal thickening and extension in central Anatolia, Turkey; the Nig˘de metamorphic core complex: Journal of Petrology, v. 39, p. 1385–1403, doi: 10.1093/ PETROLOGY/39.7.1385. (Nig˘de dome, Turkey) Wintsch, R.P., 1979, The Willimantic Fault; a ductile fault in eastern Connecticut: American Journal of Science, v. 279, p. 367–393. (Willimantic, Lyme domes, Connecticut, United States) Yan, D.-P., Zhou, M.-F., Song, H., and Fu, Z., 2003, Structural style and tectonic significance of the Jianglang Dome in the eastern margin of the Tibetan Plateau, China: Journal of Structural Geology, v. 25, p. 765–779, doi: 10.1016/S0191-8141(02)00059-7. (Jianglang, Motianling, Qiaoziding, Qiasi, Goncai, Yaside, Xunlongbao domes, East Tibetan Plateau, China) Zwart, H.J., 1978, Metamorphism in the Pyrenees, in Zwart, H.J., Sobolev, V.S., and Niggli, E., eds., Metamorphic map of Europe: Leiden and Paris, Sub-
Gneiss Domes and Crustal Flow
19
commission for the Cartography of the Metamorphic Belts of the World and the United Nations Educational, Scientific and Cultural Organization, scale 1:25,000, explanatory text, p. 104–108. (Agly, Bosost, Chiroulet domes, Pyrenees, France–Spain) MANUSCRIPT ACCEPTED BY THE SOCIETY APRIL 8, 2004
Printed in the USA