A Review of Mesoscopic Magmatic Structures and Their Potential For Evaluating The Hypersolidus Evolution

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Structural Geology 125 (2019) 134–147

Contents lists available at ScienceDirect

Journal of Structural Geology


journal homepage: www.elsevier.com/locate/jsg

A review of mesoscopic magmatic structures and their potential for T


evaluating the hypersolidus evolution of intrusive complexes
Scott R. Patersona,∗, Katie Ardilla, Ron Vernonb, Jiří Žákc
a
Department of Earth Sciences, University of Southern California, 3651 Trousdale Parkway, Los Angeles, CA 90089-0740, USA
b
Department of Earth and Planetary Sciences, Macquarie University, Sydney, NSW 2109, Australia
c
Institute of Geology and Paleontology, Faculty of Science, Charles University, Albertov 6, Prague, 12843, Czech Republic

A R T I C LE I N FO A B S T R A C T

Keywords: The perception that plutonic bodies are structurally simple is disappearing with the recognition of an array of
Fabric magmatic structures useful for constraining hypersolidus temporal histories, evolving rheologies, strain fields,
Granite flow directions, growth and cooling patterns, tilting and syn-emplacement tectonism. These histories provide a
Intrusive complex powerful means of testing an array of growth, emplacement, chamber evolution and tectonic models. Important
Magmatic structures
points include that: (1) many structures form in “hydrogranular” or congested magma slurries during magma
Crystal sorting
mush avalanching, local convection and late hypersolidus strain challenging the notion that magmas must have
≤55% crystals to convect/fractionate and form compositional diversity in upper crustal magma chambers; (2)
magmatic fabrics reflect transient strain in these slurries rather than flow directly: the latter must be inferred by
linking geometries with temporal and kinematic information; (3) caution is needed when using anisotropy of
magnetic susceptibility and similar quantitative tools that characterize, through a single ellipsoid, preferred
orientations of mineral grains given the increasing recognition of multiple fabrics in plutons; and (4) that future
studies are particularly needed in plutons focusing on the distribution and styles of compositionally defined
structures, magmatic folding, shear zones and faults, multiple fabrics and of the physical/chemical behaviors in
hydrogranular slurries.

1. Introduction Didier, 1973; Moore and Lockwood, 1973; Wiebe, 1974), but still
lagged behind the depth and understanding of structural work in me-
The earliest study of mesoscale structures in plutons is often traced tamorphic rocks, possibly because of the perception that many grani-
back to at least the seminal works by H. Cloos (1925), E. Cloos (1936, toids lacked mesoscale internal structures, and certainly because the
1946) and Balk (1937), who emphasized the value of studying mag- complex nature of the hypersolidus to subsolidus behaviors of plutons
matic foliation, lineation, layering and joints in plutons. Subsequently, (Fig. 1) inhibited a full understanding of plutonic structures. Great
the conspicuous layering, unconformities and erosional features, in- advances have been made over the last 40 years, as the 'granite com-
cluding 'cross-bedding' and magmatic shears in mafic–ultramafic in- munity' has meshed detailed field studies of igneous systems with
trusions attracted much attention (e.g., Wager and Brown, 1968; technologically-advanced structural, chemical and modeling tools.
Jackson, 1971; Irvine, 1980, 1987). These were explained by crystal Below we provide a survey of the rich diversity of mesoscale (visible
settling, marginal crystallization combined with rhythmic super- at outcrop scale, sometimes aided with hand lens) magmatic structures
saturation (e.g., McBirney and Noyes, 1979), and deposition from in plutons (Fig. 2) and discuss their value for unravelling hypersolidus
crystal-mush density currents. Similar features were noted in grani- histories of granitoids. To evaluate structures in plutons, one must first
toids. For example, Gilbert (1906) and E. Cloos (1936) described ac- distinguish magmatic versus solid state structures and then determine
cumulations of alkali-feldspar megacrysts in single-mineral aggregates whether they were formed by internal or tectonic processes (Fig. 3).
and modally-sorted layers of schlieren, commonly with graded-bedding After a brief introduction to these steps, we survey five groups
and cross-bedding. (Figs. 2–5) of magmatic structures: (1) internal contacts; (2) composi-
By 1977 (the initiation of Journal of Structural Geology), structural tionally defined structures; (3) preferred orientations; (4) deformation
analyses of granitoids were beginning to expand (e.g., King, 1964; structures; and (5) local indicators of growth, younging, vertical and
Wager and Brown, 1968; Pitcher and Berger, 1972; Pitcher et al., 1985; kinematics, followed by mention of outstanding challenges and a few


Corresponding author.
E-mail address: paterson@usc.edu (S.R. Paterson).

https://doi.org/10.1016/j.jsg.2018.04.022
Received 3 January 2018; Received in revised form 14 April 2018; Accepted 16 April 2018
Available online 12 May 2018
0191-8141/ © 2018 Elsevier Ltd. All rights reserved.
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

Fig. 1. Schematic diagram showing the complex transition from magmatic to solid state behavior, its dependence on magma crystallinity, strain rate, and changing
deformation mechanisms (color coded from lower left to upper right of diagram). Magmatic, plastic and brittle processes can occur simultaneously in the transitional
'submagmatic' zone, particularly at different strain rates. Common structures in each transitional zone are depicted as follows: (a) euhedral crystals with no preferred
orientation; (b) magmatic foliation; (c) magmatic shear; (d) brittle fracture of minerals; (e) mineral-mineral interactions forming synneusis or glomerocrysts; (f)
minerals with melt-filled fractures; (g) deformation twinning; (h) veins and fractures cross-cutting minerals; (i) chessboard subgrain formation in quartz; (j) de-
formation creep in phenocrysts. Three dashed lines show possible paths of tectonically driven magma deformation (slow strain rates), internal magmatic processes,
and processes during eruption (fast strain rates).

particularly useful modern approaches for addressing these challenges. filled pores just above the solidus (Arzi, 1978; Vigneresse, 2008; Costa
The equally rich world of microstructures in plutons (visible micro- et al., 2009; Brown, 2013; Bergantz et al., 2017). Thus, the magma's
scopically in thin sections) and of meso- and microstructures in the rheological state undergoes a continuous evolution (Vigneresse et al.,
coupled host rocks (e.g., Vernon, 2004; Passchier and Trouw, 1996) 1996; Scaillet et al., 2000; Petford, 2009) that is sometimes broadly
will not be emphasized, although all three should be integrated in divided into magmatic, submagmatic and solid-state fields (see below
practice. We conclude with interpretations on how magmatic structures and references in Paterson et al., 1989; Vernon, 2004).
form (Fig. 6) and their potential for unravelling hypersolidus growth The magmatic rheologic field involves displacement of magma with
histories of intrusive complexes. enough melt for objects (crystals, enclaves, rock fragments) to move
and rotate into alignment without significant interactions nor de-
2. Deciphering magmatic vs. solid-state and internal vs. tectonic forming internally. At the mesoscale, the most robust observations for
processes recognizing structures formed during magmatic flow are those asso-
ciated with: (1) igneous shapes (euhedral for selected crystals), internal
The first step in evaluating structures in plutons is to determine the crystal features (concentric zoning, growth twinning; e.g., Streck,
rheological state of the deforming material when the structures formed 2008), and crystal distributions (clustered and anticlustered; e.g.,
(Fig. 1). This is challenging because as the crystal-to-melt proportion Jerram et al., 2003); (2) plutonic structures of identical age that intrude
increases, magmas change from a dilute crystal suspension in melt, to or crosscut them; (3) growth of magmatic features (comb layering,
“hydrogranular” or congested magma slurries with a transient crystal orbicular structure; Moore and Lockwood, 1973); (4) evidence of high
framework and connected melt network, and finally to isolated melt- strain with no solid-state microstructures in crystals, implying that melt

135
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

Fig. 2. Photos of mesoscale magmatic structures (scale: rulers = 15 cm, hammer ∼22 cm, notebook ∼15 cm): (a) cross-cutting contacts and internal layering; (b)
graded schlieren; (c) different styles of layering including from right to left leucogranite dike, enclave swarm, schlieren, troughs, schlieren + felsic layers, “ridge and
pillar”; (d) migrating tube structure and reintrusion by host magma; (e) diapir; (f) cross-cutting schlieren troughs, locally folded (white arrow) or faulted (red arrow),
height of outcrop ∼5 m; (g) plume with schlieren-rich head and gradational tail, arrow = 2D movement direction; (h) mafic ellipsoid; (i) magmatic “dish and pillar”
structures due to upward melt migration; (j) feldspar UST structure; (k) folded plagioclase + amphibole magmatic foliation with new magmatic hornblende parallel
to axial plane; (l) folded aplite dike (thin dashed line) with magmatic axial planar foliation (thick line) defined by plagioclase, biotite, hornblende and enclaves; (m)
two magmatic preferred orientations at a single outcrop. Older enclave alignment = solid line. Younger plagioclase, hornblende, biotite alignment = dashed line.
Enclave tips bending from old to new; (n) magmatic mullion at margin of rhyolite plug intruding marine metasediments; (o) K-feldspar megacryst cluster; (p) pipe-
shaped K-feldspar megacryst cluster; (q) complex normal faulting, trough truncation, and folding of schlieren. Comb layering evident in some light-colored layers.

takes up the strain (Park and Means, 1996); and (5) contemporaneous transfer of melt into sites of low mean pressure (Paterson et al., 1989;
melt relocation structures (e.g., melt pockets in low stress sites; Hibbard Vigneresse et al., 1996; Vernon, 2004; Bergantz et al., 2017). Discrete-
and Watters, 1985). Microstructural criteria for magmatic flow are element numerical modeling by Bergantz et al. (2017) and Schleicher
summarized by Paterson et al. (1989) and Vernon (2000, 2004). et al. (2016) elegantly display the complex fluid to visco-plastic beha-
The submagmatic rheologic field involves flow/strain of hydro- vior of these flows, including the ephemeral development of 'crystal
granular magma slurries with transient crystal frameworks and melt force chains' (see also Philpotts and Dickson, 2000) that can transmit
networks by a complex combination of processes, such as melt-assisted stresses, drive intracrystalline deformation and change local melt
grain-boundary sliding, grain-boundary migration assisted by contact pressures. This field is the most challenging to recognize at the me-
melting, strain partitioning into melt-rich zones, melt-enhanced em- soscale, since it involves a mixture of both magmatic and solid-state
brittlement, high-temperature intracrystalline plastic deformation, and features across short spatial and temporal scales. Perhaps the best

136
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

Fig. 3. Cartoon showing three 'endmember' preferred orientation patterns in plutons (pink) and surrounding host rock modified from Paterson et al. (1998). (a)
Completely decoupled system (internally complex); (b) partially coupled system (margin parallel); (c) completely coupled system (rectilinear or folded). Black
lines = magmatic foliation, lineation, or folds. Grey lines = host rock structures. Filled dots = vertically plunging lineations; arrows = plunging lineations.

criteria are the evidence of patches of crystal plasticity or elastic phe- 1989; Johnson et al., 2003).
nomena (subgrains, faulting) contemporaneous with melt relocation Complications can occur if regional deformation destabilizes
phenomena. For example, the presence of primary mineral aggregates growing crystal mush zones in chambers, which then collapse and form
(namely quartz, alkali-feldspar, and sodic plagioclase) in con- local magmatic structures, such as the magmatic folds, faults and
temporaneous fractures (Bouchez et al., 1992; Vernon, 2004), indicates troughs. We know of no reliable criteria to easily identify such cases as
that late-stage melt travelled to low-pressure sites during brittle de- being of tectonic origin beyond the permissive observation of being
formation of crystals. spatially linked to a syn-intrusive fault (Emelus and Troll, 2014).
At the mesoscale, solid-state deformation is characterized by: (1)
evidence of grain-shape change and grainsize reduction by internal 3. Inventory of magmatic structures
deformation and recrystallization; (2) lenticular recrystallized ag-
gregates and 'tails' of minerals; (3) structures passing through solid Space restrictions only allow a brief introduction below of 5 pro-
xenoliths and/or continuous with structures in solid host rock; (4) fine- minent groups of structures that are the most common and most useful
grained folia anastomosing around less deformed 'relic' aggregates; (5) for unravelling hypersolidus granitoid histories (Fig. 2).
boudinage of strong minerals or clasts, with recrystallized aggregates of
weaker minerals (e.g., quartz and mica) between the boudins; (6) het- 3.1. Structures defining regional internal contacts and layering
erogeneous strain with local mylonitic zones; and (7) fractures and
cataclasites. Microstructural criteria for solid-state flow are summarized Incremental emplacement of compositionally-distinct magma pulses
by Paterson et al. (1989), Passchier and Trouw (1996) and Vernon in intrusive complexes results in a range of magmatic structures, such as
(2000, 2004). dikes and gradational to sharp unit contacts (Fig. 2a, b, c, o; Pitcher and
Once hypersolidus structures are identified and characterized, Berger, 1972; Hardee, 1982; Paterson et al., 2011). Flow, mixing and
consideration must be given to whether they reflect localized internal mingling along these contacts commonly leads to hybrid zones and
processes within an active magma chamber or reflect regional tectonic layered intrusive sequences (e.g., Fernandez and Gasquet, 1994; Wiebe
deformation imposed on a magma chamber. Magmatic and submag- and Collins, 1998; Barbey, 2009).
matic structures can form by either process, since regional deformation Internal contacts are also defined by a wide range of igneous
can act on a crystallizing magma chamber. We view 'magma chambers' layering, particularly schlieren, in granitoids, ranging from kilometer-
as structural geologists, that is not just eruptible magma, but any mush scale layered zones (Fig. 2b; Lucus and St-Onge, 1995; Clarke and
that contains enough interconnected melt to allow magmatic to sub- Clarke, 1998; Žák et al., 2009; Burgess and Miller, 2008) in which
magmatic structures, including late fabrics to form. Two reliable cri- layering maintains subparallel orientations, mineral grading and local
teria for recognizing tectonically-formed hypersolidus structures in- cross-bedding, to more complexly-shaped, 10 cm-to 10 m-scale struc-
clude: (1) coupling or continuity with regional tectonic structures in tures discussed below (Fig. 2d, e, f, q; e.g., Barrière, 1981; Parsons,
host rock (Fig. 3c); and (2) continuous structures crosscutting or 1986; Paterson et al., 2008; Pinotti et al., 2016). Mafic–ultramafic
overprinting internal contacts and/or xenoliths, indicating that magma complexes also are characterized by both internal intrusive contacts,
pulse juxtaposition predated the structure (e.g., Paterson et al., 1998). compositional layering of cumulates, and local 'unconformable' struc-
Pervasive solid-state ductile structures in plutons are typically caused tures resembling cross-bedding (Fig. 2; e.g., Wager and Brown, 1968;
by regional deformation due to the mechanical challenge of lower Irvine, 1980; Naslund and McBirney, 1996). These layers can be mod-
viscosity magmas deforming higher viscosity solid rock (e.g., Ramsay, ally, or grain and density sorted, and have graded or sharp contacts.

137
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

Fig. 4. Key elements used to define 'compositionally defined' structures (modified from Paterson, 2009). (a) Schlieren-bound troughs showing truncations, mineral
grading, local mineral clusters (Photos Fig. 2f, p); (b) stationary tubes; (c) migrating tubes (Fig. 2d); (e) mesoscale diapir (Photo Fig. 2e); (f) plumes with schlieren
'heads' and open 'tails' (Fig. 2g); (g) mafic ellipsoids with internal margin parallel foliation (Fig. 2h); (h) irregular mineral clusters; (i) pipes.

138
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

Fig. 5. Cartoon of local, mesoscale structures used to infer growth and younging directions, paleovertical, and kinematics of magma movement (the latter modified
from Philpotts and Asher, 1994): (a) cross-cutting, schlieren-bound troughs and mineral grading (Photos Fig. 2c, f, p); (b) asymmetric (chilled versus hybrid) contacts
after Wiebe and Collins (1998). Enclaves are flattened near chilled margin and irregular near hybrid; (c) undeformed tube or pipe axes; (d) concave up enclave
channels with enclaves molding minerals at channel base after Wiebe and Collins (1998); (e) magmatic flame structures and load casts; (f) layering deflected by
sinking block; (g) (g) granophyre segregations attached to phenocrysts; (h) imbricate minerals; (i) crystal orientation patterns due to flow in dike; (j) broken and
sheared phenocrysts; (k) granophyre wisps and back-folding by differential flow; (l) granophyre wisps; (m) magmatic S–C structures; (n) ramp structures; (o) shear
bands; (p) vesicles elongated parallel to the axial surface of folded flow layers; (q) asymmetric micro-folds in flow layering formed during rotation of a phenocryst or
glomerocryst; (r) Riedel shears; (s) asymmetric folds of magmatic foliation or layering.

139
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

Fig. 6. Cartoon depicting mesoscale mag-


matic structure formation in an in-
crementally grown (purple units) crystal-
lizing magma chamber intruded by new
magma pulses (red and orange); modified
from Paterson et al. (2016). Crystal-rich
mush zones form and collapse along mar-
gins of older pulses. Arriving new pulses
drive convection and 'return flow' (grey ar-
rows) in chamber. Examples of how in-
dividual structures may form in this system:
(a) cross-cutting troughs during margin
collapse; (b) migrating tubes during con-
vection; (c) diapirs and plumes (d) moving
sideways as new pulse enters chamber; (e)
mafic ellipsoids and cognate inclusions
formed from collapsing or mingling margin
material with new pulse; (f) mineral clus-
ters during flow sorting; (g) pipe structures
by flow sorting during rising or sinking of
magmas; (h) magmatic faults cooling/
shrinking and collapsing margins; (i) dikes
and veins as late melts migrate through
marginal cracks.

Some have non-planar geometries, akin to schlieren-defined structures channels in 3D, with channels either nested inwards (stationary
in silicic plutons. tubes; Fig. 4b) or forming dike-like zones of crescent-shaped,
crosscutting schlieren (migrating tubes or ladder dikes; Fig. 4c; Reid
3.2. Localized compositionally defined structures et al., 1993; Weinberg et al., 2001; Žák and Klomínský, 2007;
Paterson, 2009; Dietl et al., 2010). Schlieren truncations provide
An increasingly diverse collection of compositionally defined, local younging or migration directions of tube margins. Some tube
magmatic structures are recognized in granitoids (Figs. 2 and 4). Many structures are re-intruded, rotated and broken during and after
of these are partly characterized by schlieren, that is, centimeter-to formation (e.g., Paterson, 2009; Hodge et al., 2012). As with trough
meter-scale layers typically with sharp basal contacts, diffuse upper structures, the younging direction of tubes may show systematic
contacts and downward modal increases of dense major and accessory patterns across plutons (Memeti et al., 2014; Ardill et al., 2017;
minerals (e.g., Barrière, 1981; Reid et al., 1993; Burgess and Miller, Wiebe et al., 2017).
2008; Paterson, 2009). Minerals within schlieren layers are strongly (3) Mesoscale diapirs (Figs. 2e and 4d) are 10 cm-to 10 m-scale, ellip-
aligned parallel to the basal schlieren orientation (Fig. 2). Schlieren soidal to tear-drop shaped, compositionally variable and distinct
form in widely ranging orientations, including vertical (Wiebe et al., from the host magma, and with narrow tails and schlieren-bound
2007; Žák and Klomínský, 2007; Paterson, 2009), indicating that hy- 'heads'. They have been described in both mafic (Larsen and Brooks,
drogranular flow sorting of crystals, and not just gravity, plays an im- 1994) and felsic (Paterson, 2009) plutons. The schlieren at the
portant role in schlieren formation. diapir head is thought to form by filter pressing and accumulation
Schlieren-bound structures include the following: of mafic and accessory minerals (Weinberg et al., 2001). The trend
of a line through the tail and symmetrically dividing the head
(1) Troughs (Figs. 2f, q and 4a) are open-ended, schlieren-bound provides a 2D movement direction. Where examined in detail,
channels comprised of multiple stacked and cross-cutting layers diapir 'movement directions' are found in all orientations and so are
(e.g., Wager and Brown, 1968; McBirney and Noyes, 1979; Barrière, not simply buoyancy driven (Weinberg et al., 2001; Paterson,
1981; Reid et al., 1993; Burgess and Miller, 2008). The perpendi- 2009).
cular direction to a layer truncation defines a younging or trough (4) Plume structures (Figs. 2g and 4e) have modally distinct (from host
growth direction (Fig. 2). magma), schlieren-bound 'heads' and broad tails, sometimes mi-
(2) Tubes (Figs. 2d and 4b, c) are closed, cylindrical, schlieren-bound micking mushroom or jellyfish shapes, that compositionally grade

140
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

into the host granitoid. Their broad, gradational tails make them observation of these relationships has the power to document complex
distinct from diapirs and indicate that the source of plume material strain histories in magma chambers (e.g., Fernandez, 1988; Žák et al.,
is the host magma. The trend of a line symmetrically splitting the 2007; Cao et al., 2015).
tail and the head provides a 2D movement direction, which in cases
where measured also have variable orientations. 3.4. Deformation-related structures
(5) Mafic ellipsoids (Figs. 2h and 4f) are 1–3 m long, elongate 3D el-
lipsoids of fine-grained, more mafic plutonic rock distinct from Once formed, all magmatic structures can be altered by continued
their host, with a weak schlieren border (Memeti et al., 2014). They deformation, resulting in transposition of these older structures during
can resemble microgranitoid enclaves, except for the schlieren rims formation of a variety of magmatic folds, faults and shear zones (Fig. 2).
and because mineral alignment within the mafic ellipsoids usually Both brittle and ductile faults can develop in magma (Hibbard and
defines a well-developed margin-parallel pattern. Watters, 1985; Dingwell, 1997; Katz et al., 2006; Bergantz et al., 2017).
(6) Mineral clusters (Fig. 2o and p, 4g) range from millimeter-scale Magmatic brittle faults and 'ductile' shears typically do not show high
joined crystals (synneusis of Vance, 1969), centimeter-scale crystal strain, elongate crystal shapes and strong crystal preferred orientations
clusters and glomerocrysts (Seaman, 2000; Cashman et al., 2017), (Fig. 2f, q), since most displacement is accommodated by melt (Park
to decimeter-scale clusters with highly varied shapes, from irre- and Means, 1996; Paterson et al., 1998). Faults also may draw in ad-
gular, dike-shaped, trough-shaped, to pipe-like, (Paterson et al., ditional melt (Geshi, 2001; Pinotti et al., 2016), become sealed by
2005; Memeti et al., 2014). Mesoscale clusters are dominated by a continued crystal growth (Park and Means, 1996) and therefore may be
single mineral, but may include microgranitoid enclaves and xe- cryptic, unless offset markers are present.
noliths, and may or may not have bounding schlieren. Mineral Although magmatic folds are briefly described in several publica-
preferred orientations in clusters are highly variable. tions (Brown et al., 1986; Harm, 1991; Smith and Houston, 1994;
(7) Pipes (Figs. 2o and 4h) are closed, cylindrical, often sub-vertical 3D Paterson et al., 1998), detailed studies of magmatic folds are rare. Ex-
structures with little internal layering. Pipe compositions are typi- amples of folded layering, dikes, enclaves, and magmatic foliations are
cally dominated by more evolved (buoyant) compositions (Wiebe common in both volcanic (e.g., Vernon, 1987; Iezzi and Ventura, 2000;
and Collins, 1998) or by one mineral and thus a type of mineral Castro et el. 2002; Ventura, 2004) and plutonic rocks (e.g., Rosenberg
cluster (e.g., Paterson, 2009). Their margins with the host magma et al., 1995; Paterson et al., 1998; Yoshinobu et al., 2009). Variable fold
may or may not be marked with weak schlieren. Internal mineral profiles, fold-related geometric and mineral lineations, presence or
orientation occurs parallel to pipe margins, although these folia- absence of an axial planar foliation and refolding all occur (Fig. 2f, k, i,
tions may rotate to 'regional' fabric orientations in pipe centers. q).
Rocher et al. (in press) describe meter-scale, vertically elongated K- Mullions, another response to strain (Fletcher, 1982; Price and
feldspar accumulations with inward-dipping concentric foliation Cosgrove, 1990; Urai et al., 2001; Schmalholz and Schmid, 2012), also
and steeply plunging lineations in their upper parts that change to occur in magmatic rocks (Paterson et al., 1998; Yoshinobu et al., 2009),
flat foliation with enclaves at their lower ends. and would benefit from further detailed studies. Fig. 2n shows a mul-
lion formed at the contact between an Ordovician rhyolite plug and
3.3. Structures defined by preferred orientation marine metasediments in the Chaschuil region, NW Argentina (Lusk
et al., 2017). Magmatic mullions have been described along dike-host
Strain of crystal mushes (Benn, 1994) caused by gradients during rock margins and in some cases the tip of the mullion can tighten and
magma flow or tectonic deformation, lead to the alignment of crystals, feed veins intruding subparallel to the axial planes of host rock folds
microgranitoid enclaves and xenoliths (Fig. 2k, l, m), as well as trans- (Paterson et al., 1998; Yoshinobu et al., 2009). Mullion axes are parallel
position of older magmatic structures during folding and faulting, to regional fold axes and the shortening directions implied by the
leading to the development of one or more magmatic foliations and mullions are consistent with regional shortening implied by the host
lineations (Fig. 3). A common misconception is that only a single fabric foliation. These observations document syn-magmatic regional de-
(foliation ± lineation) forms in intrusive complexes, whereas close formation during which the magma mush had a higher effective visc-
inspection often establishes the presence of two or more sets of fabrics osity than its host.
that may have contemporaneous or overprinting relationships (e.g., Žák
et al., 2007; Kratinová et al., 2010; Macchioli Grande et al., 2015; 3.5. Magmatic indicators of local solidification, younging, paleovertical,
Alasino et al., 2017). Different mineral populations may show different and kinematics
intensities of alignment and/or different orientations (Blumenfeld and
Bouchez, 1988; Schulmann et al., 1997; Paterson et al., 1998). Fur- A variety of local magmatic structures provide additional informa-
thermore, microgranitoid enclaves may have internal mineral align- tion about structures, such as their initial orientation, growth/crystal-
ments parallel to, or distinct from, alignments in the host matrix and lization direction and kinematics of associated deformation.
parallel to, or distinct from, the enclave long axis (Paterson et al., Unidirectional solidification (or crystallization) structures (Fig. 2j)
2004). Xenoliths may preserve older magmatic or metamorphic struc- (UST's of Shannon et al., 1982) include: (1) comb layering; (2) crenulate
tures. Local layering or other magmatic structures discussed above may layering; (3) dentritic growth patterns, similar to crescumulate and
have mineral alignments distinct from preferred orientations in nearby spinifex microstructures; and (4) intergrowth layering. Besides cross-
host granitoids (Paterson, 2009; Alasino et al., 2017). cutting intrusive boundaries (Fig. 2a), mesoscale magmatic structures
Magmatic fabric patterns may or may not show refraction across available to provide local growth directions include: (1) local cross-
compositional boundaries (Compton, 1955; Hines et al., in press), be cutting boundaries in compositionally defined structures (Fig. 2d, f, q,
parallel to fold axial planes and fold axes (Pitcher and Berger, 1972) 5a); (2) mineral grading (Fig. 2b, f, 5a); (3) magmatic load cast and
and show deflections in magmatic shear zones. Regional patterns can be flame structures (Fig. 5e; Puziewicz and Wojewoda, 1984); (4) asym-
quite varied, although Paterson et al. (1998) defined three simplified metric contact characteristics (Figs. 2b and 5b), such as the chilled
'endmember' patterns: (1) internally complex, decoupled (Fig. 3a); (2) versus hybrid margins of Wiebe and Collins (1998); and (5) enclave
margin parallel or partially coupled (Fig. 3b); and (3) rectilinear, in- channels (Figs. 2c and 5d) with mineral molding at base (Wiebe and
ternally folded and coupled with host rock (Fig. 3c. Type #1 and 2 Collins, 1998). Proposed 'way-up' or 'paleo-vertical' indicators include:
foliation patterns can locally show steepening lineation patterns, which (1) load casts, flame structures (Puziewicz and Wojewoda, 1984;
have been used to infer 'magma feeder zones' (e.g., Vigneresse and Vaughan et al., 1995); (2) block sinking patterns in non-convecting
Bouchez, 1997; Olivier et al., 1997; Talbot et al., 2004). Careful magma (Fig. 5f; Paterson et al., 1998; Emeleus and Troll, 2014); (3)

141
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

axes of undeformed pipes and tubes (Fig. 2d, p, 5c; Wiebe and Collins, (2016) show the complex simultaneous, viscous–plastic–brittle beha-
1998; Paterson, 2009), and (4) 'dish and pillar' structures during which vior of deforming hydrogranular crystal mushes and argue against
rising melts deflect layering upwards (Fig. 2i). Kinematics caused by abrupt rheological transitions. Therefore, a key issue is to recognize
deviatoric strain during movement of magma, can be established from: that crystal mushes can continue to deform at low melt proportions
(1) crystal orientation patterns defining S–C structures, shear bands, during which melt-aided, crystal-plastic, and even brittle deformation
and hyperbolic patterns in dikes (Fig. 5i, m, n, o; Komar, 1972; Shelley, mechanisms increasingly overlap (Bergantz et al., 2017; Holness et al.,
1985; Vernon, 1987; Miller and Paterson, 1994; Correa-Gomes et al., 2017).
2001; Geoffroy et al., 2002); (2) imbricate phenocrysts (Fig. 5h;
Blumenfeld and Bouchez, 1988); (3) broken, sheared and rotated phe- 4.2.2. Layering
nocrysts (Fig. 5g, q) (Vernon, 1987; Blumenfeld and Bouchez, 1988); Attempts to categorize layering (e.g., Wager and Brown, 1968;
(4) granophyre wisps (Fig. 5i; Philpotts and Asher, 1994); (5) asym- Irvine, 1980; Barbey, 2009; Pinotti et al., 2016) collectively suggest
metric folds (Fig. 5s); (6) segregations attached to phenocrysts (Fig. 5g); that the following styles form by processes that often occur simulta-
(7) Riedel shears (Fig. 5r); and (8) ramp structures (Fig. 5n; Philpotts neously in growing plutons: (1) intrusion of newly arriving magma
and Asher, 1994). pulses or of locally derived magma from other parts of the chamber
(e.g., late melts forming dikes) resulting in internal contacts and/or
4. Discussion planar zones of mingling and mixing (e.g., layering in hybrid zones
along contacts; Fig. 2a; Hutton, 1992; Paterson et al., 2008; Barbey,
4.1. Hypersolidus structural histories 2009); (2) fractionation driven by hydrodynamic and gravity-driven
sorting and crystal–melt separation resulting in modally defined
The wealth of magmatic structures described above emphasizes layering (Fig. 2b; Irvine, 1980; Barrière, 1981). Petford (2009) also
that, comparable to metamorphic structures, a growing set of structural noted that igneous layering may form during increasing shear rate and
'tools' exist for unravelling hypersolidus histories. For example, cross- shear thinning in congested magma slurries when crystals arrange
cutting or overprinting structures help to establish temporal histories. themselves into layers parallel with the mean flow direction, a phe-
Layering, foliation patterns, folds, and faults all provide information nomenon observed both experimentally and in computer simulations
about hypersolidus strain fields. And when combined with kinematic (Stickel and Powell, 2005); (3) thermally/chemically driven fractiona-
information they can be used to infer magma flow directions. Structures tion and flow forming diffuse layering (Boudreau, 1995); (4) de-
indicating paleovertical in chambers can be used to evaluate local formation-assisted (magmatic or tectonic) strain (e.g., attenuated en-
magmatic tilting of layers or tectonic tilting of entire intrusive com- claves forming schlieren), crystal-melt segregation, or collection of
plexes. There also is great potential for constraining changing rheolo- crystals and rock fragments into planar zones (Fig. 2c, o, q; Tobisch
gies and evolving internal gradients as we better understand structures, et al., 1997); and (5) chemical and textural mineral modifications of
particularly the non-vertical motions of structures like diapirs and evolving layers as bulk compositions and environmental conditions
plumes and the mechanisms and kinematics of magmatic folding and shift (Boudreau, 1995).
faulting. The increasing mesoscale evidence of structures younging to- Only the first of the processes listed above to form internal contacts/
wards contacts with older units (unit ages established through geo- layers involves the arrival of a new pulse of magma. A challenge for
chronology) and statistically preferred migration directions are likely futures studies is to apply criteria for establishing which contacts reflect
driven by internal motions of magma during incremental growth and juxtaposition of new pulses versus internally formed boundaries (e.g.,
crystallization of intrusive complexes. Finally, in our experience, all Miller and Paterson, 2001). Thus two critical issues regarding internal
plutons have at least limited magmatic structures (e.g., Bouchez, 1997), contacts/layering are the following: (1) that many internal contacts
although some intrusive complexes are much richer in magmatic form by processes in already emplaced magma and are not contacts
structures than others (e.g., Weinberg et al., 2001; Pinotti et al., 2016). between arriving pulses; and (2) the degree to which internal contacts/
Why this is so will provide exciting information about the construction layering can be removed.
and behavior of these systems. Thus, the results of magmatic structural Some recent publications examining incremental growth of in-
studies provide powerful constraints on models of magmatic systems trusive complexes conclude that evidence for the numerous original
and need to be fully incorporated into geochemically dominated dis- intrusive contacts have been removed by subsequent processes (e.g.,
cussions of these systems. Coleman et al., 2004, 2012). To test this idea, it is important to examine
the processes proposed to remove contacts. Older internal contacts/
4.2. Key issues about magmatic structures layering may be removed by: (1) being entirely transported out of the
map section or recycled by a newly arriving magma pulse; (2) wide-
4.2.1. Hypersolidus deformation mechanisms spread convection and homogenization in a magma chamber; (3) re-
One unique aspect of magmatic structures is that they form across a working of contacts during widespread subsolidus deformation and
wide range of temperatures, effective viscosities and deformation me- metamorphism; and (4) pervasive annealing during widespread near-
chanisms associated with evolving physical and chemical processes solidus, thermally driven (possibly fluid aided) recrystallization.
(e.g., Cruden, 1990; Rosenberg and Handy, 2005). It has been tradi- Evidence of magmatic erosion resulting in sharp truncations of older
tional to identify key, rather abrupt, transitions in these behaviors (e.g., structures along new intrusive contacts, the presence of cognate blocks
van der Molen and Paterson, 1979; Vigneresse et al., 1996; Rosenberg of older phases in younger, and the presence of antecrysts in younger,
and Handy, 2005), and to conclude that magma with less than ca 55% particularly compositionally uniform units support #1 and #2
melt becomes 'static', which we believe typically means to others that (Paterson et al., 2016; Gaschnig et al., 2017). Widespread convection
magma is hard to erupt or vigorously convect. However, many of the (#2) is supported by the extensive dispersal of foreign objects (zircon
preserved structures noted in this paper likely formed during this antecrysts, microgranitoid enclaves, chemically distinct mineral popu-
transition from 55% to < 20%. Nicolas and Ildefonse (1996), Park and lations) throughout new pulses and homogenization of the originally
Means (1996) and Paterson et al. (1998) argue that magma can deform distinct pulses. The above mechanisms require a large, active magma
with even less than 20% melt through grain alignment and sliding on chamber.
melt films aided by dissolution or melting at grain impingements Removal of magmatic structures without an active chamber can
(called contact melting by Park and Means, 1996). Benn (1994) es- occur by #3 or #4. Extensive deformation and recrystallization at
tablished that only small increments of strain are needed to form amphibolite facies conditions (#3) results in easily identifiable gneisses
magmatic fabrics. And Bergantz et al. (2017) and Schleicher et al. with strong metamorphic fabric and/or compositional banding, and loss

142
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

of igneous microstructures, such as oscillatory zoning. Grains are (Paterson et al., 2016), and avalanching of crystal piles down into re-
characterized by polygonal to irregularly xenoblastic quartz and feld- gions of the chamber with lower effective viscosities (e.g., Irvine, et al.,
spar, and rounded inclusions of quartz in feldspar and vice versa 1998; Bergantz, 2000; Žák and Paterson, 2009).
(Vernon, 2000, 2004). Mesoscale diapirs and plumes show all the characteristics of
Conceivably, contacts could be obscured by #4. Grain boundary Raleigh–Taylor instabilities: but one surprising recent observation has
adjustments and mineral reactions are recognized in mafic/ultramafic been the wide variation in movement directions of these structures,
magmas where temperatures are significantly higher (Boudreau, 1995; many of which show non-vertical directions (Paterson, 2009). Alter-
Holness et al., 2017). If so, these processes should remove mineral natively, they may record local flow gradients/displacement fields
primary igneous microstructures such as oscillatory zoning, either by (stress, strain, rheology) in crystal mush systems, potentially driven by
grain-boundary migration or advance of a diffusion front. Although either regional deformation of magma chambers and/or convective
limited mineral alteration has been documented in granitoids (Barnes movement in reservoirs during arrival of new magma batches.
et al., 2017), considerable doubt exits that modifications extensive Mafic ellipsoids have only recently been recognized as such (Memeti
enough to remove evidence of older structures occurs without wide- et al., 2014) but may be comparable to microgranitoid enclaves and
spread subsolidus deformation (Vernon and Paterson, 2008b; Holness largely formed by magma mingling. The greatest uncertainties are the
et al., 2017). Furthermore, the common preservation of euhedral grain original source of magmas, the type of structure prior to mingling
shapes, euhedral inclusions, disequilibrium dihedral angles, and oscil- (mafic dikes, pipes, sheets) and where initial mingling occurred (e.g.,
latory zoning in the minerals in granitoids lacking solid-state de- Barbey et al., 2008; Paterson et al., 2016).
formation is strong evidence that extensive subsolidus grain-shape Although individual types have been examined, the full range of
changes did not occur during cooling of granitic magmas (Vernon and sizes and shapes of mineral clusters have not been systematically stu-
Paterson, 2008b; Holness et al., 2017). Finally, in typical granitoid died. Many are thought to have formed by the combined physical
plutons, internal structures, formed during both the construction of the processes of nucleation and surface energy effects, flow sorting, and
plutons and subsequent internal processes, are typically well preserved filter pressing (e.g., Vance, 1969; Barrière, 1981; Jerram et al., 2003;
(Fig. 2). This is true even for the granitoids emplaced at amphibolite Vernon and Paterson, 2008a). This clustering also implies that stress
facies conditions. Consequently, magmatic structures are unlikely to be transfers between crystals, leading to crystal plasticity, may be more
obliterated by intergranular changes, unless pervasive subsolidus tec- common in these hypersolidus systems than previously recognized.
tonic deformation occurs. Petford (2009) and Bergantz et al. (2017) summarized processes
such as the following that occur in flowing hydrogranular magma
4.2.3. Compositionally defined magmatic structures slurries because of their complex times-transgressive rheologic beha-
It is our expectation that knowledge of compositionally defined vior: (1) crystal disorder-order transition to form compositional
structures will continue to grow, both in the recognition of new types layering parallel to the flow direction; (2) either crystal jamming
and in a better understanding of formation. Mechanisms of formation (forming clumps) or dilatancy (drawing in melts); and (3) transient
presently remain controversial (e.g., compare Weinberg et al., 2001; particle pressures that fluidize crystal mushes resulting in an in-
Paterson, 2009; Hodge et al., 2012; and Wiebe et al., 2017). Although stantaneous pressure decrease and in situ bubble formation in the melt,
crystallization, fractionation, and mixing certainly play a role in promoting chamber-wide instability and formation of melt pipes and
forming crystals or the melts from which crystals grow, we suggest that channels, disrupted layering, crystal clustering, synmagmatic faulting,
existing evidence favors that physical processes dominate in the for- and collapse and remixing of mush zones. These phenomena could form
mation of these structures: below we present some of the more popular a number of the magmatic structures listed above.
ideas for their formation.
Wiebe and Collins (1998), Weinberg et al. (2001), Paterson (2009) 4.2.4. Preferred orientations of mineral grains
suggested that tubes and pipes represent cylindrical channels of magma Two outstanding issues regarding the interpretation of preferred
flow through crystal mushes, and that the main flow direction is par- orientations are the following: (1) the need to recognize that they re-
allel to the tube walls. Schlieren along the tube walls form by combined flect transient hydrogranular strain and are only parallel to flow planes/
flow sorting and filter pressing. Weinberg et al. (2001) also noted that directions under certain conditions; and (2) that there are many plutons
coarse K-feldspar aggregates occurred on the upstream side of 'dike with more than one foliation/lineation. The first issue is discussed ex-
necks' (pipes), as inferred from crosscutting relationships in the dike. tensively by Paterson et al. (1998) who noted that foliation/lineation
They suggested that flow necking (a consequence of melt extraction, patterns (Fig. 3) must be combined with kinematic and other in-
crystal growth and flow sorting) led to a 'megacryst logjam’, which formation to establish flow patterns. And Bergantz et al. (2017) ex-
continued to grow by the filtering out of megacrysts upstream (see also amine what fabrics record at the mineral population scale. The re-
Clarke and Clarke, 1998; Rocher et al., in press). Magma flow through cognition of multiple fabrics in granitoids is typically ignored in AMS
tubes and pipes, potentially either up or down, can be explained by and other quantitative fabric analyses in which one ellipsoid per station
thermal or compositional buoyancy (e.g., Griffiths, 1986; Martin et al., is determined. This forces multiple preferred orientations to be aver-
1987; Weinberg et al., 2001), Raleigh–Taylor fingering (Wiebe and aged and the loss of information about temporal histories. These
Collins, 1998; Perugini and Poli, 2005; Schofield et al., 2010), rising quantitative studies will be more valuable if combined with the re-
volatile-rich magmas (Clarke et al., 2013; Weinberg et al., 2001; cognition of multiple fabrics.
Memeti et al., 2014) or falling objects (Castro et al., 2008; Rocher et al.,
in press). 4.2.5. Deformation structures
Schlieren-bounded troughs are generally viewed as open (towards It is our interpretation that the 'deformation structures' listed above
more crystal poor magma) mush channels analogous to flow channels form in hydrogranular “submagmatic” mushes either at fast strain rates
with a 'bed-load' and formed in a porous media (e.g., Wahrhaftig, 1979; or as the magma is approaching its solidus and thus represent an im-
Dufek and Bergantz, 2007). Magma flow is interpreted to be parallel to portant record of the rheological behavior of magma chambers under
trough axes, inferred by aligned hornblendes in basal schlieren in these conditions.
granitoids (Paterson, 2009) or aligned, elongate plagioclase and olivine Magmatic fold morphology depends on the rheologic state of ma-
parallel to the trough axes in mafic rocks (Holness et al., 2017). The terial being folded, particularly whether or not an effective viscosity
intra-chamber channel flow is potentially caused by the disturbance contract exists between different folded phases (Fernandez and
and collapse of growing magma mush zones along margins (Humphreys Gasquet, 1994), and the relative crystal/melt proportion. Folding in
and Holness, 2010; Marsh, 2015) or erosion of previous pulses crystal poor magmas may be more akin to soft sediment folding in

143
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

which a fluid phase is present. Magmatic folding can obviously occur at 5. Conclusions
very fast strain rates (see volcanic examples and Albertz et al., 2005).
Examples below exemplify the potential of detailed studies of magmatic (1) A complex array of magmatic structures occur in plutons and pro-
folding. Vernon (1987) described magmatic folds in glassy rhyolitic vide useful tools for constraining hypersoidus temporal histories,
obsidian in which flow banding displays limb thinning (relative to evolving rheology, strain fields and inferred flow directions, growth
hinges) with crystallites and stretched vesicles defining an axial planar and cooling information, local and regional tilting, and syn-em-
foliation, reflecting strain during folding. However, the sense of relative placement tectonism.
rotation of the phenocrysts (Fig. 5p) remains constant around the folds (2) The above structural histories provide a powerful means of testing a
indicating that they record a pre-folding shear strain. Paterson et al. wide array of incremental emplacement, magma chamber growth
(1998) describe a magmatic fold in an ∼90 Ma pluton in the Cascades and evolution and syn-emplacement tectonic models.
core, Washington, where a magmatic foliation, defined by plagioclase (3) Many of these magmatic structures form in “hydrogranular” magma
and hornblende, is folded and a new magmatic foliation defined by slurries during local convection, collapse of mushes and strain re-
poikilitic hornblende grew parallel to the axial plane (Fig. 2k). The sulting in local structural diversity. Thus, the notion that magmas
number of overgrown grains define a minimum crystal percent during must have ≤55% crystals to convect, mix, fractionate or strain to
folding of ca 70%. Žák et al. (2007) describe a magmatically folded form compositional and structural diversity at these crustal levels,
leucogranite dike that has parasitic folds in both limbs and hinge, hinge needs to be revised.
thickening, and an axial planar magmatic foliation (Fig. 2l). Line length (4) Researchers using preferred orientations to indicate magma flow
method of 'unfolding' the dike indicates ∼28% shortening. must remember that these reflect strain rather than flow directly:
We are not aware of any systematic study of different types of the latter must be inferred from structural geometries and temporal
magmatic faults and shear zones in plutons, indicating that this is an and kinematic information.
important topic for future research. Published descriptions of individual (5) The presence of multiple fabrics in plutons requires caution in the
fault types suggest that there are three main groups, namely those interpretation of Anisotropy of Magnetic Suceptibility and other
caused by: (1) differential flow/strain gradients during local or quantitative tools that define preferred orientations at the sampling
chamber wide convection (e.g., Hibbard and Watters, 1985; Neufeld site through use of a single ellipsoid.
and Wettlaufer, 2008; Pinotti et al., 2016; Bergantz et al., 2017); (2) (6) Future structural studies are particularly needed in plutons ex-
marginal processes in growing crystal mush zones, particularly re- amining the distribution, styles and formation of compositional
sulting in normal faults during gravitational collapse of mushes defined structures, magmatic folding, shear zones, brittle faults, and
(Fig. 2q; e.g., Irvine et al., 1998; Solgadi and Sawyer, 2008; Paterson multiple fabrics in hydrogranular slurries.
et al., 2008) and cooling induced fractures (Geshi, 2001; John and
Stünitz, 1997; Žák et al., 2009; Humphreys and Holness, 2010); and (3) Acknowledgements
synchronous tectonism during magma chamber crystallization (e.g.,
Blumenfeld and Bouchez, 1988; John and Blundy, 1993; Zibra, 2012). We thank Marian Holness and Jean-Luc Bouchez for helpful reviews
Studies of magmatic faults are an underutilized tool to evaluate late and Joao Hibbertt for editorial assistance. Scott Paterson and Katie
strain fields in plutons, the collapse of crystal rich margins and localized Ardill acknowledge National Science Foundation support through
convection in magma chambers, which are otherwise resisting further grants EAR-0537892 and EAR-1624847. Paterson acknowledges dis-
flow or strain. cussions with Jorn Kruhl and George Bergantz about magmatic systems.
Jiří Žák acknowledges financial support from the Czech Science
Foundation through Grant No. 16-11500S and from the Charles
4.3. Modern approaches for studying magmatic structures University through Center for Geosphere Dynamics (UNCE/SCI/006)
and project PROGRES Q45.
A combination of the following techniques provide a powerful
means to address the key issues outlined above: (1) GIS-based, geor- References
eferenced grid-mapping provides key spatial information about the
type, orientation, and distribution of magmatic structures within plu- Albertz, M., Paterson, S.R., Okaya, D., 2005. Fast strain rates during pluton emplacement:
tons (e.g., Olivier et al., 1997; Launeau and Cruden, 1998; Peternell magmatically folded leucocratic dikes in aureoles of the Mount Stuart
Batholith,Washington, and the Tuolumne Intrusive Suite, California. Geol. Soc. Am.
et al., 2011; Rocher et al., in press): (2) crystal size distribution data Bull. 117, 450–465.
provide a quantitative measure of cooling and crystallization histories Alasino, P.H., Larrovere, M.A., Rocher, S., Dahlquist, J.A., Basei, M.A., Memeti, V.,
of plutonic bodies (e.g., Cashman and Marsh, 1988; Zieg and Marsh, Paterson, S.R., Galindo, C., Grande, M.M., Neto, M.D.C.C., 2017. Incremental growth
of an upper crustal, A-type pluton, Argentina: evidence of a re-used magma pathway.
2002; Jerram et al., 2003): (3) Electron Back-Scatter Diffraction pro- Lithos 284, 347–366.
vides mineral lattice preferred orientations and when combined with Archanjo, C.J., Launeau, P., Bouchez, J.L., 1995. Magnetic fabric vs. magnetite and biotite
qualitative mineral composition is useful for evaluating deformation shape fabrics of the magnetite-bearing granite pluton of Gameleiras (Northeast
Brazil). Phys. Earth Planet. Inter. 89, 63–75.
mechanisms and kinematic information (Žák et al., 2008; Holness et al., Ardill, K.E., Paterson, S.R., Barnes, C., Stanback, J., Alasino, P., Werts, K., Memeti, V.,
2017); (4) anisotropy of magnetic susceptibility is widely used to ana- Teruya, L., King, J., Crosbie, S., 2017. Magmatic structures in the Tuolumne Intrusive
lyze pluton fabrics (e.g., Tarling and Hrouda, 1993; Archanjo et al., Complex record physical differentiation of magmas at the emplacement level of an
upper crustal batholith. Geol. Soc. Am. 49http://dx.doi.org/10.1130/abs/2017AM-
1995; Bouchez, 1997; Borradaile and Henry, 1997; Kruckenberg et al.,
302303. Abstracts with Programs.
2010; Dietl et al., 2010; Tomek et al., 2017; Žák et al., 2017); (5) mi- Arzi, A.A., 1978. Critical phenomena in the rheology of partially melted rocks.
neral-scale electron microprobe and laser ablation-inductively coupled Tectonophysics 44, 173–184.
plasma mass spectrometry analyses, leading to the application of crystal Balk, R., 1937. Structural Behavior of Igneous Rocks, vol. 5. Geological Society of
America Memoir, pp. 177.
thermometers and chemometers are powerful tools to evaluate the Barbey, P., 2009. Layering and schlieren in granitoids: a record of interactions between
origin, timing and rheologic evolution of compositionally defined magma emplacement, crystallization and deformation in growing plutons. Geol. Belg.
magmatic structures (McBirney, 1993; Solgadi and Sawyer, 2008; 12, 109–133.
Barbey, P., Gasquet, D., Pin, C., Bourgeix, A.L., 2008. Igneous banding, schlieren and
Davidson et al., 2008; Barnes et al., 2016, 2017; Putirka, 2016; Zhang mafic enclaves in calc-alkaline granites: the Budduso pluton (Sardinia). Lithos 104,
et al., 2017); and (6) discrete element, computational fluid dynamics, 147–163.
numerical simulations of a hydrogranular magma mushes (e.g. Bergantz Barnes, C.G., Memeti, V., Coint, N., 2016. Deciphering magmatic processes in calc-alka-
line plutons using trace element zoning in hornblende. Am. Mineral. 101, 328–342.
et al. (2017). Barnes, C.G., Berry, R., Barnes, M.A., Ernst, W.G., 2017. Trace element zoning in

144
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

hornblende: tracking and modeling the crystallization of a calc-alkaline arc pluton. 119–145.
Am. Mineral. 102, 2390–2405. Emeleus, C.H., Troll, V., 2014. The rum igneous centre, Scotland. Mineral. Magaz. 78.
Barrière, M., 1981. On curved laminae, graded layers, convection currents and dynamic http://dx.doi.org/10.1180/minmag.2014.078.4.04.
crystal sorting in the Ploumanac'h (Brittany) subalkaline granite. Contrib. Mineral. Fernandez, A., 1988. Strain Analysis from Shape Preferred Orientation in Magmatic
Petrol. 77, 214–224. Flows, vol. 14. Bulletin of the Geological Institute, University of Uppsala, pp. 61–67.
Benn, K., 1994. Overprinting of magnetic fabrics in granites by small strains: numerical Fernandez, A.N., Gasquet, D.R., 1994. Relative rheological evolution of chemically con-
modeling. Tectonophysics 233, 153–162. trasted coeval magmas: example of the Tichka plutonic complex (Morocco). Contrib.
Bergantz, G.W., 2000. On the dynamics of magma mixing by reintrusion: implications for Mineral. Petrol. 116, 316–326.
pluton assembly processes. J. Struct. Geol. 22, 1297–1309. Fletcher, R.C., 1982. Analysis of the flow in layered fluids at small, but finite, amplitude
Bergantz, G.W., Schleicher, J.M., Burgisser, A., 2017. On the kinematics and dynamics of with application to mullion structures. Tectonophysics 81, 51–66.
crystal-rich systems. J. Geophys. Res. Solid Earth 122. Gaschnig, R.M., Vervoort, J.D., Tikoff, B., Lewis, R.S., 2017. Construction and preserva-
Blumenfeld, P., Bouchez, J.L., 1988. Shear criteria in granite and migmatite deformed in tion of batholiths in the northern US Cordillera. Lithosphere 9, 315–324.
the magmatic and solid states. J. Struct. Geol. 10, 361–372. Geoffroy, L., Callot, J.P., Aubourg, C., Moreira, M., 2002. Magnetic and plagioclase linear
Borradaile, G.J., Henry, B., 1997. Tectonic applications of magnetic susceptibility and its fabric discrepancy in dykes: a new way to define the flow vector using magnetic
anisotropy. Earth Sci. Rev. 42, 49–93. foliation. Terra Nova. 14, 183–190.
Bouchez, J.L., 1997. Granite is never isotropic: an introduction to AMS studies of granitic Geshi, N., 2001. Melt segregation by localized shear deformation and fracturing during
rocks. In: Bouchez, J.L., Hutton, D.H.W., Stephens, W.E. (Eds.), Granite: from crystallisation of magma in shallow intrusions of the Otoge volcanic complex, central
Segregation of Melt to Emplacement Fabrics. Kluwer Academic Publishers, Japan. J. Volcanol. Geotherm. Res. 106, 285–300.
Dordrecht, pp. 95–112. Gilbert, G.K., 1906. Gravitational assemblage in granite. Bull. Geol. Soc. Am. 17,
Bouchez, J.L., Delas, C., Gleizes, G., Nédelec, A., Cuney, M., 1992. Submagmatic micro- 321–328.
fractures in granites. Geology 20, 35–38. Griffiths, R.W., 1986. Thermals in extremely viscous fluids, including the effects of
Boudreau, A.E., 1995. Crystal aging and the formation of fine-scale igneous layering. temperature-dependent viscosity. J. Fluid Mech. 166, 115–138.
Contrib. Mineral. Petrol. 54, 55–69. Hardee, H.C., 1982. Incipient magma chamber formation as a result of repetitive intru-
Brown, M., 2013. Granite: from genesis to emplacement. Bull. Geol. Soc. Am. 125, sions. Bull. Volcanol. 45, 41–49.
1079–1113. Harm, S., 1991. Linear dilatation structures and syn-magmatic folding in granitoids. J.
Brown, P.E., Chambers, A.D., Becker, S.M., 1986. A large soft-sediment fold in the Lilloise Struct. Geol. 13, 625–634.
Intrusion, East Greenland. In: In: Parsons, I. (Ed.), Origins of Igneous Layering, NATO Hibbard, M.J., Watters, R.J., 1985. Fracturing and diking in incompletely crystallized
ASI Series C: Mathematical and Physical Sciences, vol. 196. D. Reidel Publishing granitic plutons. Lithos 18, 1–12.
Company, Dordrecht, pp. 125–144. Hines, R., Paterson, S.R., Memeti, V., Chambers, J.A., 2018;al., in press. Nested incre-
Burgess, S.D., Miller, J.S., 2008. Construction, solidification and internal differentiation mental growth of zoned upper crustal plutons in the Southern Uplands Terrane, UK:
of a large felsic arc pluton: cathedral Peak granodiorite, Sierra Nevada Batholith. fractionating, mixing, and contaminated magma fingers. J. Petrol. http://dx.doi.org/
Geol. Soc. Lond. Spec. Publ. 304, 203–233. 10.1093/petrology/egy034. (in press).
Cao, W., Paterson, S., Memeti, V., Mundil, R., Anderson, J.L., Schmidt, K., 2015. Tracking Hodge, K.F., Carazzo, G., Montague, X., Jellinek, A.M., 2012. Magmatic structures in the
paleodeformation fields in the Mesozoic central Sierra Nevada arc: implications for Tuolumne Intrusive Suite, California: a new model for the formation and deformation
intra-arc cyclic deformation and arc tempos. Lithosphere 7, 296–320. of ladder dikes. Contrib. Mineral. Petrol. 164, 587–600.
Cashman, K.V., Marsh, B.D., 1988. Crystal size distribution (CSD) in rocks and the kinetics Holness, M.B., Vukmanovic, Z., Mariani, E., 2017. Assessing the role of compaction in the
and dynamics of crystallization II. Makaopuhi lava lake. Contrib. Mineral. Petrol. 99, formation of adcumulates: a microstructural perspective. J. Petrol. 58, 643–673.
292–305. Humphreys, M.C.S., Holness, M.B., 2010. Melt-rich segregations in the Skaergaard
Cashman, K.V., Sparks, R.S.J., Blundy, J.D., 2017. Vertically extensive and unstable Marginal Border Series: tearing of a vertical silicate mush. Lithos 119, 181–192.
magmatic systems: a unified view of igneous processes. Science 355, eaag3055. Hutton, D.H.W., 1992. Granite sheeted complexes: evidence for the dyking ascent me-
Castro, A., Martino, R., Vujovich, G., Otamendi, J., Pinotti, L., D'Eramo, F., Tibaldi, A., chanism. Trans. R. Soc. Edinb. Earth Sci. 83, 377–382.
Viñao, A., 2008. Top-down structures of mafic enclaves within the Valle Fértil Iezzi, G., Ventura, G., 2000. Kinematics of lava flows based on fold analysis. Geophys.
magmatic complex (early Ordovician, san juan, Argentina). Geol. Acta 6, 217–229. Res. Lett. 27, 1227–1230.
Castro, J., Manga, M., Cashman, K.V., 2002. Dynamics of obsidian flows inferred from Irvine, T.N., 1980. Magmatic density currents and cumulus processes. Am. J. Sci. 280,
microstructures: insights from microlite preferred orientations. Earth Planet. Sci. Lett. 1–58.
199, 211–226. Irvine, T.N., 1987. Appendix I. Glossary of terms for layered intrusions. In: In: Parsons, I.
Clarke, D.B., Clarke, G.K.C., 1998. Layered granodiorites at chebutco head, South (Ed.), Origins of Igneous Layering, NATO ASI Series C: Mathematical and Physical
mountain batholith, Nova Scotia. J. Struct. Geol. 20, 1305–1324. Sciences, vol. 196. D. Reidel Publishing Company, Dordrecht, pp. 641–647.
Clarke, D.B., Grujic, D., McCuish, K.L., Sykes, J.C.P., Tweedale, F.M., 2013. Ring Irvine, T.N., Andersen, J.C.O., Brooks, C.K., 1998. Included blocks (and blocks within
schlieren: description and interpretation of field relations in the Halifax pluton, South blocks) in the Skaergaard intrusion: geologic relations and the origins of rhythmic
mountain batholith, Nova Scotia. J. Struct. Geol. 51, 193–205. modally graded layers. Geol. Soc. Am. Bull. 110, 1398–1447.
Cloos, E., 1936. Der Sierra Nevada Pluton in Californien. Neues Jahrbuch für Mineralogie. Jackson, E.D., 1971. The origin of ultramfic rocks by cumulus processes. Fortschritte für
Geol. Paleontol. 76, 355–450. Mineral. 48, 8–74.
Cloos, E., 1946. Lineation: a critical review and annotated bibliography. Geol. Soc. Am. Jerram, D.A., Cheadle, M.J., Philpotts, A.R., 2003. Quantifying the building blocks of
Mem. 18. http://dx.doi.org/10.1130/MEM18. igneous rocks: are clustered crystal frameworks the foundation? J. Petrol. 44,
Cloos, H., 1925. Einführung in die tektonische Behandlung magmatischer Erscheinungen: 2033–2052.
Das Riesengebirge in Schlesien. Borntraeger, Berlin. John, Barbara E., Blundy, Jonathan D., 1993. Emplacement-related deformation of
Coleman, D.S., Gray, W., Glazner, A.F., 2004. Rethinking the emplacement and evolution granitoid magmas, southern Adamello Massif, Italy. Geol. Soc. Am. Bull. 105 (12),
of zoned plutons: geochronologic evidence for incremental assembly of the Tuolumne 1517–1541.
Intrusive Suite, California. Geology 32, 433–436. John, B.E., Stünitz, H., 1997. Magmatic fracturing and small-scale melt segregation
Coleman, D.S., Bartley, J.M., Glazner, A.F., Pardue, M.J., 2012. Is chemical zonation in during pluton emplacement: evidence from the Adamello massif (Italy). In: Bouchez,
plutonic rocks driven by changes in source magma composition or shallow-crustal J.L., Hutton, D., Stephens, W.E. (Eds.), Granite: from Segregation of Melt to
differentiation? Geosphere 8, 1568–1587. Emplacement Fabrics. Kluwer Academic Publishers, Dordrecht, pp. 55–74.
Compton, R., 1955. Trondhjemite batholith near bidwell bar, California. Geol. Soc. Am. Johnson, S.E., Fletcher, J.M., Fanning, C.M., Vernon, R.H., Paterson, S.R., Tate, M.C.,
Bull. 66, 9–44. 2003. Structure, emplacement and lateral expansion of the San Jose tonalite pluton,
Correa-Gomes, L.C., Souza Filho, C.R., Martins, C.J.F.N., Oliveira, E.P., 2001. Peninsular Ranges batholith, Baja California, Mexico. J. Struct. Geol. 25, 1933–1958.
Development of symmetrical and asymmetrical fabrics in sheet-like igneous bodies: Katz, R., Spiegelman, M., Holtzman, B.K., 2006. The dynamics of melt and shear locali-
the role of magma flow and wall-rock displacements in theoretical and natural cases. zation in partially molten aggregates. Nature 442, 676–679.
J. Struct. Geol. 23, 1415–1428. King, B.C., 1964. The nature of basic igneous rocks and their relations with associated
Costa, A., Caricchi, L., Bagdassarov, N., 2009. A model for the rheology of particle bearing acid rocks. Part IV. Sci. Prog. 52, 282–292.
suspensions and partially molten rocks. Geochem. Geophys. Geosys. 10http://dx.doi. Komar, P.D., 1972. Flow differentiation in igneous dikes and sills: profiles of velocity and
org/10.1029/2008GC002138. Q03010. phenocryst concentration. Geol. Soc. Am. Bull. 83, 3443–3448.
Cruden, A., 1990. Flow and fabric development during the diapiric rise of magma. J. Geol. Kratinová, Z., Ježek, J., Schulmann, K., Hrouda, F., Shail, R.K., Lexa, O., 2010.
98, 681–698. Noncoaxial K-feldspar and AMS subfabrics in the Land's End granite, Cornwall: evi-
Davidson, J.P., Font, L., Charlier, B.L.A., Tepley, F.J., 2008. Mineral-scale Sr isotope dence of magmatic fabric decoupling during late deformation and matrix crystal-
variation in plutonic rocks: a tool for unraveling the evolution of magma systems. lization. J. Geophys. Res. 115 B09104.
Earth Environ. Sci. Trans. R. Soc. Edinb. 97, 35–67. Kruckenberg, S.C., Ferré, E.C., Teyssier, C., Vanderhaeghe, O., Whitney, D.L., Seaton,
Didier, J., 1973. Granites and Their Enclaves: the Bearing of Enclaves on the Origin of N.C.A., Skord, J.A., 2010. Viscoplastic flow in migmatites deduced from fabric ani-
Granites. Elsevier, Amsterdam. sotropy: an example from the Naxos dome, Greece. J. Geophys. Res. 115, B09401.
Dietl, C., de Wall, H., Finger, F., 2010. Tube-like schlieren structures in the Fürstenstein Larsen, R., Brooks, C.K., 1994. Origin and evolution of gabbroic pegmatites in the
Intrusive Complex (Bavarian Forest, Germany): evidence for melt segregation and skaergaard intrusion, east Greenland. J. Petrol. 35, 1651–1679. http://dx.doi.org/10.
magma flow at intraplutonic contacts. Lithos 16, 321–339. 1093/petrology/35.6.1651.
Dingwell, D.B., 1997. The brittle–ductile transition in high-level granitic magmas: ma- Launeau, P., Cruden, A., 1998. Magmatic fabric acquisition mechanisms in a syenite:
terial constraints. J. Petrol. 38, 1635–1644. results of a combined anisotropy of magnetic susceptibility and image analysis study.
Dufek, J., Bergantz, G.W., 2007. Suspended load and bed-load transport of particle-laden J. Geophys. Res. 103, 5067–5089.
gravity currents: the role of particle–bed interaction. Theor. Comput. Fluid Dyn. 21, Lucus, S.B., St-Onge, M.R., 1995. Syn-tectonic magmatism and the development of

145
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

compositional layering, Ungawa Orogen (northern Quebec, Canada). J. Struct. Geol. Putirka, K., 2016. Amphibole thermometers and barometers for igneous systems and
17, 475–491. some implications for eruption mechanisms of felsic magmas at arc volcanoes. Am.
Lusk, A.D.J., Paterson, S.R., Ratschbacher, B.C., Larrovere, M., Alasino, P., Memeti, V., Mineral. 101, 841–858.
Cawood, T.K., Hernandez, R., 2017. Deformation of the uppermost Famatinian Puziewicz, J., Wojewoda, J., 1984. Origin of load structures and graded beddings of dark
orogen: mapping and structural analysis from the Sierra de Navaráez and Sierra de minerals in the Strzegom granites (SW Poland): a sedimentological and petrological
Las Planchadas, NW Argentina. Geol. Soc. Am. 49http://dx.doi.org/10.1130/abs/ approach. Neues Jahrb. für Mineral. Monatsh. 8, 353–364.
2017AM-304220. Abstract with Programs. Ramsay, J.G., 1989. Emplacement kinematics of a granite diapir: the Chindamora
Macchioli Grande, M., Alasino, P.H., Rocher, S., Larrovere, M.A., Dahlquist, J.A., 2015. Batholith, Zimbabwe. J. Struct. Geol. 11, 191–209.
Asymmetric textural and structural patterns of a granitic body emplaced at shallow Reid, J., John, B., Murray, D.P., Hermes, O.D., Steig, E.J., 1993. Fractional crystallization
levels: the La Chinchilla pluton, northwestern Argentina. J. S. Am. Earth Sci. 64, in granites of the Sierra Nevada: how important is it? Geology 21, 587–590.
58–68. Rocher, S., Alasino, P.H., Macchioli Grande, M., Larrovere, M.A., Paterson, S.R., 2018;al.,
Marsh, B.D., 2015. Magmatism, Magma, and Magma Chambers. Elsevier. in press. K-feldspar megacryst accumulations formed by mechanical instabilities in
Martin, D., Griffiths, R.W., Campbell, I.H., 1987. Compositional and thermal-convection magma chamber margins, Asha pluton, NW Argentina. J. Struct. Geol. 112, 154–173.
in magma chambers. Contrib. Mineral. Petrol. 96, 465–475. http://dx.doi.org/10.1016/j.jsg.2018.04.017. ISSN 0191-8141.
McBirney, A.R., Noyes, R.M., 1979. Crystallization and layering of the Skaergaard in- Rosenberg, C., Berger, A., Schmid, S., 1995. Observations from the floor of a grantoid
trusion. J. Petrol. 20, 487–554. pluton: inferences on the driving force of final emplacement. Geology 23, 443–446.
McBirney, A.R., 1993. Igneous Petrology, second ed. Jones and Bartlett Publishers, Rosenberg, C.L., Handy, M.R., 2005. Experimental deformation of partially melted
Boston. granite revisited: implications for the continental crust. J. Metamorph. Geol. 23,
Memeti, V., Paterson, S., Putirka, K., 2014. Formation of the Sierra Nevada Batholith: 19–28.
Magmatic and Tectonic Processes and Their Tempos, vol. 34. Geological Society of Scaillet, B., Whittington, A., Martel, C., Pichavant, M., Holtz, F., 2000. Phase equilibrium
America Field Guide, pp. 116. constraints on the viscosity of silicic magmas 2: implications for mafic-silicic mixing
Miller, R.B., Paterson, S.R., 1994. Transition from magmatic to high-temperature solid- processes. Trans. R. Soc. Edinb. Earth Sci. 91, 61–72.
state deformation, Mount Stuart batholith, Washington. J. Struct. Geol. 16, 853–865. Schleicher, J.M., Bergantz, G.W., Breidenthal, R.E., Burgisser, A., 2016. Time scales of
Miller, R.B., Paterson, S.R., 2001. Construction of mid-crustal sheeted plutons: examples crystal mixing in magma mushes. Geophys. Res. Lett. 43.
from the North Cascades, Washington. Geol. Soc. Am. Bull. 113, 1423–1442. Schmalholz, S.M., Schmid, D.W., 2012. Folding in power-law viscous multi-layers. Philos.
Moore, J.G., Lockwood, J.P., 1973. Origin of comb layering and orbicular structure, Trans. R. Soc. A 370, 1798–1826.
Sierra Nevada batholith, California. Geol. Soc. Am. Bull. 84, 1–20. Schofield, N., Stevenson, C., Reston, T., 2010. Magma fingers and host rock fluidization in
Naslund, H.R., McBirney, A.R., 1996. Mechanisms of formation of igneous layering. In: the emplacement of sills. Geology 38, 63–66.
Cawthorn, R.G. (Ed.), Layered Intrusions. Elsevier, Amsterdam, pp. 1–43. Schulmann, K., Ježek, J., Venera, Z., 1997. Perpendicular linear fabrics in granite: mar-
Neufeld, J.A., Wettlaufer, J.S., 2008. An experimental study of shear-enhanced convec- kers of combined simple shear and pure shear flows? In: Bouchez, J.L., Hutton, D.,
tion in a mushy layer. J. Fluid Mech. 612, 363–385. Stephens, W.E. (Eds.), Granite: from Segregation of Melt to Emplacement Fabrics.
Nicolas, A., Ildefonse, B., 1996. Flow mechanism and viscosity in basaltic magma Kluwer Academic Publishers, Dordrecht, pp. 159–176.
chambers. Geophys. Res. Lett. 23, 2013–2016. Seaman, S.J., 2000. Crystal clusters, feldspar glomerocrysts, and magma envelopes in the
Olivier, P., de Saint Blanquat, M., Gleizes, G., Leblanc, D., 1997. Homogeneity of granite Atascosa Lookout lava flow, southern Arizona, USA: recorders of magmatic events. J.
fabrics at the metre and decametre scale. In: Bouchez, J.L., Hutton, D.H.W., Stephens, Petrol. 41, 693–716.
W.E. (Eds.), Granite: from Segregation of Melt to Emplacement Fabrics. Kluwer Shannon, J.R., Walker, B.M., Carten, R.B., Geraghty, E.P., 1982. Unidirectional solidifi-
Academic Publishers, Amsterdam, pp. 113–127. cation textures and their significance in determining relative ages of intrusions at the
Park, Y., Means, D., 1996. Direct observation of deformation processes in crystal mushes. Henderson Mine, Colorado. Geology 10, 293–297.
J. Struct. Geol. 18, 847–858. Shelley, D.M., 1985. Determining paleo-flow directions from groundmass fabrics in the
Parsons, I., 1986. Origins of Igneous Layering. NATO ASI Series, Series C: Mathematical Lyttleton radial dykes, New Zealand. J. Volcanol. Geotherm. Res. 25, 69–79.
and Physical Sciences, V. 196. D. Reidel Publishing Company, Dordrecht, Holland. Smith, J.V., Houston, E., 1994. Folds produced by gravity spreading of a banded rhyolite
Passchier, C.W., Trouw, R.A.J., 1996. Microtectonics. Springer, Berlin. lava flow. J. Volcanol. Geotherm. Res. 63, 89–94.
Paterson, S.R., 2009. Magmatic tubes, troughs, pipes, and diapirs: late-stage convective Solgadi, F., Sawyer, E.W., 2008. Formation of igneous layering in granodiorite by gravity
instabilities resulting in compositional diversity and permeable networks in crystal- flow: a field, microstructure and geochemical study of the Tuolumne Intrusive Suite
rich magmas of the Tuolumne Batholith, Sierra Nevada, California. Geosphere 5, at Sawmill Canyon, California. J. Petrol. 49, 2009–2042.
496–527. Stickel, J.J., Powell, R.L., 2005. Fluid mechanics and rheology of dense suspensions.
Paterson, S.R., Vernon, R.H., Tobisch, O.T., 1989. A review of criteria for identification of Annu. Rev. Fluid Mech. 37, 129–149.
magmatic and tectonic foliations in granitoids. J. Struct. Geol. 11, 349–363. Streck, M.J., 2008. Mineral textures and zoning as evidence for open system processes.
Paterson, S.R., T. Fowler, K., Schmidt, K., Yoshinobu, A., Yuan, S., 1998. Interpreting Rev. Mineral. Geochem. 69, 595–619.
magmatic fabric patterns in plutons. Lithos 44, 53–82. Talbot, J.Y., Martelet, G., Courrioux, G., Chen, Y., Faure, M., 2004. Emplacement in an
Paterson, S.R., Pignotta, G., Vernon, R.H., 2004. The significance of microgranitoid en- extensional setting of the Mont Lozére–Borne granitic complex (SE France) inferred
clave shapes and orientations. J. Struct. Geol. 26, 1465–1481. from comprehensive AMS, structural and gravity studies. J. Struct. Geol. 26, 11–28.
Paterson, S.R., Vernon, R.H., Žák, J., 2005. Mechanical instabilities and physical accu- Tarling, D., Hrouda, F., 1993. Magnetic Anisotropy of Rocks. Chapman and Hall.
mulation of K-feldspar megacrysts in granitic magma, Tuolumne Batholith, Tobisch, O.T., McNulty, B.A., Vernon, R.H., 1997. Microgranitoid enclave swarms in
California, USA. J. Virtual Explor. 18 Paper 01. granitic plutons, central Sierra Nevada, California. Lithos 40, 321–339.
Paterson, S.R., Žák, J., Janoušek, V., 2008. Growth of complex magmatic zones during Tomek, F., Žák, J., Verner, K., Holub, F.V., Sláma, J., Paterson, S.R., Memeti, V., 2017.
recycling of older magmatic phases: the Sawmill Canyon area in the Tuolumne Mineral fabrics in high-level intrusions recording crustal strain and volcano–tectonic
Batholith, Sierra Nevada, California. J. Volcanol. Geotherm. Res. 177, 457–484. interactions: the Shellenbarger pluton, Sierra Nevada, California. J. Geol. Soc. Lond.
Paterson, S., Okaya, D., Memeti, V., Economos, R., Miller, R., 2011. Magma addition and 174, 193–208.
flux calculations of incrementally constructed magma chambers in continental Urai, J.L., Spaeth, G., van der Zee, W., Hilger, C., 2001. Evolution of mullion (boudin)
margin arcs: combined field, geochronologic, and thermal modeling studies. structures in the Variscan of the Ardennes and Eifel. J. Virtual Explor. 3 Paper 1.
Geosphere 7, 1439–1468. van der Molen, I., Paterson, M.S., 1979. Experimental deformation of partially melted
Paterson, S., Memeti, V., Mundil, R., Žák, J., 2016. Repeated, multiscale, magmatic granite. Contrib. Mineral. Petrol. 70, 299–318.
erosion and recycling in an upper-crustal pluton: implications for magma chamber Vance, J.A., 1969. On synneusis. Contrib. Mineral. Petrol. 24, 7–29.
dynamics and magma volume estimates. Am. Mineral. 101, 2176–2198. Vaughan, A.P.M., Thistlewood, L., Millar, I.L., 1995. Small-scale convection at the in-
Perugini, D., Poli, D., 2005. Viscous fingering during replenishment of felsic magma terface between stratified layers of mafic and silicic magma, Campbell Ridges, NW
chambers by continuous inputs of mafic magmas: field evidence and fluid-mechanics Palmer Land, Antarctic Peninsula: syn-magmatic way-up criteria. J. Struct. Geol. 17,
experiments. Geology 33, 5–8. 1071–1075.
Peternell, M., Bitencourt, F.M., Kruhl, J.H., 2011. Combined quantification of anisotropy Ventura, G., 2004. The strain path and kinematics of lava domes: an example from Lipari
and inhomogeneity of magmatic rock fabrics in an outcrop scale analysis recorded in (Aeolian Islands, Southern Tyrrhenian Sea, Italy). J. Geophys. Res. 109 B01203.
high resolution. J. Struct. Geol. 33, 609–623. Vernon, R.H., 1987. A microstructural indicator of shear sense in volcanic rocks and its
Petford, N., 2009. Which effective viscosity? Mineral. Mag. 73, 167–191. relationship to porphyroblast rotation in metamorphic rocks. J. Geol. 95, 127–133.
Philpotts, A., Asher, P., 1994. Magmatic flow-direction indicators in a giant diabase Vernon, R.H., 2000. Review of microstructural evidence of magmatic and solid-state flow.
feeder dike, Connecticut. Geology 22, 363–366. Electron. Geosci. 5, 1–23.
Philpotts, A.R., Dickson, L.D., 2000. The formation of plagioclase chains during con- Vernon, R.H., 2004. A Practical Guide to Rock Microstructure. Cambridge University
vective transfer in basaltic magma. Nature 406, 59–61. Press.
Pinotti, L.P., D'Eramo, F.J., Weinberg, R.F., Demartis, M., Tubía, J.M., Coniglio, J.E., Vernon, R.H., Paterson, S.R., 2008a. Mesoscopic structures resulting from crystal accu-
Radice, S., Maffini, M.N., Aragón, E., 2016. Contrasting magmatic structures between mulation and melt movement in granites. Trans. R. Soc. Edinb. Earth Sci. 97,
small plutons and batholiths emplaced at shallow crustal level (Sierras de Córdoba, 369–381.
Argentina). J. Struct. Geol. 92, 46–58. Vernon, R.H., Paterson, S.R., 2008b. How extensive are subsolidus grainshape changes in
Pitcher, W.S., Berger, A.R., 1972. The Geology of Donegal: a Study of Granite cooling granites? Lithos 105, 42–50.
Emplacement and Unroofing. John Wiley, London. Vigneresse, J.L., Bouchez, J.L., 1997. Successive granitic magma batches during pluton
Pitcher, W.S., Atherton, M.P., Cobbing, E.J., Beckinsale, R.D., 1985. Magmatism at a Plate emplacement: the case study of Cabeza de Araya, Spain. J. Petrol. 38, 1767–1776.
Edge: the Peruvian Andes. John Wiley, London. Vigneresse, J.L., Barbey, P., Cuney, M., 1996. Rheological transitions during partial
Price, N.J., Cosgrove, J.W., 1990. Analysis of Geological Structures. Cambridge University melting and crystallization with application to felsic magma segregation and transfer.
Press. J. Petrol. 37, 1579–1600.

146
S.R. Paterson et al. Journal of Structural Geology 125 (2019) 134–147

Vigneresse, J.L., 2008. Granitic batholiths: from pervasive and continuous melting in the Geotherm. Res. 164, 254–267.
lower crust to discontinuous and spaced plutonism in the upper crust. Trans. R. Soc. Žák, J., Paterson, S.R., 2009. Magmatic erosion of the solidification front during re-
Edinb. Earth Sci. 97, 311–324. intrusion: the eastern margin of the Tuolumne batholith, Sierra Nevada, California.
Wager, L.R., Brown, G.M., 1968. Layered Igneous Rocks. Oliver and Boyd, Edinburgh. Int. J. Earth Sci. (Geol Rundsch) 99 (4), 801–812. http://dx.doi.org/10.1007/
Wahrhaftig, C., 1979. Significance of asymmetric schlieren for crystallization of granites s00531-009-0423-7.
in the Sierra Nevada batholith, California. Geol. Soc. Am. 11, 133 Abstracts with Žák, J., Paterson, S.R., Memeti, V., 2007. Four magmatic fabrics in the Tuolumne bath-
Programs. olith, central Sierra Nevada, California (USA): implications for interpreting fabric
Weinberg, R.F., Sial, R.N., Pessoa, R.R., 2001. Magma flow within the Tavares pluton, patterns in plutons and evolution of magma chambers in the upper crust. Geol. Soc.
northwestern Brazil: compositional and thermal convection. Geol. Soc. Am. Bull. 113, Am. Bull. 119, 184–201.
508–520. Žák, J., Verner, K., Týcová, P., 2008. Grain-scale processes in actively deforming magma
Wiebe, R.A., 1974. Coexisting intermediate and basic magmas, Ingonish, cape Breton mushes: new insights from electron backscatter diffraction (EBSD) analysis of biotite
Island. J. Geol. 82, 74–87. schlieren in the Jizera granite, Bohemian Massif. Lithos 106, 309–322.
Wiebe, R.A., Collins, W., 1998. Depositional features and stratigraphic sections in granitic Žák, J., Paterson, S.R., Janoušek, V., Kabele, P., 2009. The Mammoth Peak sheeted
plutons: implications for the emplacement and crystallization of granitic magma. J. complex, Tuolumne batholith, Sierra Nevada, California: a record of initial growth or
Struct. Geol. 20, 1273–1289. late thermal contraction in a magma chamber? Contrib. Mineral. Petrol. 158,
Wiebe, R.A., Jellinek, M., Markley, M.J., Hawkins, D.P., Snyder, D., 2007. Steep schlieren 447–470.
and associated enclaves in the Vinalhaven granite, Maine: possible indicators for Žák, J., Verner, K., Tomek, F., Johnson, K., Schwartz, J.J., 2017. Magnetic fabrics of arc
granite rheology. Contrib. Mineral. Petrol. 153, 121–138. plutons reveal a significant Late Jurassic to Early Cretaceous change in the relative
Wiebe, R.A., Jellinek, A.M., Hodge, K.F., 2017. New insights into the origin of ladder plate motions of the Pacific Ocean basin and North America. Geosphere 13, 11–21.
dikes: implications for punctuated growth and crystal accumulation in the Cathedral Zhang, J., Humphreys, M.C.S., Cooper, G.F., Davidson, J.P., Macpherson, C.G., 2017.
Peak granodiorite. Lithos 277, 241–258. Magma mush chemistry at subduction zones, revealed by new melt major element
Yoshinobu, A.S., Wolak, J.M., Paterson, S.R., Pignotta, G.S., Anderson, H.S., 2009. inversion from calcic amphiboles. Am. Mineral. 102, 1353–1367.
Determining relative magma and host rock xenolith rheology during magmatic fabric Zibra, I., 2012. Syndeformational granite crystallisation along the Mount Magnet
formation in plutons: examples from the middle and upper crust. Geosphere 5, Greenstone Belt, Yilgarn Craton: evidence of large-scale magma-driven strain loca-
270–285. lisation during Neoarchean time. Aust. J. Earth Sci. 59, 793–806.
Žák, J., Klomínský, J., 2007. Magmatic structures in the Krkonoše–Jizera Plutonic Zieg, M.J., Marsh, B.D., 2002. Crystal size distributions and scaling laws in the quanti-
Complex, Bohemian Massif: evidence for localized multiphase flow and small-scale fication of igneous textures. J. Petrol. 43, 85–101.
thermal-mechanical instabilities in a granitic magma chamber. J. Volcanol.

147

You might also like