Vous êtes sur la page 1sur 33

Downloaded from gsabulletin.gsapubs.

org on February 27, 2013

Geological Society of America Bulletin

Low-temperature deformation mechanisms and their interpretation


RICHARD H. GROSHONG, JR.

Geological Society of America Bulletin 1988;100, no. 9;1329-1360


doi: 10.1130/0016-7606(1988)100<1329:LTDMAT>2.3.CO;2

Email alerting services click www.gsapubs.org/cgi/alerts to receive free e-mail alerts when new articles cite
this article
Subscribe click www.gsapubs.org/subscriptions/ to subscribe to Geological Society of America
Bulletin
Permission request click http://www.geosociety.org/pubs/copyrt.htm#gsa to contact GSA

Copyright not claimed on content prepared wholly by U.S. government employees within scope of their
employment. Individual scientists are hereby granted permission, without fees or further requests to GSA, to
use a single figure, a single table, and/or a brief paragraph of text in subsequent works and to make
unlimited copies of items in GSA's journals for noncommercial use in classrooms to further education and
science. This file may not be posted to any Web site, but authors may post the abstracts only of their articles
on their own or their organization's Web site providing the posting includes a reference to the article's full
citation. GSA provides this and other forums for the presentation of diverse opinions and positions by
scientists worldwide, regardless of their race, citizenship, gender, religion, or political viewpoint. Opinions
presented in this publication do not reflect official positions of the Society.

Notes

Geological Society of America


Downloaded from gsabulletin.gsapubs.org on February 27, 2013

Low-temperature deformation mechanisms


and their interpretation
CENTENNIAL ARTICLE

RICHARD H. GROSHONG, J r . Department of Geology, The University of Alabama, Tuscaloosa, Alabama 35487-1945

ABSTRACT section. Interpretations are given in terms of the stress or strain that causes
the features to develop and the stress or strain that may be inferred from
Low-temperature deformation is characterized by heterogeneous the features. Low-temperature deformation generally means deformation
strain in which the bulk of the material clearly retains its primary at one-third or less of the melting temperature of the framework minerals.
texture. Deformation is by grain-scale crystal plasticity, rotation, frac- Under these conditions, individual crystals are strained less than about
ture, and pressure solution, and by transgranular mechanisms that 15%-20%, a magnitude that leaves the rock looking relatively undeformed
crosscut numerous grains. The important low-temperature crystal- (Fig. 1). There is minimal recrystallization except in fine-grained, water-
plastic features are twin lamellae, deformation bands, and undulatory wet materials which may recrystallize as a result of either diagenesis or
extinction. Subgrain formation by recrystallization or crystal-plastic deformation. Framework grains do not react to form new minerals. Phyl-
strain of more than 15% marks the upper limit of the low-temperature losilicates may change species but retain their constituent sheets of silica
regime. Grain rotation may produce foliations in soft sediments or tetrahedra (Oertel, 1983). Uncemented grains may be free to slide past one
rocks. Microscopic to mesoscopic kinks and crenulations of bedding another under shallow-burial or high-pore-pressure conditions. Large
occur in soft clay and shale. Transgranular features include Lfiders' strains usually occur by the transgranular mechanisms of extension frac-
bands, cooling and desiccation cracks, joints, extension-fracture tures, faults, stylolites, or solution cleavage. It is the material between these
cleavage, clastic dikes, mineral-filled veins of several types, recrystalli- transgranular features that remains relatively undeformed. Strain usually
zation/replacement veins, vein arrays, boudins, faults, stylolites, slick- reduces porosity and permeability in high-porosity materials but may in-
olites, solution cleavages that range from widely spaced to slaty and crease it in low-porosity materials.
pencil cleavage. Pressure fringes form adjacent to relatively rigid Low-temperature deformation is often characterized as brittle, where
grains and have fabrics analogous to those in veins. Faults include brittle phenomena include a number of processes that are dominated by
conjugate fault pairs (Andersonian faults) multiple simultaneous con- fracture as a mechanism or are controlled by the presence of pre-existing
jugates (Oertel faults), and Riedel shear-zone configurations. The fractures (Logan, 1979). A fracture is defined as a surface along which a
sense of fault displacement is determined from bends, steps, trails, loss of cohesion has taken place (Griggs and Handin, 1960). Brittle defor-
tails, and feather fractures. Superplasticity, especially if aided by diffu-
sion in grain-boundary water, might be important at low tempera-
tures. Fault textures are diagnostic of the environment of deformation
but have yet to be uniquely correlated with the presence or absence of
earthquakes. Riedel shears and pseudotachylite may form in earth-
quake source regions, although pseudotachylite is evidently rare in
brittle fault zones. The best indicators of stress magnitudes are the
critical resolved shear stress for deformation twinning and the pres-
ence of tensile fractures. Strain magnitudes and stress and strain ten-
sor orientations can be determined with a variety of methods that are
based on mechanical twins, platy grain orientation, grain center distri-
bution, and fault geometry and slip directions. Different deformation
mechanism associations, expressed by the partitioning of the total
strain into different mechanisms, are related to the ductility and envi-
ronment of deformation. Deformation fronts separating different
mechanism associations are defined on the basis of changes in the
crystal-plastic component of strain.

Rocks do not suffer deformation; they enjoy it. Figure 1. Low-temperature deformation of the Upper Cretaceous
Rob Knipe, 1982 Schrattenkalk, eastern Switzerland (sample 11, Groshong and others,
1984). Based on the twin strain, the rock has enjoyed a shortening of
INTRODUCTION 6% at a temperature on the order of 150 °C (from the vitrinite reflec-
tance of about 3.0% in nearby rocks, using the well-standard curve of
This Centennial article reviews the identification and interpretation of Bostick, 1974). The fossil was discovered and photographed by Laurel
low-temperature deformation mechanisms as seen in outcrop and thin Pringle-Goodell.

Geological Society of America Bulletin, v. 100, p. 1329-1360,2 figs., September 1988.

1329
Downloaded from gsabulletin.gsapubs.org on February 27, 2013
1330 R. H. GROSHONG, JR.

mation is usually contrasted with ductile deformation; however, the term at a different position than the host grain. Thin twins do not contain
"ductile" is used in several different senses (Rutter, 1986): (1) plastic distinguishable twinned material under the optical microscope but appear
deformation of crystals, (2) homogeneous deformation by any mechanism as thin dark lines (Knopf, 1949) that were called "untwinned" lamellae by
(the uniform flow of Griggs and Handin, 1960), and (3) more than a Turner (1953). Conel (1962) showed that "untwinned" lamellae exhibit
specified amount of permanent strain (l%-2%; Logan, 1979) prior to the interference fringes predicted to be associated with twins, and under
faulting. A problem in using the terms in the field is whether soft-sediment the scanning electron microscope they have been seen to be twins (Spang
faults and deformation by transgranular pressure solution should be con- and others, 1974). Twins formed at low temperatures in calcite are thin
sidered brittle or ductile. Current usage favors the former term, and so (1-10 fi) and plane sided, whereas twins formed at high temperatures are
"brittle" is here redefined as deformation with less than l%-2% crystal- thicker and tend to be convex sided (Schmid and others, 1980).
plastic strain prior to transgranular failure. "Semi-brittle" (suggested by Twin gliding occurs on a specific glide plane and in a specific direc-
C. Simpson, 1987) is deformation with 2%-20% crystal-plastic strain prior tion along the glide line. The direction and sense of shear can always be
to transgranular failure, and "ductile" is deformation with >20% crystal- uniquely determined for a given twin set (Bell, 1941; Turner and others,
plastic strain with or without transgranular strain. Low-temperature strain 1954a), which leads to the use of twins in dynamic and strain analysis,
is brittle to semi-brittle. Large homogeneous deformation may occur by discussed later. The twinning geometry in dolomite was initially proposed
soft-sediment flow, cataclastic flow (Stearns, 1969), or pressure solution, fromfieldsamples by Fairbairn and Hawkes (1941) and later confirmed in
each being brittle or semi-brittle if the crystal-plastic strain is small. Thus laboratory experiments by Handin and Fairbairn (1955) and Turner and
"brittle" is the opposite of "crystal-plastic" (Rutter, 1986), rather than the others (1954b, p. 483). Mechanical Dauphine twinning in quartz produces
opposite of "flow." strong preferred orientations but causes no permanent strain (Tullis, 1970;
The review begins with crystal-plastic deformation features and Tullis and Tullis, 1972). According to Vance (1961), deformation twin-
moves on to the mainly transgranular mechanisms of extension fracture, ning was first recognized in plagioclase by van Werveke in 1883 and first
pressure solution, and faulting. The stress, strain, and sense-of-shear inter- produced experimentally by Miigge and Heide in 1931. Deformation
pretations of the mechanisms are discussed along with their descriptions. twinning occurs according to (a) the albite law, for which the twin plane is
Techniques for inferring stress or strain from the bulk deformation are (010) and the glide line is irrational and (b) the pericline law, for which the
reviewed next. The final discussions are of strain partitioning, flow laws, twin plane is a rhombic section and the glide line is <010>, both with
deformation fronts, and deformation-mechanism associations. Many field positive sense of shear (Borg and Heard, 1970). The two directions of
observations have not been clearly understood until reproduced in the twinning are about 94° apart, and both develop during experimental de-
laboratory. An attempt is made here to bring together data from the field formation (Borg and Handin, 1966; Borg and Heard, 1970, p. 387). Me-
and the laboratory. Numerous illustrations would be desirable but are chanical twinning has been suggested for microcline (Marmo and Binns, in
precluded by length constraints, and so the references cited in the text Spry, 1969) but has not yet been produced in the laboratory. The twin
include page numbers that refer to key photographs, diagrams, or original laws for 69 different crystal species are listed by Handin (1966,
definitions. p. 241-243).
Deformation versus growth as the cause of twinning has long been
CRYSTAL PLASTICITY debated. Growth twins tend to be single twins that divide a crystal into two
units; deformation twins are lamellar or polysynthetic (Spry, 1969). In
Crystal plasticity is the permanent deformation without loss of cohe- feldspar, lamellar twins are attributed to both growth and deformation.
sion of single crystals by internal deformation mechanisms. Mechanisms Vance (1961) suggested that growth twins in plagioclase are wide and
quantitatively important at low temperatures are twin glide and translation highly variable in thickness, occur in small numbers, and often change
glide. Higher-temperature dislocation climb and diffusion mechanisms re- width or terminate abruptly and independently of other twins, whereas
sult in polygonization and recrystallization, mentioned here because they deformation twins are thin, numerous, and, if they die out within a crystal,
define the upper limit of low-temperature deformation. Grain-boundary tend to do so along a kink boundary or parallel to the trend of the
diffusion may be important at both low and high temperatures. The term undulatory extinction. On the other hand, this difference closely resembles
"plastic" is also widely used for the permanent deformation without loss of the difference between high- and low-temperature deformation twins pro-
cohesion of the bulk rock, approximately synonymous with "flow" duced experimentally in calcite (Schmid and others, 1980). Spry (1969)
(Griggs, 1942). Stress-strain curves typically show a yield stress that marks concluded that most polysynthetic twins in metamorphic plagioclase, mi-
the change from recoverable to permanent strain (Nadai, 1931), and so crocline, calcite, and dolomite are the result of deformation. Calcite and
"plastic" is often used in "those cases where the presence of a yield value dolomite twins have been very useful in the study of deformation; more
can be assumed" (Burgers and Burgers, 1935). The first experimental proof attention should be given to the feldspars.
that normally brittle rocks could be made to flow was by Adams and
Nicolson (1901), and the importance of both temperature and pressure Deformation Lamellae and Undulatory Extinction
was demonstrated by Adams and Coker (1910) and von Kdrmin (1911).
The modern era of experimentation began with the high-pressure experi- Translation glide indirectly results in deformation lamellae, certain
ments of Griggs (1936). Analysis of plastic deformation usually requires types of undulatory extinction, and kinks in single crystals. Handin (1966,
oriented samples; see Prior and others (1987) for collecting procedures. p. 238-240) gave the translation glide systems for 81 different crystal
species. Deformation lamellae are parallel sets of very thin, optically dis-
Twin Lamellae tinct bands within a crystal that are not mineral cleavage, open fractures,
or twins. In the optical microscope, one type of deformation lamellae in
Twin lamellae are thin bands of twinned material usually found in quartz consists of narrow, closely spaced subplanar features having slightly
sets of many parallel lamellae. First recognized in calcite by Huygens in the different refractive indices from the host quartz (Fairbairn, 1941, PI. 2;
late 1600s, they were shown by Reusch in the mid-1800s to be the result Ingerson and Tuttle, 1945; Carter and others, 1964, PI. 2; Christie and
of deformation according to Klassen-Neklyudova (1964). Sorby (1908, PL others, 1964), causing them to appear as parallel bands of brighter and
18) recognized them as a deformation feature in limestone. A thick twin is darker material that terminate at crystal boundaries. They are subbasal or
an optically visible domain that goes to extinction under crossed polarizers basal in quartz (Carter, 1971). Certain inclusion-rich lamellae in quartz
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

LOW-TEMPERATURE DEFORMATION MECHANISMS 1331

have been called "Böhm" or "Boehme" lamellae (Böhm, 1883); however, compared to the sharp boundaries of subgrains (Tullis and Yund, 1987,
this usage is vague and includes healed fractures (Griggs and Bell, 1938) p. 607). Patchy extinction has been produced experimentally in the
and so should not be used (Fairbairn, 1941). I suggest the term "Fairbairn low-temperature cataclastic flow of albite (Tullis and Yund, 1987).
lamellae" for deformation lamellae resulting from a difference in refractive The fabric seems to penetrate the entire grain, rather than forming first
index relative to the host. In quartz they are usually associated with, and at at grain boundaries. This texture should form during low-temperature
a high angle to, parallel bands of undulatory extinction. Irrational Fair- deformation.
bairn lamellae have been produced in plagioclase (Borg and Heard, 1970).
Quartz contains rational Fairbairn lamellae that are usually parallel to the Subgrains and Recrystallized Grains
basal plane or, less commonly, parallel to a rhomb. TEM reveals that the
basal lamellae may be thin zones of glass with associated dislocation loops Subgrains are regions within a crystal that are misoriented from the
(S. White, 1973c; Christie and Ardell, 1974) or paired stacking faults host by an angle of greater than one degree (Nicolas and Poirier, 1976,
interpreted as Brazil twins (S. White, 1973a, p. 27) and that the prismatic p. 139). Under the microscope, they are small, clear areas separated from
lamellae are zones of high dislocation density (Twiss, 1974). Subbasal one another by distinct, low-angle boundaries (Tullis and Yund, 1987).
Fairbairn lamellae in quartz are not parallel to any crystallographic plane Polygonization texture was defined by Nicolas and Poirier (1976, p. 137)
but usually make an angle of - 3 0 ° to the basal plane. Ingerson and Tuttle as having subgrains misoriented from the host by angles of 1°-15° and a
(1945) recognized on the basis of field data that these lamellae formed in recrystallization texture as having grains that are misoriented from the host
planes of high shear stress, a fact later demonstrated experimentally by by more than 15°. The typical polygon size is 50-100 /urn but may be up
Hansen and Borg (1962). In TEM, irrational lamellae are seen to have to 1 mm (Nicolas and Poirier, 1976, p. 139). White (1976, p. 74) accepted
several different origins (S. White, 1973c, p. 27; Christie and Ardell, 1974; a 10° misorientation as sufficient to define a new grain boundary that
Ardell and others, 1976; Twiss, 1974,1976): (1) walls of dislocations that represents recrystallization. Subgrains produce a blocky undulatory extinc-
form narrow, elongated basal or prismatic subgrains and are decorated tion. In the early stages of development, subgrains are likely to be localized
with bubbles; (2) subgrains that are too thin to be optically resolvable; along older deformation bands or grain boundaries (White, 1973b, 1976),
(3) dislocation tangles; and (4) alternating slabs of high- and low- forming serrated boundaries (Hobbs and others, 1976, p. 115). This should
dislocation density. Early experiments (Griggs, 1936; Griggs and others, be a relict texture in rocks deformed at low temperature.
1960b) required unusually high stresses to produce this type of deforma- Subgrain formation occurs by dislocation-glide-related processes that
tion in quartz. Griggs and Blacic (1965) accidentally discovered that small are thermally activated: dynamic recovery and dynamic recrystallization
amounts of water drastically weaken the crystals. This hydrolitic weaken- (Tullis and Yund, 1985). Dynamic recovery (polygonization of Poirier,
ing results in more realistic experimental stresses. The relationship of the 1985, p. 174) results in the formation of subgrains with less than 10°
lamellae orientation to the crystal structure is a function of pressure, misorientation (Tullis and Yund, 1985). Progressive subgrain rotation can
temperature, and strain rate (Carter, 1971) with the rational lamellae being produce higher angle mismatches within this regime (Poirier, 1985; Tullis
the result of shock deformation and irrational lamellae being formed under and Yund, 1985). Dynamic recrystallization occurs as either rotation re-
normal tectonic strain rates. crystallization or migration recrystallization (Poirier, 1985). In rotation
Sander (1930) made the first quantitative measurements of undula- recrystallization, low-angle-boundary subgrains rotate to high-angle mis-
tory extinction attributed to deformation in quartz. The first experimental matches (Poirier, 1985, p. 181). In migration recrystallization, grain
study was by Griggs and Bell (1938), who produced undulatory extinction boundaries migrate between undeformed subgrains and deformed larger
but no other deformation features. Lattice bending results in a sweeping grains (Poirier, 1985, p. 182). Syntectonic migration recrystallization tends
extinction, and the formation of new boundaries within the crystal leads to to produce jigsaw-puzzle-like grain boundaries (Hobbs, 1968, p. 387),
blocky and patchy extinction. whereas static migration recrystallization produces a foam texture. Dy-
Kinking in a single crystal results from a slip instability on a weak namic recovery and recrystallization textures are generally relict in the
direction within the crystal structure and results in extinction bands. Con- low-temperature environment, but a foam texture might develop at low
jugate kink bands may form in crystals with very strong planar anisotropy, temperatures in veryfine-grainedrocks.
such as brucite and gypsum (Turner and Weiss, 1965, p. 359-364) or In a mortar texture (Spry, 1969, PI. XXVIII) or core and mantle
mica (Borg and Handin, 1966, p. 331). A single direction of kinking is structure (White, 1976), large original grains are surrounded by and in-
characteristic of crystals with a weaker anisotropy (Carter, 1968), such as vaded by small subgrains or recrystallized grains (Carter and others, 1964,
dolomite (Higgs and Handin, 1959), basal glide in quartz (Christie and Pis. 8-10; Neumann, 1969, PI. 1; Tullis and others, 1973, p. 305). A
others, 1964), r glide in calcite (Turner and Heard, 1965; Carter, 1976, similar texture may also occur within large host grains which recrystallize
p. 311)/glide in calcite (Wenk and others, 1973; Nicolas and Poirier, along high-energy sites at kink-band boundaries or twin lamellae (Hobbs,
1976), or slip on (010) in plagioclase (Seifert, 1965, p. 1470-1471). 1968, p. 356-380). The orientation of the new grains tends to be related to
Gentle curvature at the kink boundaries gives a linear undulatory extinc- that of the host, such as by a rotation about an a-axis in quartz (White,
tion, whereas sharp curvature produces well-defined kink bands (Carter 1976). New grain formation at grain boundaries indicates that the recrys-
and others, 1964, Pis. 4 and 5). Thin, well-defined bands that have a tallization occurred in situ and was not inherited from the source terrane.
different crystallographic orientation from the host may be called "defor- The size of dynamically recrystallized grains is inversely related to
mation bands" (Carter and others, 1964; Christie and others, 1964). differential stress in single-phase materials (Twiss, 1977; Christie and Ord,
Strongly developed deformation bands may cause a single crystal to look 1980; Kohlstedt and Weathers, 1980; Etheridge and Wilkie, 1981; Ord
like several very elongate individual crystals (Carter and others, 1964, PI. and Christie, 1984; Ranalli, 1984; Poirier, 1985), although the exact rela-
5). The deformation origin of this type of undulatory extinction was first tionship remains in doubt. If grain sizes are stabilized by a second phase,
shown by lattice distortions seen in X-ray patterns (Bailey and others, such as mica, then there may be no relationship to stress, and the relation-
1958). ship may be a function of temperature and strain rate (White, 1979).
Submicroscopic microcracking and crushing between relatively un- Subgrain size may reflect the maximum stress in a history of variable stress
deformed zones can produce a patchy undulatory extinction that resembles magnitudes (Poirier, 1985). In a rotational strain experiment, Friedman
poorly defined subgrains (Tullis and Yund, 1987). Patchy extinction in- and Higgs (1981, p. 20) found that the larger recrystallized grains showed
volves more continuous and irregular changes in extinction position as the most consistent relationship between grain size and magnitude of stress.
Downloaded from gsabulletin.gsapubs.org on February 27, 2013
1332 R. H. GROSHONG, JR.

The Ambiguous Foam Texture tively correlate the degree of preferred orientation in shale to the perfection
of the fissility.
In a foam texture (Stanton, 1972, p. 233), the grains are plane-sided The parallel orientation of platy minerals has been attributed to depo-
polygons having triple-point boundaries, as seen in a closely packed foam. sition, deformation, compaction, and oriented grain growth. Clay may be
The triple-point junctions between 3 grains are usually at 120° angles deposited with a horizontal preferred orientation from fresh water but is
(Voll, 1960) when viewed normal to the boundary. The angle may vary usually deposited with an open house-of-cards framework of small packets
from 120°, depending on the relative surface energies of the materials in of clay flakes (Moon, 1972; Collins and McGowan, 1974; Oertel, 1983).
contact (Stanton, 1972). This texture is characteristic of foam, com- It has been demonstrated that constant-volume strain can cause randomly
pressed lead shot, and a variety of other organic and inorganic materials oriented clay flakes to become aligned perpendicular to the direction of
(Folk, 1965), certain undeformed and deformed sedimentary rocks, and is shortening experimentally by Sorby (1853,1908), Clark (1970), and Fer-
well known as the end result of the high-temperature static recrystalliza- nandez (1987). Shortening with dewatering produces only a slightly
tion of deformed rocks. Griggs and others (1960a, PI. 1) published the first greater degree of preferred orientation than does shortening alone (Clark,
example of annealing in quartz. Annealing recrystallization does not pro- 1970). Oertel and Curtis (1972) demonstrated that compaction was the
duce a crystallographic orientation fabric, although traces of the pre- cause of the preferred orientation seen in one shale, and it is a likely cause
existing fabric remain because the orientation of the new crystals is related for the fabric in many others. The fabrics obtained by syndeformational
to that of the old (Griggs and others, 1960a; Hobbs, 1968, p. 357). synthesis or recrystallization experiments on mica produce the same degree
In unmetamorphosed limestones, foam texture is characteristic of of preferred orientation as strain without recrystallization (Means and
micrite (2- to 3-fx grain size) and microspar (5- to 10-^ grain size) illus- Paterson, 1966; Tullis, 1976), leading these authors to conclude that rota-
trated by Folk (1965, p. 34, 38). Microspar tends to be loaf shaped (Folk, tion was the predominant orienting mechanism.
1965). Many examples of micrite and microspar are interpreted to be the Slumping or tectonic deformation of soft sediments can produce
result of diagenetic recrystallization of unstrained magnesian calcite or planar and/or linear fabrics in the unconsolidated materials. In this case,
aragonite needles in an environment with a low Mg/Ca ratio or where Mg the experiments of Sorby (1853, 1908) and Clark (1970) apply directly,
is removed from the rock (Folk, 1974). Diagenetic recrystallization of rather than as rock analogs. A well-developed phyllosilicate foliation is
fine-grained calcite has been accomplished experimentally by Baker and seen in clay-rich rocks from a subduction zone (Lundberg and Moore,
others (1980) who found that the presence of clay minerals retarded the 1986, p. 19) where it is associated with a slickensided scaley fabric. Soft-
process. This suggests that relatively purefine-grainedmaterials will read- sediment slump and buckle folds develop foliations that are caused by grain
ily recrystallize at low temperatures and may likewise anneal or deform at rotation (Moore and Geigle, 1974, p. 509; Brodzikowski and others,
low temperatures by mechanisms usually considered to be restricted to 1987). Williams and others (1969, p. 421-422) described both a single
high temperatures. Keller and others (1985, p. 1354-1356) showed foam- axial-plane parallel-mineral foliation in folded fine-grained shale (PI. 2)
textured chert (in the Ouachita thrust belt) in which the grain size system- and double planar-mineral foliations, separated by up to about 20° and
atically increases with degree of metamorphism. The thermal data of bisected by the axial plane. Slump folds in a shale-siltstone sequence may
Guthrie and others (1986) show that the initial recrystallization began at a have an axial-plane preferred orientation of platy minerals that is penetra-
vitrinite reflectance of about 1.5% with a l-jim crystal size and that crystal tive in thin section (Woodcock, 1976). In a soft-sediment fold in sandstone
size increased up to 4 jum or more at a reflectance of 3%. This represents described by Yagishita and Morris (1979, p. 109), the quartz grains have a
temperatures of 150-200 °C according to the "well-standard" correlation long-axis lineation parallel to the fold axis and weak axial-plane foliation
of Bostick (1974, p. 9). All of the rocks are folded or thrusted. Syntaxial defined by the intermediate dimensions of ellipsoidal grains. On the other
cement may also produce a foam texture. hand, soft-sediment folds are usually characterized as not having cleavage
Recently a foam-like texture has been recognized as characteristic of (Helwig, 1970; Pickering, 1987). My personal experience is that axial-
superplastic deformation in which the dominant deformation mechanism plane mineral foliations are reasonably common in folded soft sediments
is diffusion-aided, grain-boundary sliding (see discussion below of super- and that sediments or poorly consolidated rocks will usually break into
plastic fault zones). In superplastic deformation, grains are relatively unde- rough slabs parallel to the foliation. This foliation does not resemble the
formed internally, the crystallographic fabric is weak, and the grains tend near-penetrative smooth planes of splitting that characterize slaty cleavage
to be square or loaf shaped. or the well-defined compositional banding of spaced cleavage, nor are the
rocks lithified to the same degree as slate.
GRAIN ROTATION Rotation of pebbles in a conglomerate having a fine-grained matrix
produces a tectonic polish and aligned microstriations (Judson and Barks,
Parallel Fabric 1961, p. 373). Clifton (1965, p. 871) documented an area in which the
striations from all clasts were oriented parallel to dip and concluded that
Primary ellipsoidal or platy grains may show a high degree of paral- the striations were the result of structural deformation, although striations
lelism within a rock. This is commonly observed in shales (Ingram, 1953; also occur in conglomerates that are not deformed tectonically (Judson
Moon, 1972) where the fabric is contrasted to that of mudstones in which and Barks, 1961; Wiltschko and Sutton, 1982). External rotation of a
the grain alignment is absent. Parallelism of platy minerals is characteristic pebble due to slip along internal shear planes was found by Tyler (1975,
of slates (Sorby, 1853; flow cleavage of Leith, 1905), and parallelism of p. 507) in a small fault zone. Mosher (1987) attributed a major component
pebbles and cobbles is seen in deformed conglomerates (Leith, 1905). In of the preferred clast orientation in a deformed conglomerate to rotation
fine-grained, phyllosilicate-rich rocks, the fabric is observed in thin section during deformation.
as a tendency for the whole section to go to extinction at the same time
(Sorby, 1880; Dale and others, 1914). The mineral fabric can be observed Interweaving Fabric
directly using the SEM (Davies, 1982, p. 65) and indirectly as point
concentrations in X-ray diffraction patterns (Odom, 1967, p. 612; Oertel, A mixture of equant grains with platy grains can result in an anas-
1970; Oertel and Curtis, 1972, p. 2601). Sorby (1853) correlated the tomosing fabric that might be interpreted as being the result of multiple
degree of preferred phyllosilicate orientation in slate to the quality of the deformation. The fabric is, however, common in natural soils (Brewer,
cleavage, and Ingram (1953) appears to have been the first to quantita- 1964, p. 311-333; Collins and McGowan, 1974, p. 238-241) and is
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

LOW-TEMPERATURE DEFORMATION MECHANISMS 1333

shown to occur in certain clays as well as in silt-clay mixtures. Brewer Geiser and Engelder, 1983, p. 163; Nickelsen, 1986, p. 365). The symmet-
(1964) attributed the fabric to rotation of platy minerals caused by move- ric geometry of the crenulations is favored when the maximum compres-
ments resulting from expansion and contraction caused by wetting and sive stress is parallel to layering (Reches and Johnson, 1976).
drying. Large phyllosilicate grains in a matrix of finer grains may also
show oblique orientations (Weaver, 1984, p. 13). Lennox (1987) pro- EXTENSION FRACTURE
duced a similar fabric experimentally by deforming a mica-quartz-sand
mixture under conditions favoring recrystallization, showing that, just as An extension fracture (Griggs and Handin, 1960) forms perpendicu-
with parallel orientations, similar fabrics can be found in soils and meta- lar to the least principal stress (03). A tensile fracture (mode I of Lawn and
morphic rocks and that both are probably caused primarily by deforma- Wilshaw, 1975) is formed if the least principal stress is tensile and the
tion of a mixture of platy and equant grains. crack opens perpendicular to the fracture plane. Tensile stresses may occur
in a compressive environment as a result of stress concentrations (Griffith,
Microlamination 1924) or by the increase of pore-fluid pressure to produce a tensile effec-
tive stress (Secor, 1965). Some fractures may form as mixed modes with
The fissility or property of splitting along approximately parallel both opening and shear displacement across the surface. A mode-II frac-
surfaces in a number of shales and slates correlates with the presence of ture (sliding mode, Lawn and Wilshaw, 1975) results from displacement
microlartiinations rather than grain fabric. The fissility of shale is signifi- of the crack walls parallel to the surface in a direction normal to the crack
cantly enhanced by fine-scale interbedding of clay with bands of organic front, the slip being analogous to the movement of an edge dislocation. A
material (Ingram, 1953; Gipson, 1965; Odom, 1967; Curtis and others, mode-Ill fracture (tearing mode, Lawn and Wilshaw, 1975) results from
1980, p. 336) or the presence or organic-rich and organic-poor clay layers displacement of the crack walls parallel to the surface and to the crack
(Spears, 1976). The fissility of slate may be caused by closely spaced front, the slip being analogous to the movement of a screw dislocation.
laminae of different compositions (Sorby, 1880, p. 73; Plessman, 1965,
p. 77). Fissility along such spaced cleavage (see discussion later) is more Microfractures
the result of compositional difference than grain orientations. Microlami-
nation fabrics are common, and sofissilityshould not be assumed to be the Microfractures are microscopic cracks that are confined to a single
result of grain orientation only. crystal, grain, or grain boundary or to only a few crystals, pains, or grain
boundaries. Cracks that are restricted to single grains imply a strong con-
Kinks and Crenulations trast in material properties between grain and matrix or grain and pore
space. Such cracks usually connect grain-to-grain contacts and thus have a
Asymmetrical kink bands and symmetrical crenulations occur as wide range of orientations within a single thin section (Gresley, 1895;
microscopic- to hand-specimen-scale deformation mechanisms in well- Gallagher and others, 1974, p. 234-236; Aydin, 1978, p. 920; Batzle and
foliated but unlithified clay and sand. The first report of kinked clay others, 1980, p. 7074-7085). Through-going cracks that continue across
appears to be the field examples and laboratory experiments of Tchalenko several grains without change in orientation (Reik and Currie, 1974,
(1968, p. 169) in which kinks occur in and adjacent to fault zones. His p. 1257; Aydin, 1978, p. 926) imply low contrasts in physical properties
experimental kinks were produced in a shear box in which the clay fabric between grains and matrix or grains and cement and require some degree
began parallel to the shear plane. During shear, the foliation rotated in the of prior cementation in porous rocks (Gallagher and others, 1974). Micro-
direction of the shear couple and then kinked with the short limbs rotating fractures that have healed without filling remain visible as planes of solid
in a sense opposite to that of the shear couple associated with faulting. This and fluid inclusions called "Tuttle lamellae" (Spry, 1969, citing N. Rast)
is evidently the same type of kink band that has been produced experimen- after Tuttle (1949), who thoroughly described this feature. Experimental
tally in card decks, slate (Gay and Weiss, 1974, p. 291, 294), and rubber work shows that high temperature (Lemmleyn and Kliya, 1960, p. 127)
strips (Honea and Johnson, 1976, p. 207; Reches and Johnson, 1976, and the presence of a reactive pore fluid greatly speed the healing process
p. 305, 307) where it is the result of unstable bedding-plane slip. In the and reduce the time required for crack closure in quartz (Smith and Evans,
experiments of Tchalenko (1968), the kink bands grew in width until all of 1984, p. 4128), implying that open microfractures are young and have
the clay foliation had the kink orientation, about normal to the maximum remained at relatively low temperature (Kowallis and others, 1987). Field
compressive stress within the fault zone. At an intermediate stage, the studies show that well-oriented microfracture sets are parallel to joint sets
structure looks like kinks with the opposite sense of asymmetry. Kinks (Tuttle, 1949; Bonham, 1957; Roberts, 1965; Friedman, 1969; Dula,
have also been produced experimentally in clay by Foster and De (1971, 1981). Filled microfractures that represent microveins may be invisible in
Fig. 13) and Maltman (1977, p. 430) and prove to be relatively common plane light or crossed polars. Microveins may stand out in cathodolumi-
in the deformed accretionary prism sediments of modern subduction zones nescence (Sippel, 1968; Sibley and Blatt, 1976; Sprunt and Nur, 1979;
(Lundberg and Moore, 1986, p. 22), as well as in other areas (Reches and Kanaori, 1986, p. 137-140; Narahara and Wiltschko, 1986, p. 161) re-
Johnson, 1976, p. 298). Van Loon and others (1984) illustrated natural vealing the presence of many more veins than anticipated. These micro-
centimeter-scale folds from unconsolidated but well-bedded sands that veins do not resemble Tuttle lamellae. Elongate grains that formed in
have the geometry of chevron folds (p. 355), kinks (p. 357,359,368), and a low-temperature environment may owe their shape to invisible filled
kinks with highly attenuated steep limbs (p. 358) resembling faults. The microfractures rather than to plastic strain (Narahara and Wiltschko,
latter geometry is characteristic of the early stages of displacement on faults 1986, p. 161).
in experiments with unconsolidated sand (Hubbert, 1951, PI. 2) and indi- Grain fracturing has been the most conspicuous result of compaction
cates the close connection between faults and certain folds having a kink experiments on quartz sands (Maxwell and Verrall, 1954; Borg and Max-
geometry. well, 1956; Maxwell, 1960, PI. 1; Borg and others, 1960, PI. 2). There has
Maltman (1977, p. 422), using a pure-shear experimental configura- been some discussion about whether microfractures, especially those con-
tion with clay, produced symmetric crenulations (p. 426, 430) and very fined to single grains, solitary fractures, or single fracture sets were initially
tight hinged folds that he called creases (p. 432). Crenulations are found caused by extension or shear (Borg and others, 1960). A wide range of
naturally in clay-rich accretionary prism sediments (Lundberg and Moore, orientations is typical (Borg and Maxwell, 1956). It is now generally
1986, p. 23) and in sedimentary rocks (Nickelsen, 1979, p. 258-259; accepted that the origin is extensile (Gallagher and others, 1974; Kranz,
Downloaded from gsabulletin.gsapubs.org on February 27, 2013
1334 R. H. GROSHONG, JR.

1983) and that shear offsets result from later movement, such as the sliding plane, a tension crack must intersect it at right angles (Anderson, 1942;
of a fragment into an adjacent pore, or from noncoaxial strain, as in a Lachenbruch, 1962). A tension fracture propagating at some other angle
fault zone. Brown and Macaudiere (1984, p. 580-581), however, show will curve to intersect a free face at 90° (Lachenbruch, 1962). Branching of
examples of conjugate microfractures within single crystals of plagioclase propagating tensile fractures has been observed in time-lapse photography
having the proper directions and offsets to establish that they are conjugate of developing mudcracks to cause a 120°-angle intersection (Anderson
faults. This pattern is quite different from the microfractures seen in a and Everett, 1965), an observation that demonstrates that large-angle
typical sandstone. The compressibility of porous materials increases with branching need not result from high-speed fracture propagation as sug-
grain size as a result of the greater stress concentrations at the smaller gested by Lachenbruch (1962). Surface striations on columnar basalt indi-
number of grain contacts present in coarse-grained materials (Borg and cate tensile fracture propagation. According to Ryan and Sammis (1978),
others, 1960; Friedman, 1967). the striations were first noticed by Iddings (1886) and first related to the
Open microfractures have attracted much attention in recent years fracture process by James (1920). A striation consists of a smooth and
because of their association with brittle faulting (Kranz, 1983) and as rough band resulting from the propagation and cessation of propagation of
possible porosity- and permeability-enhancing mechanisms. In experi- the growing tensile fracture (Ryan and Sammis, 1978, p. 1296-1299). The
ments, microfractures are produced by either mechanical or thermal stress concentration at the pre-existing fracture segment controls the posi-
stresses. Brittle macroscopic shear failure of low-porosity rocks is usually tion of the next fracture increment (DeGraff and Aydin, 1987). A small
preceded by distributed microfracturing that causes dilatancy, a volume component of shear parallel to the fracture (combined fracture modes I
increase (Brace and others, 1966). Significant increases in microfracture and II) causes relief on the fracture surfaces called "fracture lances" (Ryan
density along the trend of the fault occur just before faulting (Gramberg, and Sammis, 1978, p. 1300) which trend parallel to the fracture propaga-
1965, p. 40-41). The microfractures nucleate at a variety of stress concen- tion direction. Contraction cracks in soils may form as parallel sets or as
trations (Brace and others, 1966; Olsson and Peng, 1976, p. 55; Beeré, orthogonal systems (Lachenbruch, 1962; Brewer, 1964, p. 139) mimicking
1978, Pis. 1 and 2; Abdel-Gawad and others, 1987, p. 12914). Thermal systematic joints in rocks.
stresses cause microfracturing because of anisotropic thermal expansion of Syneresis fractures are extension fractures resulting from bulk volume
adjacent grains, a process that is enhanced by the presence of quartz reduction due to desiccation and are randomly oriented in three dimen-
because of its large and anisotropic coefficients of thermal expansion sions (Nelson, 1979, p. 2216). The fractures have a scale and spacing that
(Kranz, 1983). Some microfractures form upon stress release when a rock causes them to resemble chicken wire on a plane surface, and so they are
is removed from deep burial and can be closed by subjecting the sample to commonly called "chicken-wire fractures" (Nelson, 1979).
a confining pressure equivalent to the released stress (Wang and Simmons, Systematic Joints. These joints occur as a subparallel set (Hodgson,
1978). Stress-release microfractures may form parallel to existing joint sets 1961b). Nearly ubiquitous in surface exposures, they are readily visible on
(Carlson and Wang, 1986). Grain boundary microcracks seem to be a the scale of the outcrop, and a single joint may extend more than 400 ft
common result of surficial weathering and disappear in fresh rock (R. J. (Hodgson, 1961b). In outcrop and thin section, uncemented joints are thin
Kuryvial, 1976, personal commun.). cracks that cut straight through most grains (Nickelsen and Hough, 1967,
Measured microfracture porosities are in the range of 0.01%-2.4% PI. 6; Jamison and Stearns, 1982, p. 2594; Segall and Pollard, 1983a,
(Brace and others, 1966; Carlson and Wang, 1986; Abdel-Gawad and p. 568) or may, in part, follow grain boundaries (Nelson, 1985, p. 32).
others, 1987). Abundant weathering-related grain-boundary microfrac- Microscopic shear offsets of grain boundaries are not reported from thin
tures in a sandstone resulted in a porosity increase of zero to 1% and a sections of joints, indicating little if any shear displacement (Narr and
permeability increase that was usually negligible but ranged upward to 5 Burruss, 1984, p. 1091; Ramsay and Huber, 1987, p. 642).
millidarcys (measured under atmospheric conditions; R. J. Kuryvial, 1976, Systematic joints are characterized by their geometric relationships to
personal commun.). Microfracture porosity and permeability are much one another and general surface morphology. Cross joints (Hodgson,
less under elevated confining pressure because of crack closure, unless the 1961b, p. 18) extend between and are approximately normal to a system-
cracks are held open by surface irregularities (Batzle and others, 1980; atic joint set, have irregular surfaces, and commonly terminate on bedding
Nelson, 1985). surfaces and against other joints. Nickelsen and Hough (1967, PI. 4)
pointed out that this term had been used previously by Balk (1937) for
Joints joints perpendicular to lineation in igneous rocks and renamed them
"truncated joints." I prefer a more general terminology, calling the very
A joint is a fracture surface with no visible displacement parallel to planar joints "planar systematic" and the rougher-surface-textured joints
the surface (Bates and Jackson, 1980). It is a feature seen on the scale of an "rough systematic," because both types occur in sets, and all rough system-
outcrop or greater. Wise (1964) defined microjoints as macroscopic sub- atic sets are not necessarily perpendicular to a planar set. The standard
parallel fractures spaced closer than 3 mm apart. It seems likely that all deviation of a planar set in a single outcrop is small, on the order of 2.4°,
joints form as cracks; that is, they are cohesionless at their inception. Later so that 95% of the joints should fall within a range of 9° (Groshong, 1965;
events might cause slip parallel to the surface or cementation to form veins. Engelder and Geiser, 1980). Rough systematic joints may have rough
Cooling and Desiccation Cracks. Contraction due to drying causes surfaces or be rippled on a scale of centimeters; the orientation variability
mudcracks in shaly or clay-rich sediments (Neal and others, 1968; Kahle is usually twice or more that of a planar set. Hodgson (1961b) defined
and Floyd, 1971), and contraction due to cooling causes columnar joints nonsystematic joints as joints that display random rather than oriented
in basalts (Iddings, 1886; Peck and Minakami, 1968; DeGraff and Aydin, patterns in plan and section. He considered cross-joints to be a type of
1987) and in permafrost (Lachenbruch, 1962). When fractures propagate nonsystematic joint, an interpretation incompatible with their being per-
from a planar surface, two-dimensional polygonal arrays (Neal and others, pendicular to a systematic set. I consider that the best use of the terminol-
1968) to rectangular arrays (Lachenbruch, 1962) are developed. Fracture ogy is to define nonsystematic joints as random, not occurring in sets, and
intersection angles and, where visible, fracture surface markings indicate a as excluding truncated joints.
tensile fracture origin for these features. Typical fracture intersection angles The small-scale surface morphology of joints has played a major role
are 60°, 90°, and 120° (Lachenbruch, 1962; Peck and Minakami, 1968). in the interpretation of joint origin since the review by Hodgson (1961a).
A 90° intersection occurs where a propagating tension fracture meets the Forms are concentric ridges or ribs (Price, 1966, PI. 2; Syme Gash, 1971,
free surface of another open fracture. Because the free surface is a principal p. 351) and radial ridges and valleys called "hackle" or "hackle marks"
Downloaded from gsabulletin.gsapubs.org on February 27, 2013
LOW-TEMPERATURE DEFORMATION MECHANISMS 1335

(Price, 1966; Syme Gash, 1971, p. 351; DeGraff and Aydin, 1987, p. 607) relief on plumose markings. They interpreted Parker's system I joints to be
or "striations" (Bahat, 1979). Bahat (1979, p. 82; 1986a, p. 201) termed two independent sets, one overprinting the other, the major alternative to
the portion of a joint between concentric ridges that has a rough surface a the conjugate hypothesis. Parker (1969) replied that individual joints of
"hackle," a different use of the term. This will be here termed a "grainy system I could be found changing direction along strike from one trend to
surface" to avoid confusion. Plumose markings (Woodworth, 1896, Pis. the other, indicating contemporaneity. In the latest interpretation of system
1-6; Hodgson, 1961a, p. 494) are hackle marks that consist of small ridges I by Nickelsen (1979), Engelder and Geiser (1980), Engelder (1985), and
and valleys curving away from a central axis like the geometry of a feather. Engelder and Oertel (1985), there are two independent tensile joint direc-
The edges of joints with plumose markings, especially at bedding planes, tions formed in three different episodes. Deep in the undercompacted
may show what Hodgson (1961a) called a "fringe of small en echelon Devonian deltaic sequence (Engelder and Oertel, 1985), the more north-
fractures" that curve obliquely away from the main joint face and conse- westerly set (lb) is cut by a spaced cleavage that is contemporaneous with
quently are called "en echelons" by Bahat (1986a, p. 199-200; 1986b, the more northeasterly set (la). Higher in the sequence, set lb joints termi-
p. 185-186). The initiation point of the fracture may be a very smooth nate against la joints, indicating that set la is here older, and thus indicating
"mirror" (Bahat, 1979, 1986a, p. 202). Hodgson (1961b, p. 23) pointed a third episode of jointing controlled by residual stress (Evans and En-
out that "the interlocking nature of the plumose patterns on the opposed gelder, 1986) that produced joints parallel to those of the first episode.
faces also precludes transcurrent movement; no such movement can occur In summary, it appears that joints are extensional in origin and may
without obliterating the plumose pattern." Hodgson (1961b) believed that form at a variety of different times (Engelder, 1985, 1987) with orienta-
the plumose pattern arose from an extension fracture propagating along tions controlled by the modern stress field (Engelder, 1982), residual elas-
the direction of the plume toward the tip of the feather. Gramberg (1965) tic strain (Friedman, 1972; Reik and Currie, 1974; Evans and Engelder,
and Price (1966) cited experimental evidence that documents this interpre- 1986), original rock fabrics (Brewer, 1964, p. 329; Nelson and Stearns,
tation. The en echelon fractures in the fringe region may be the result of a 1977), older deformation fabrics (Engelder and Geiser, 1980), or topog-
consistent slight rotation of the tensile stress direction (Pollard and others, raphy (Bradley, 1963; Nelson, 1979). Apparent conjugate relationships of
1982) or may be mixed-mode fractures analogous to the fracture lances joint (not fault) sets are best explained by overprinting; shear interpreta-
seen on basalt columns (Ryan and Sammis, 1978; DeGraff and Aydin, tions based on pattern analysis (Parker, 1942; Hancock, 1985) need addi-
1987). Conchoidal ridges or ribs (Bahat, 1979, Pis. 3 and 4; Bahat and tional documentation of the origin of the features before being accepted.
Engelder, 1984, p. 302) have been termed "arrest marks" by Bahat and
Joints may have a significant effect on bulk porosity and permeabil-
Engelder (1984) because the same surface morphology is characteristic of
ity, depending on their degree of opening and spacing. Joint widths tend to
the start-stop propagation of fatigue fractures in the laboratory. Fracture
be in the range of 0.001-0.05 cm, although reduced by a factor of 10 to
propagation is in the direction of convexity of the conchoidal ridges. A
1,000 by confining pressures appropriate to hydrocarbon reservoir depths
grainy surface represents small-scale branching of a rapidly propagating
(Nelson, 1985), but may be wider by a factor of 10 to 100 if held open by
extension fracture (Bahat, 1986a, p. 201-203). Bahat and Engelder (1984,
partial vein filling (Lucas and Drexler, 1976, p. 129-130; Nelson, 1985,
p. 302) and Bahat (1987a) pointed out that different joint sets in the same
p. 51, 55). Maximum fracture porosity is usually less than 2% (Nelson,
area may be characterized by different plume types. Inferences have been
1985) but is 6% in the Monterey chert in California (Weber and Bakker,
made about fracture velocity based on plume type, but all of the features
1981). Not all fracture porosity or permeability is related to joints; faults
evidently occur on both fast- and slow-moving fractures. Plumes with
may be an important factor. Joint permeabilities range from nearly zero to
fringes are observed in the laboratory on rapidly moving fractures (Bahat,
400 millidarcys at the surface (Shuaib, 1973; Nelson, 1985) but tend to be
1979) but have been observed on mudcracks by Bahat (1979, PL 16) and
less than 300 millidarcys under reservoir confining pressures (Nelson,
are interpreted by Pollard and others (1982) as occurring on slowly mov-
1985), a value large enough to have a very significant effect on the fluid
ing fractures. Conchoidal arrest marks may occur on rapidly moving frac-
flow in low-porosity rocks. Permeability increases approximately as the
tures in association with seismic energy release (Ryan and Sammis, 1978)
cube of the width of the opening (Engelder and Scholz, 1981).
but represent a start-stop movement which implies that the growth of the
complete fracture is slow.
VEINS AND RELATED FEATURES
The characteristics of joint intersections provide additional informa-
tion about the origin of the fractures (Hancock, 1985, p. 447; Bahat, A vein is a relatively thin, normally tabular, rock mass of distinctive
1987b, p. 308-316). Hodgson (1961b, p. 14) pointed out that joints of the lithologic character, usually crosscutting the structure of the host rock
same set intersect by curving to end at right angles against each other. This (Dennis, 1967). This definition does not specify the origin of the vein
is a common, although not universal, phenomenon and indicates the inter- material; the filling could be from an external source such as an intrusive
section of a tensile fracture with a free surface (principal plane) as in the igneous vein or an epigenetic ore vein (Bates and Jackson, 1980), or could
propagation of mudcracks or basalt columns. Low-angle intersections pose be the result of in situ deformation, as in a fault zone. The distinction
a more difficult interpretation problem. between externally sourced quartz veins and fault zones containing in situ
Bucher (1920) proposed a shear origin for pairs of joint sets that have cataclasticfillingcan be difficult in the field and might require thin-section
a conjugate shear-fracture orientation. The Appalachian Plateau of New analysis.
York and Pennsylvania provides a classic example of the alternatives. In Dilation veins represent openings that arefilledby external materials.
an influential paper, Parker (1942) interpreted the New York Plateau as Active intrusions result from the forceful injection of material into exten-
having three sets ofjoints. Parker's Set I (better termed"system I") contains sion fractures. Passive intrusions are pulled into low-pressure zones created
two conjugate directions separated by a 16° angle. Parker's descriptions by the formation of extension fractures. Nondilational veins (Hobbs and
show the joints of system I to be planar systematic with abundant plumose others, 1976, p. 292) result from the alteration or replacement of the wall
markings. Parker interpreted these joints as shear fractures formed with a rock.
tensile <73 and concluded that the plumose markings must indicate a shear
origin (see also Roberts, 1961). Discussing the adjacent Pennsylvania Pla- Clastic Dikes
teau, Nickelsen and Hough (1967) reached the contrary conclusion that
the planar systematic joints are extension fractures because of their open- A clastic dike is a vein filled with sedimentary debris (Newson,
ness, lack of tangential movements, and the interlocking character of the 1903). Neptunian dikes (Hancock, 1985) are surface features such as
Downloaded from gsabulletin.gsapubs.org on February 27, 2013
1336 R. H. GROSHONG, JR.

mudcracks, filled with sediments from above (Heron and others, 1971). A special type of antitaxial vein has been called the "product of vein
Some Neptunian dikes may fill pre-existing fractures that have structural coalescence" by Misik (1971) and a "crack-seal vein" by Ramsay (1980b,
significance but are not related to thefillingepisode (Smith, 1952), where- p. 136-138). This vein type contains multiple screens of wall rock called
as other surface fractures form during an event that controls their "inclusion bands" (Ramsay and Huber, 1987, p. 576) marking successive
orientation, such as faulting (Vintanage, 1954; Harms, 1965), landsliding, openings. The "stretched crystal" veins of Durney and Ramsay (1973,
or glacial movement (Dionne and Shilts, 1974). Active or passive subsur- p. 77) are interpreted by Ramsay and Huber (1983) as a type of crack-seal
face sediment intrusions may extend upward or downward from the vein. Inclusion trails are isolated crystals in the vein that are generally of
source bed and are commonly oriented by the contemporaneous stress the same species and in optical continuity (Ramsay and Huber, 1987,
field in folds or adjacent to faults (Diller, 1890; Harms, 1965; Peterson, p. 576). Inclusion trails track the direction of extension and may be
1966; Plessman and Spaeth, 1971; Winslow, 1983). Clastic sills are also oblique to the fiber direction (Cox, 1987, p. 781,783), indicating that the
known (Bielenstein and Charlesworth, 1965; Truswell, 1972). Injected fibers do not necessarily parallel the opening direction.
dikes may show compositional layering, graded layering, and alignment of A composite vein (Durney and Ramsay, 1973, p. 76) has one min-
grains parallel to the vein walls (Diller, 1890; Harms, 1965; Peterson, eral species at the vein wall and another at the center (but no evidence of
1968, p. 180-187; Winslow, 1983, p. 1078). void-filling textures). Both mineral phases of the vein are interpreted to
have grown simultaneously at the phase boundaries within the vein (Dur-
Mineral-Filled Veins ney and Ramsay, 1973), a notably different interpretation from the com-
Dilation veinsfilledwith minerals precipitated from aqueous solutions mon assumption that the two phases represent two different episodes of
are of primary interest here. Mosaic-filled andfiber-filledtypes are recog- infilling. The chemistry of this has not been explained.
nized. Mosaic-filled veins are characterized by interlocking anhedra of Bedding-parallel veins are found in undeformed sedimentary rocks in
quartz (Adams, 1920) or calcite spar cement (Misik, 1971) either covering which the fibers are approximately perpendicular to bedding. The variety
an early fiber-type filling (druse) or perhaps completely filling the vein known as "beef' has parallel fibers (Marshall, 1982). The unusual mor-
(Groshong, 1975b, p. 1368; Spang and Groshong, 1981, p. 333). This is phology called "cone-in-cone structure" (Sorby, 1860; Marshall, 1982,
called a "rapid opening fabric" by Misik (1971) because the fracture p. 618-619) usually contains crystals that diverge at low angles to produce
obviously opened faster than it could befilled.This fabric is characteristic a cone-shaped cross section. The interpreted origin of both types is similar
of the lower temperatures of deformation. Fiber-veins arefilledwith single to that of tectonic fiber veins (slow opening of the vein with simultaneous
crystals that do not show prominent crystal-face boundaries and that are filling), but no explanation is given for the conical morphology which is
elongate oblique to the vein wall. Not uncommonly the outer margin of a absent in tectonic veins.
vein consists of fibers and a mosaic-filled center, suggesting an increase in The description of cone-in-cone structure superficially resembles that
opening rate (Misik, 1971) or a decrease in the production of the filling of a shatter cone. A shatter cone is a distinctively striated conical fragment
(Ramsay and Huber, 1987). An opening rate increase may coincide with a of rock (or fracture trace), ranging in length from less than a centimeter to
change from extension normal to the vein wall to oblique opening (Gro- several meters, along which fracturing has occurred; it is generally found in
shong, 1975b, p. 1368; Beach, 1975, p. 259). nested or composite groups (after Dietz, 1959, Pis. 1-8). Shatter cones are
Macroscopic stylolites normal to veins are common (Nelson, 1981, associated with high-velocity impact; the cones point toward the center of
p. 2421), and microstylolites may be found on fiber boundaries within impact.
veins (Beach, 1977, p. 217). Both relationships indicate shortening per-
pendicular to the length of the vein. Many examples have curved fibers. Recrystallization/Replacement Veins
The optic axis of a crystal comprising a fiber is typically straight, not bent,
and lacks evidence for large internal plastic deformation, and so the curva- Recrystallization/replacement veins show no evidence for separation
ture is interpreted to be the result of growth, not the later deformation of of the wall rock (such as offsets of bedding) and appear to be the result of
originally straight fibers. recrystallization or replacement of the host. MSik (1971) described two
In a paper that triggered a great expansion of research in this area, types: (1) veins with ghost fabrics in which the grain size of the host is
Durney and Ramsay (1973) subdivided fiber veins into three types based changed and (2) decolorizing veins. A decolorizing vein contains unaltered
upon the mode of growth: syntaxial, antitaxial, and composite (Ramsay host fabric that has been lightened in color, usually being nearly clear in
and Huber, 1983, p. 241). In a syntaxial vein (Durney and Ramsay, 1973, plane transmitted light (Misik, 1971, p. 452). In a vein with ghost fabric,
p. 71) the fibers show a break in optical continuity near the center of the the vein crystals no longer conform to the original host-grain fabric and
vein, interpreted to be the result of mid-point refracturing and crystal perhaps not even to the host-grain mineralogy, but undeformed ghosts of
growth from the wall to the vein center. This fabric, however, is also the original fabric remain (MSik, 1971, p. 455). Tectonically oriented
characteristic of some diagenetic pore-filling cements (Sandberg, 1985, ghost-fabric recrystallized veins have been recognized by Spang and Gro-
p. 35, 50; Pierson and Shinn, 1985, p. 160), where clearly a void is being shong (1981, p. 323) in the same rock with dilation veins; they have
filled, and so syntaxial veins may be of the void-filling type. Slightly more ill-defined boundaries in thin section yet are clearly recognizable as veins
drusy cement would cause the cement texture in Figure 1 to resemble that and are normal to the extension direction inferred from calcite twinning.
of a syntaxial vein. Where the mineralogy of the veinfillingand wall rock An analogous structure has been reported in plagioclase crystals by
is the same, the fibers are optically continuous with the wall-rock grains Hanmer (1981, p. T55-T57) where the vein orientation was at least partly
(Misik, 1971; Durney and Ramsay, 1973) and are usually free of wall- controlled by structure within the crystal.
rock inclusions. In an antitaxial vein (Durney and Ramsay, 1973, p. 73;
Ramsay and Huber, 1983, p. 241), thefibersare interpreted to have grown Vein Arrays
by refracturing at one or both of the vein edges. The fibers are optically
continuous across the vein but not optically continuous with the wall-rock Veins quite commonly occur in en echelon arrays, and two end-
grains and are likely to contain inclusions of the wall rock. Cox and member configurations are common: veins at nearly 45° to the array
Etheridge (1983, p. 154-155) show that such veins might contain fibers boundary and veins at 10°-20° to the array boundary (Hancock, 1972) or
grown syntaxially over the wall rock and might grow from one side only at even lower angles (Pollard and others, 1982, p. 1292) with a rather
(p. 155). The vein crystals may become wider in the direction of growth continuous range of examples in between. There are at least four docu-
due to the elimination of narrow fibers (Cox, 1987, p. 781). mented origins for single arrays. (1) Small-angle-change kink bands re-
Downloaded from gsabulletin.gsapubs.org on February 27, 2013
LOW-TEMPERATURE DEFORMATION MECHANISMS 1337

quire stretching parallel to bedding or dilation normal to bedding within formed rocks by Heard (1960, PI. 1). Heard found that Liiders' bands in
the kink band and consequently may be the sites of bedding-parallel arrays the Solenhofen Limestone occurred in the brittle-ductile transition region,
within the kink band (Anderson, 1968, p. 204-205) or approximately which is also characteristic of their occurrence in mild steel (Nadai, 1950).
bedding-normal arrays (termed "feather fractures" by Cloos, 1932, Experiments by Friedman and Logan (1973, p. 1466) resulted in two
1947b, p. 899; see also Garnett, 1974, p. 132). (2) Veins at 45° to the array conjugate sets with a dihedral angle of 75°-103°. The dihedral angle is
boundary are at exactly the angle expected for tension fractures if the bisected by the a\ axis, and the angle increases with confining pressure. In
maximum shear stress is parallel to the array boundary, such as in a shear both limestone and sandstone, the effect occurs in the brittle-ductile transi-
zone (Cloos, 1955, PI. 3). (3) Veins at a low angle to the array boundary tion regime but continues into the ductile regime in sandstone. The features
may be the result of the breakdown of a propagating planar vein that is are not faults, as they have zero offset, no gouge, and a dihedral angle that
parallel to the array boundary where the vein encounters a slight change in is at least 20° greater than the faults that form late in the same experi-
the direction of the maximum tensile stress (Pollard and others, 1982). ment (Friedman and Logan, 1973, p. 1466). Optical and SEM analyses by
(4) Veins oblique to a pre-existing surface may be produced as a result of Friedman and Logan reveal the bands to be zones of grain-size reduction
slip on the surface (Conrad and Friedman, 1976; Nemat-Nasser and Horii, (1973, p. 1470) or intense grain microfracturing (1973, p. 1473) with a
1982). width of about two original grain diameters. A natural example from
Conjugate vein arrays are common. The conjugate array boundaries sandstone (1973, p. 1475) shows prominent bands on outcrop that in thin
make angles of about 40° to more than 90° to one another, usually having section are zones of grain fracture without offset. Liiders' bands in rock
the configuration of conjugate faults (Shainin, 1950, PL 1). Roering (1968) thus appear to be arrays of extensional microfractures.
recognized and Beach (1975) defined two categories of conjugate array:
(1) conjugate zones in which the undistorted portions of the veins in one Pressure Shadows and Pressure Fringes
zone are not parallel to those in the other zone and (2) conjugate zones in
which the undistorted portions of the veins in both zones are parallel. In a Pressure shadows and pressure fringes are veins, fillings, and relict
type 2 zone, the veins are all parallel to the bisector of the acute angle materials found at the opposite ends of relatively rigid grains or fossils. In a
between the zones (Beach, 1975, p. 248; Ramsay and Huber, 1987, pressure shadow (Spry, 1969), the foliation of the host rock wraps around
p. 629). At the intersection of type 1 conjugate arrays, there is commonly a the boundary of the shadow, and the grains in the shadow are unoriented
vein or veins parallel to the conjugate angle bisector (Choukroune and or may be relics of the host rock. Since the work of Spry (1969), Chouk-
S6guret, 1968, p. 243), but veins of opposite trends may cross in this area roune (1971), and Durney and Ramsay (1973), a great deal of attention
(Shainin, 1950, PI. 1). It is generally agreed that the far-field ay axis bisects has been paid to pressure fringes, in which the foliation of the host rock
the acute angle between the vein arrays (Hancock, 1985), but the exact abuts the boundary of the fringe, and the fringe contains oriented mineral
cause of the vein geometries and especially the evolution of the geometry growths (fibers) according to the definition of Spry (1969). These are
are matters of controversy. well-known fabrics in low-grade metamorphic rocks, but pressure shad-
Vein arrays have the geometry of shear zones. Displacement parallel ows have been recognized in deformed soft sediments (Knipe, 1986,
to the array boundary should rotate the veins (Ramsay and Graham, p. 80), and short pressure fringes have been recognized in deformed
1970) leading to a sigmoidal vein shape and implying large strains in the sedimentary rocks (Fellows, 1943, PL 7; Geiser, 1974, p. 1404; Spang and
wall rock between the veins (Nicholson and Pollard, 1985). Large wall- others, 1979, p. 1111; Ramsay and Huber, 1983, p. 267). Such pressure
rock strains are usually indicated in the field by pressure solution within fringes may be much more common than would be inferred from the
the array in excess of that found outside the array (Beach, 1974; Rickard paucity of published descriptions. Rutter (1983, p. T31) appears to have
and Rixon, 1983). Later veins that are parallel to the original trend and been the first to produce a pressure fringe experimentally.
that obliquely crosscut the earlier wider veins have been observed (Shai- Durney and Ramsay (1973) subdivided pressure fringes in a fashion
nin, 1950, p. 516; Durney and Ramsay, 1973; Rickard and Rixon, 1983, analogous to their classification of veins. A syntaxial or pyrite-type (Ram-
p. 574), supporting the concept of vein rotation. Type 2 array veins may say and Huber, 1983) fringe is interpreted to have grown from the
open perpendicular to the vein wall without significant rotation. The sig- fiber/host-rock boundary. Syntaxial fringes are characteristically seen on
moidal vein shapes can be caused by tip interactions between growing en pyrite and are found where there is a strong contrast in mineral species
echelon cracks (Pollard and others, 1982, p. 1299) or by the formation of between the object being overgrown and the host rock, such as between a
curved country-rock bridges between dilating planar veins (Gorlov, 1971, pyrite grain and argillite host. A more complex form of pyrite-type fringe is
p. 149; Beach, 1975, p. 256; Nicholson and Pollard, 1985, p. 584-585). called "face controlled" by Ramsay and Huber (1983, p. 269), in which
The angle between conjugate type 2 arrays is explained by Ramsay and the fibers are oriented perpendicular to the pyrite face and for which the
Huber (1987, p. 629) as being related to the dilation; a small conjugate extension direction is oblique to the fiber trend. Inclusion trails and the
angle implies volume gain, and a large conjugate angle implies volume loss boundaries between fibers growing from adjacent faces mark the true
associated with pressure solution. If the differential stress is small, it is extension direction. An antitaxial or crinoid-type (Ramsay and Huber,
possible that some veins will form as conjugate shears of very low dihedral 1983) fringe grows from the rigid-grain/fiber boundary. This form is
angle (Muehlberger, 1961) and open obliquely (Hancock, 1972). Veins common on crinoids and is seen where the object and the host material
that show evidence for shear offset of external markers and irrotational have the same mineralogy. A composite fringe has more than one mineral
oblique opening based onfibersmight have this origin (Hancock, 1972, PI. present, and growth occurs at the phase boundary.
4; Beach, 1977, p. 204-205; Burg and Harris, 1982, p. 353; Cox, 1987).
Methods for computing the strain associated with en echelon veins are Boudins
given in Ramsay and Huber (1983,1987) and Collins and DePaor (1986).
The original boudins (Lohest, 1909) are in a quartzite between schist
Liiders' Bands beds and are barrel shaped in cross section with veins forming the top and
bottom of the barrel. Bedding-normal veins, regularly spaced at about one
Liiders' bands (or lines) were originally described from deformed soft or two times the thickness of the bed in which they occur and restricted to
steel by Liiders in 1854 according to Nadai (1950). They are seen on the the bed (Cloos, 1947a, p. 626), provide a mechanism for the formation of
surface of a sample as conjugate sets of prominent parallel lines in planes boudins in the low-temperature deformation of rocks. In his review, Cloos
of high shear stress. They were first described from experimentally de- (1947a) pointed out that investigators generally agreed that boudins were
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

1338 R. H. GROSHONG, JR.

caused by extension of a brittle layer between more ductile layers but most to least soluble as (1) halite and potassium salts; (2) calcite;
noted that the deformation implied by the barrel shape was an unresolved (3) dolomite; (4) anhydrite; (5) gypsum; (6) amphibolite and pyroxene;
problem in brittle rocks. The barrel shape results from strain in the stiffer (7) chert; (8) quartzite; (9) quartz, glauconite, rutile, and hematite; (10)
layer (Ramberg, 1955) and has been produced in rock experiments of a feldspars and cassiterite; (11) mica and clay minerals; (12) arsenopyrite;
more brittle layer by cataclastic thinning (Gay and Jaeger, 1975b). Richter (13) tourmaline and sphene; (14) pyrite; (15) zircon; and (16) chromite. It
(1963, p. 245) and more recently Mullenax and Gray (1984, p. 65-69) seems possible that local pore-fluid chemistry could alter the relative solu-
have shown examples in limestone in which the barrel shape is due to bilities, and so this order might not always be followed. It should be noted
differential removal of material by pressure solution. that an unreactive pore fluid such as liquid hydrocarbon prevents pressure
The opening of vein sets of more than one orientation can lead to the solution in the laboratory (Griggs, 1940) and in the field (Dunnington,
chocolate tablet form of boudinage (Wegmann, 1932). The veins may 1967) and inhibits precipitation (Hawkins, 1978) as well. Trurnit
open at different times, in different directions, and at different rates leading (1968) found that sutured contacts are characteristic of stylolites with ma-
to complex fiber patterns (Casey and others, 1983; Ramsay and Huber, terials of equal solubility on both sides (p. 97-99) and smooth contacts
1983, p. 256-257). Burg and Harris (1982) demonstrated that in a num- (p. 104-105) are characteristic of unequal solubility materials where only
ber of examples the vein arrays developed simultaneously, oblique to the the more soluble material dissolves (Morawietz, quoted by Trurnit, 1968).
extension axis. They interpreted the deformation to begin with necking A thick zone of insoluble residue acts like a more insoluble solution
due to the formation of Liiders' bands having the appropriate orientations, partner and favors smooth solution contacts (Trurnit, 1968). Heald (1955)
followed by through-going extension fracture parallel to the bands and observed that the amplitude of, and spacing between, sutured stylolites was
vein formation. greater in quartzose sandstones than in calcareous and argillaceous sand-
stones. Solution cleavages are much more abundant in argillaceous sand-
STYLOLITES AND SLICKOLITES stones or carbonates than in purer lithologies. The rate of pressure solution
is enhanced by the presence of clay (Heald, 1956; Oldershaw and Scoffin,
Pressure solution is defined here descriptively as the process by which 1967; Thomson, 1959; Whisonant, 1970; Mossop, 1972), and by the
material is removed by solution or diffusion from along a discrete surface, presence of good solvents (Gratier and Guiguet, 1986).
the sides of which remain in close contact. This contrasts with vuggy or Mechanical indentation of a stiff material into a softer material was
cavernous solution in which a void is created. Small voids may occur proposed as a cause for pebble indentation by Gresley (1895) and revived
along pressure-solution surfaces (Spang and others, 1979, p. 1113; Tada by Deelman (1975, p. 23) as a major alternative to pressure solution.
and Siever, 1986), but large voids represent a later opening event (Wong Mechanical indentation implies large strain in the deformed material, re-
and Oldershaw, 1981, p. 518). A stylolite is a thin seam or contact surface sulting in internal shape perturbations, visible as fractures or crystal-plastic
that is interlocking by mutual interpénétration of the two sides, or irregular deformation features (Gresley, 1895; Fruth and others, 1966; Gay and
or smooth (Logan and Semeniuk, 1976). Sorby (1863) was the first to Jaeger, 1975a, p. 315; McEwen, 1981, p. 33). Stockdale (1922, p. 56-57)
recognize stylolites as being formed by the mechanism of pressure solution. pointed out that the preservation of undistorted bedding adjacent to the
Stylolites can be subdivided on the basis of their cross sections into colum- stylolite precluded this interpretation, and McEwen (1977, p. 249-250)
nar, peaked, irregular, hummocky, and smooth (Logan and Semeniuk, provided additional examples.
1976, p. 16). A stylolitic surface is nearly always marked by a seam of the Stylolite and solution cleavage surfaces are believed to form perpen-
relatively insoluble or slowly diffusing components of the adjacent rock. dicular to ai (Sorby, 1908; Weyl, 1959; Robin, 1978; Fletcher and Pol-
Stylolites are common in carbonates (Stockdale, 1922, p. 8-12; Logan and lard, 1981; Green, 1984) and are regularly observed to be at least
Semeniuk, 1976, p. 17-20) and siliciclastics (Heald, 1955; 1956, Pis. 1-4) approximately perpendicular to the axis of maximum shortening strain
but have also been found in rhyolite (Golding and Conolly, 1962, (Nickelsen, 1972; Geiser, 1974; Groshong, 1976; Engelder, 1979; Onasch,
p. 535-537), pegmatite (Bailly, 1954), chert (Glover, 1969; Cox and 1983a, 1983b).
Whitford-Stark, 1987), and granite (Burg and Ponce de Leon, 1985, The direction of shortening across a stylolite is parallel to the axis of
p. 434). Stylolite formation occurs under conditions ranging from the columns where this feature is present. Stockdale (1922, p. 39) observed
diagenesis to at least the lower greenschist grade of metamorphism (Stock- that in square-wave columnar stylolites the caps of the columns contained
dale, 1922; Park and Schott, 1968; Mimran, 1975; Kerrich, 1978; Wong thick seams of residue, whereas the sides of the columns contained much
and Oldershaw, 1981; Groshong and others, 1984a; Engelder and Mar- less material and were slickensided, indicating pressure solution across the
shak, 1985). caps and slip parallel to the column axis. The plane of a stylolite may curve
Stylolites may either occur at grain-to-grain contacts or be transgran- 20°-30°, especially to follow some compositional heterogeneity of the
ular, having lengths of tens of meters or more. Stylolites that are parallel to rock, but the columns remain parallel (Stockdale, 1922, p. 44-54;
bedding may be the result of burial diagenesis (for example, Wong and Arthaud and Mattauer, 1969; 1972, p. 13).
Oldershaw, 1981) or vertical tectonics (Dunnington, 1967, p. 347; John- Stylolites that are parallel to bedding usually nucleate along contacts
son and Budd, 1975). Transgranular stylolites oblique to bedding were first of contrasting lithology. Stylolites oblique to bedding may also nucleate at
recognized as being the result of structural deformation by Blake and Roy lithologic contrasts, such as the edges of worm tubes and burrows (Nick-
(1949, p. 784; see also Rigby, 1953, p. 268; Choukroune, 1969, p. 66; elsen, 1972, p. 109). Cleavage patterns observed by Geiser and Sansone
Arthaud and Mattauer, 1969; Plessman, 1972, p. 336) and may be referred (1981, p. 282) led them to propose that cleavage followed pre-existing
to as tectonic stylolites (Jaroszewski, 1969). Smooth or gently curved to joints. Residue in stylolites that does not have the composition of insoluble
anastomosing stylolites oblique to bedding are the major cause of spaced residue of the host rock caused Laubscher (1980, personal commun.) to
cleavage and may be referred to as solution cleavage (Alvarez and others, suggest that some stylolites in the Jura Mountains formed on older, filled-
1978, p. 364). joint surfaces. Orientation control by pre-existing discontinuities should be
The surface shape of a stylolitic contact is affected by the relative obvious if the stylolite columns are oblique to the stylolitic surface but may
solubilities of the minerals on opposite sides of the contact and by the not be obvious for irregular and smooth stylolites. Stylolites at a low angle
amount of insoluble residue present. Relative pressure solubilities have to pre-existing discontinuities occur (Barrett, 1964) but seem rare.
been determined from the minerals found within the insoluble residue and The response of a stylolite to a noncoaxial strain path has not been
from differential grain indentation, summarized by Trurnit (1968) from firmly established. Irregular to smooth stylolites or solution cleavages
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

LOW-TEMPERATURE DEFORMATION MECHANISMS 1339

might allow pressure solution to occur on planes oblique to o\, as sug- p. 613) and Gratier and Guiguet (1986, p. 852) have produced excellent
gested by the existence of oblique columnar stylolites. Solution cleavages examples. Stylolitic-like contacts in a rock containing grains and cement
not perpendicular to the axis of shortening have been observed by Borra- might represent the result of pressure solution or might be irregular bound-
daile (1977) and Spang and others (1979) and can be attributed to a aries between cement overgrowths (Land and Dutton, 1978). Evidence of
noncoaxial strain path. Stylolitic laminae may grow by the asymmetric pressure solution is provided by truncation of original contacts of fossils of
accretion of material (Oertel, 1983, p. 435). The final stylolite surface known original shape, such as ooids (Ramsay, 1967, p. 196; Logan and
reflects the strain history but is not necessarily perpendicular to the axis of Semeniuk, 1976, p. 24) or by truncation of dust rings around sand grains
finite shortening. (Heald, 1955, PI. 3; 1956, PI. 2; Rittenhouse, 1973, Fig. 4). Differential
The currently most satisfactory macroscopic description of stylolite cathodoluminescence of grains and cement facilitates the distinction be-
formation and evolution is the anticrack model, first stated by Durney tween stylolitic and overgrowth contacts (Sibley and Blatt, 1976,
(1974) and named and developed by Fletcher and Pollard (1981). Accord- p. 885-886; Houseknecht, 1987, p. 636). Where the radii of curvature of
ing to this concept, a stylolite is mechanically identical to a crack that the grains in contact differ, all other things being equal, small grains indent
shortens perpendicular to its surface by the removal of material. This leads large grains (Trurnit, 1968, p. 102-105). Stylolitic contacts between grains
to a stress and strain distribution around the stylolite identical to that of an tend to be approximately parallel to one another (Heald, 1956, PI. I;
opening crack but with opposite sign. The model suggests that stylolites Ramsay, 1967, p. 196, Houseknecht, 1987, p. 636) but with a significant
should propagate by growing at the tips. Natural examples show volume variability in orientation, as expected from the fact that not all grain
strains along the stylolite and plastic strain adjacent to the stylolite that fit contacts are normal to the bulk shortening direction (analogous to the
the model prediction (Tapp and Cook, 1988). The rotation of stress on control on microfracture orientations; Gallagher and others, 1974). A large
fold limbs causes curvature of the cleavage in perfect conformity to the amount of grain-to-grain pressure solution results in what Wanless (1979)
anticrack model (Groshong, 1975a, p. 411; Tapp and Wickham, 1987). In called a fitted fabric in which nearly all grain contacts are stylolitic (Bux-
the kinematic model of stylolite formation by Guzzetta (1984), a sutured ton and Sibley, 1981, p. 25).
stylolite geometry is obtained if the side from which material is removed Grain-to-grain pressure solution reduces pore space by compaction as
changes along the surface. If the side from which material is removed material is removed and also by cementation if the material is deposited in
remains the same through time, a square-wave stylolite evolves; but if the the nearby pore space (Rittenhouse, 1971b). This process, which may be
side changes, then the peaked form evolves. The regular spacing of stylo- called "chemical compaction," is now recognized as being very important
lites and solution cleavage was addressed by Merino and others (1983), in the field (Choquette and James, 1987). Experiments on sands have
who proposed a theory for monomineralic rock in which initial porosity resulted in porosity reductions of 45%-70%, with the greatest reduction in
variations localize pressure solution and set up an instability that controls the finest grained material (Renton and others, 1969; Sprunt and Nur,
the spacing between stylolites or cleavages; spacing is controlled by poros- 1976,1977), a relationship also observed in thefield(Houseknecht, 1984).
ity, grain size, temperature, stress, mineral compliance, molar volume, and Mimran (1977) demonstrated tectonic control of the process by showing
the dissolution rate constant. that in a naturally deformed chalk the bulk density increase due to
The exact mechanism(s) by which pressure solution takes place is a pressure-solution-related porosity loss is directly related to dip.
subject of some controversy as shown by the reviews of Elliott (1973),
Paterson (1973), de Boer (1977a), McClay (1977), Durney (1978), Ker- Transgranular Stylolites
rich (1978), Rutter (1983), and Green (1984). It seems to be generally
agreed that solid-state grain-boundary diffusion is too slow a process to Stylolites that traverse many grains are here termed "transgranular
allow significant strain to occur at geological strain rates (Fletcher and stylolites." Park and Schott (1968) used the term "aggregate stylolites" for
Hofmann, 1974; Rutter, 1983; Green, 1984). Pressure solution has been stylolites having an amplitude larger than the grain size in which they
inferred in water-wet experiments where large bulk strain has been pro- occur. In a rock lacking matrix, this type of stylolite appears to require the
duced without cataclastic or plastic strain of the crystals (Griggs, 1940; deprior existence of cement, otherwise there would be only grain-to-grain
Boer, 1977b; de Boer and others, 1977; Sprunt and Nur, 1977; Urai, stylolites. Tectonic stylolites were first used as a tool for structural analysis
1985). The two most likely models are diffusion in a water film adsorbed by Lindstrom (1962). He was soon followed by Price (1967) and Plessman
along the grain boundary (Weyl, 1959; Robin, 1978; Green, 1984) and (1972).
undercutting by free surface dissolution at grain margins followed by grain Much of our quantitative understanding of transgranular sutured
crushing (Bathurst, 1958; Weyl, 1959). The experiments by Tada and stylolites traces to the exceptionally thorough work of Stockdale (1922,
Siever (1986) demonstrate the first model for halite; the experiments of 1926) on limestones from the United States craton. He showed that the
Gratier and Guiguet (1986) on quartz sand are interpreted by them in material in the stylolite seam had the same mineralogical composition,
terms of the second model. Movement of the pore fluid, even in very low oxide ratios, and percentage of organic matter as the insoluble residue
porosity and permeability rocks, will substantially speed up the process component of the adjacent limestone, thereby establishing that the seam
(Fletcher and Hofmann, 1974). The term "solution transfer" (Durney, material is insoluble residue. Schwander and others (1981) have verified
1972) might be used where movement of pore fluid is important and the this conclusion but have also identified host material in the stylolite. In
term "diffusion transfer" where the pore fluid does not move. addition, Stockdale showed that the thickness of the seam varies directly
with the amplitude of the columns and inversely with the purity of the
Grain-to-Grain Stylolites limestone. These relationships imply that a stylolite begins as a planar
surface and that the amplitude of the suturing increases as more material is
Grain-to-grain stylolites have been observed in a variety of clastic removed (Stockdale, 1926, p. 402). The first experimental example of this
rock types of various grain sizes and imply that the rock is grain supported process has been illustrated by Gratier and Guiguet (1986, p. 852). In my
(Dunham, 1962) or that the contrast in physical properties between grains experience, where stylolites are found along lithologic contacts, the separa-
and matrix (if present) is large. Grain contacts may be sutured (Heald, tion of lithologies to opposite sides of the stylolite is always perfect. If a
1956, PI. 1; Thomson, 1959, p. 97) or smooth (Trurnit, 1968, p. 104). The stylolite began at random centers of solution that joined along small faults
first experimental deformation of a granular aggregate to show a grain-to- to produce the columns, there should be examples where the offset of the
grain stylolite was by de Boer (1977b, Fig. 5); Shinn and Robbin (1983, lithologic contact is less than the amplitude of the stylolite, something I
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

1340 R. H. GROSHONG, JR.

have yet to see. In some homogeneous units, pressure solution may be the Slickolites
cause of the bedding (Simpson, 1985). Stylolites may also be initiated
along irregular boundaries formed during compaction (Shinn and Robbin, A slickolite, a term coined by combining slickenside and stylolite,
1983). Stylolite amplitude provides a minimum estimate of the material was originally discovered in association with shallow substratal solution in
removed because there may be a component of pressure solution that is limestone (Bretz, 1940, p. 354-355). According to Bretz (1950), a slicko-
equal on both sides of the stylolite that does not produce relief on the lite consists of alternating grooves andridgesor half columns which fit into
contact. Stylolite width provides a maximum estimate because an un- ridges and grooves opposite them, the series on one face being the exact
known amount of residual-composition material may have been present reverse of the marking on the opposite face. Ends of columns are engaged
previously, for example, clay on a bedding plane. Another technique for by sockets in the termini of grooves, and the column-socket relationship on
estimating material removed by pressure solution uses the offset of planar one face is the opposite of that in the other. Bretz (1940) pointed out that
markers such as beds, veins, or large fossils. Perhaps first noted by Cony- no clayfilmsare associated with slickolites, an observation that appears to
beare (1949, p. 84), pressure solution oblique to a linear marker produces me to be generally valid, although small amounts of residual material may
an offset of the marker without slip parallel to the stylolite. The amount of be found. The slickolites observed by Bretz (1940, 1950) were mainly
offset is proportional to the angle between the marker and the stylolite and vertical, downthrown toward the center of zones of solution collapse, but
the amount of material removed. occasionally found on bedding planes and were interpreted to have formed
The total amount of material removed along stylolites may be large. under about 55 ft of overburden. Jaroszewski (1969, PI. II), Carannante
Stockdale (1922, p. 39) illustrated a square-wave stylolite with an ampli- and Guzzetta (1972), and Laubscher (1979, p. 472) called attention to the
tude of 13 inches. Stockdale (1926) reported stylolitic bed thinning of role of slickolites in structural deformation, although Laubscher did not
13%-34%, Dunnington (1967) reported 20%-25%, and Johnson and Budd use the term.
(1975) provided data that indicate a 15% bed-thickness reduction. Hor- Slickolites evidently form parallel to the displacement direction. This
tenbach (1977) determined an increase in percent pressure solution with is consistent with the vertical orientation and vertical displacement on the
depth of burial. Closed systems, from which solutions cannot escape, slickolites observed by Bretz (1940,1950), with the general lack of resid-
quickly reach thinning values of 4%-6% and remain constant, whereas ual material on the stylolite and with the right-angle relationship between
open systems show a linear thinning up to 30%. Hortenbach suggested that conventional stylolites and the "dextral stylolites" illustrated by Laubscher
the pressure solution thinning can be used to estimate maximum depth of (1979, p. 472). Blake and Roy (1949) and Jaroszewski (1969, p. 21)
burial. Cretaceous chalk folded over a salt dome has been shown by reported a continuous gradation with changing orientation from normal
Langheinrich and Plessman (1968) to have been thinned along stylolites stylolites to slickolites, and the same gradation is implied in the illustra-
in an amount up to 28% and to have a porosity that correlates inversely tion of stylolites and oblique stylolites by Arthaud and Mattauer (1969,
with the amount of stylolitic bed thinning and the dip of bedding. Restora- p. 739).
tion of a fold that has deformed primarily by pressure solution has demon-
strated a minimum pressure-solution area loss in the hinge of 18% CLEAVAGE
(Groshong, 1975a). In a fold examined by Droxler and Schaer (1979),
10%-20% of the original volume was pressure-solved and in part reprecipi- Cleavage includes all types of secondary planar parallel fabric ele-
tated in nearby veins. Large amounts of pressure solution on irregular to ments other than coarse schistosity which impart mechanical anisotropy to
smooth bedding-parallel stylolites can produce nodular bedding in which the rock without apparent loss of cohesion (Dennis, 1967). Spaced cleav-
relatively unaffected nodules are surrounded by solution residue (Richter, age, defined as having cleavage surfaces spaced at finite intervals, however
1963; Logan and Semeniuk, 1976; Garrison and Kennedy, 1977, small (Dennis, 1967), is the typical fabric. A rock having spaced cleavage
p. 115-123; Wanless, 1979, p. 442, 448). In relatively heterogeneous is divided into cleavage laminae and narrow slices of the original rock
lithologies, the result of extensive pressure solution may be a stylobreccia called "microlithons" (de Sitter, 1956).
(Logan and Semeniuk, 1976, p. 44,47).
In otherwise high-porosity rocks, stylolites are typically associated Solution Qeavage
with zones of low porosity. Very low porosity stylolitic zones in high-
porosity reservoirs segment the reservoirs. A single stylolite affects a zone The quantitative importance of pressure solution in the formation of
from 0.5 ft (Wong and Oldershaw, 1981, p. 513) to about 5 ft in width to slaty cleavage was first demonstrated by Plessman (1965) in the slate of
either side (Dunnington, 1967, p. 347; Johnson and Budd, 1975, p. 16, the Rheinisches Schiefergebirge. Plessman showed that the cleavage lami-
22). It is commonly assumed that the lack of porosity is the result of local nae were zones of pressure-solution residue with grains and fossils in the
deposition of pressure-solved material, although Nelson (1983) proposed microlithons truncated against the cleavage by pressure-solution removal
that zones of low porosity control the location of the stylolites. Buxton and (1965, p. 75). Offsets of bedding along cleavage were shown by Plessman
Sibley (1981, p. 25) found greater porosity in a zone of fitted texture (1965) to be the result of the amount of material removed. These observa-
adjacent to a transgranular stylolite than in the adjacent relatively unstylo- tions have been duplicated by Roy (1978, p. 1777) and repeated in other
litic rock, showing that the pressure-solved material is not necessarily slates by Williams (1972, p. 10), Groshong (1976, p. 1141), and Bell
deposited in the zone of pressure solution. Volume balance and chemical (1978, p. 186-187). The same origin for more widely spaced cleavages has
changes contemporaneous with pressure solution suggest that nearby re- been demonstrated for sandstone and mudstone (Nickelsen, 1972, p. 110;
deposition is a general if not universal rule. Wanless (1979, 1982) found Geiser, 1974, p. 1404; Gray, 1978, p. 579), limestone (Nickelsen, 1972;
dolomite growing adjacent to stylolites where magnesian calcite was being Groshong, 1975b, p. 1367; Alvarez and others, 1976, p. 699; Spang and
pressure-solved, a fact he attributed to pressure solution of the impure others, 1979, p. 1111), and dolomite (Schweitzer and Simpson, 1986,
calcite and precipitation as dolomite. This is called "incongruent pressure p. 782-783). A large number of examples are given in the compilation
solution" by Beach (1979). Cements contemporaneous with pressure solu- edited by Borradaile and others (1982).
tion have been observed experimentally (de Boer and others, 1977). On Many fabrics that had previously been termed "fracture cleavage" are
the other hand, changes in oxygen isotopes associated with the pressure now seen to be solution cleavages that break easily along the cleavage
solution compaction of a chalk (Mimran, 1977) suggest that some of the laminae. In many fabrics previously termed "slip cleavage," the offsets
precipitated calcite came from outside the chalk bed. along cleavage can be interpreted in terms of pressure solution normal to
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

LOW-TEMPERATURE DEFORMATION MECHANISMS 1341

axial plane cleavage (Plessman, 1965, p. 81), normal to fanning cleavage Large strains may be associated with solution-cleavage formation. In
(Groshong, 1975a, p. 413) or pressure-solution removal of fold limbs in some examples, the presence of syntectonic veins and pressure fringes
certain crenulation cleavages (Williams, 1972, p. 19; Gray, 1977b, p. 234). suggests that some or all of the material is deposited nearby (Geiser, 1974;
Pressure solution oblique to the solution laminae is also possible as a cause Mitra, 1976; Clendenen and others, 1988), but other examples show no
of offsets parallel to cleavage (Groshong, 1975a; Borradaile, 1977). Rela- evidence of extension of any type, and so the strain must be associated
tively widely spaced cleavages appear to be transitional to closely spaced with a volume loss (Wright and Piatt, 1982; Beutner and Charles, 1985).
slaty cleavages with increasing strain, depth of burial (Engelder and Mar- Shortening strains of up to 50% have been reported by Plessman (1965),
shak, 1985), and increasing clay content of the host rock (Marshak and Alvarez and others (1978), and Wright and Piatt (1982); 59% was re-
Engelder, 1985). ported by Beutner and Charles (1985), much or all of which can represent
The descriptive classification of solution cleavages parallels that volume loss.
of stylolites in having stylolitic (sutured), smooth, wavy, and anasto-
mosing geometries (Powell, 1979; Borradaile and others, 1982; Engelder Pencil Cleavage
and Marshak, 1985; Schweitzer and Simpson, 1986). The term "rough
cleavage" (Gray, 1978, p. 579) is used for the irregular subplanar spaced The term "pencil structure" was used as early as 1858 by Naumann
cleavage typical of psammitic rocks. The same cleavage plane may be for a lineation in gneiss and later for the cleavage-bedding intersection
both stylolitic and smooth as it passes through beds of different lithology lineation in slate producing pencil slate (Cloos, 1946, p. 8). Graham
(Helmstaedt and Greggs, 1980, p. 105, 107). Cleavage domains that (1978) and Engelder and Geiser (1979) introduced the term "pencil cleav-
cannot be resolved with an optical microscope are called "continuous," age" to replace what Crook (1964) called "reticulate cleavage." Crook
as is cleavage resulting from platy minerals evenly distributed throughout (1964, p. 527) described the fabric as quasi-planar fractures that are pene-
the rock (Powell, 1979). trative on the scale of hand specimen and that result (especially after
Grain-orientation fabrics associated with solution cleavage are weathering) in a characteristic mass of acutely terminated elongate po-
explained by pressure solution removal and by recrystallization. Solution lygonalfragments.Pencil cleavage is found in shale, mudstone, and siltstone
cleavages may show large grains in the microlithons elongated parallel to (Crook, 1964, p. 526; Engelder and Geiser, 1979, p. 463; Reks and Gray,
cleavage. The grains are usually dimensionally oriented but not necessarily 1982, p. 165-168).
crystallographically oriented. This is explained by removal of the grain Pencil cleavage on the Appalachian Plateau (Engelder and Geiser,
margins adjacent to the cleavage laminae by pressure solution; the width of 1979) shows almost no evidence of the insoluble residue expected for
the grain normal to cleavage is reduced, but its length parallel to cleavage solution cleavage but can be traced into solution cleavage in interbedded
remains unchanged (Williams, 1972, p. 10; Lisle, 1977; Beutner, 1978, limestones. The pencil cleavage in the inner Valley and Ridge province of
p. 11). In many examples, the rock within the microlithons is strained, as the Appalachians is caused by very thin, disconnected solution cleavage
shown by the presence of pressure fringes on rigid grains within the micro- laminae that become better connected as the cleavage becomes more
lithons (Geiser, 1974, p. 1404; Roy, 1978, p. 1779; Gray and Durney, perfect (Reks and Gray, 1982, p. 167-168). The cleavage (Reks and Gray,
1979, p. 55). Mica can rotate into parallelism with the cleavage in the 1982) developed over a range of cleavage-normal shortening strains of
cleavage laminae as the supporting framework is removed by pressure 9%-26%, being invisible at lower strains and grading into solution cleavage
solution (Williams, 1972, p. 14-15, 39; Knipe and White, 1977, p. 363, at larger strains.
366). In many examples, the lack of bent micas at the border of cleavage
laminae indicates that micas within the cleavage laminae have recrystal- Extension-Fracture Cleavage
lized by pressure solution/diffusion to achieve their preferred orientation
(Oertel and Phakey, 1972, p. 6-7; Holey well and Tullis, 1975, p. 1299; Fracture cleavage is a parting defined by closely spaced discrete
White and Knipe, 1978, p. 168-169; Wintsch, 1978; Woodland, 1982, parallel fractures, ideally independent of any planar preferred orientation
p. 108; Lee and others, 1986, p. 773). Soft-sediment rotation of micas of grain boundaries that may exist in the rock (Turner and Weiss, 1963,
during compaction has been proposed as a mechanism to form slate by p. 98). Confusion over the use of this term goes back to itsfirstuse by Leith
Maxwell (1962) and Powell (1972) because of a near parallelism between (1905) who applied it to wide veins, joints, and pressure-solution cleavage.
soft-sediment sandstone dikes and cleavage, leading to the inference of In recent years, most field examples have been shown to be solution
contemporaneity. This concept has been largely discredited by the discov- cleavage. A few published examples appear to fit the Turner and Weiss
ery that pressure solution is a viable orienting mechanism and that the definition of fracture cleavage (Knill, 1960; Hancock, 1965; Price and
sandstone dikes, ostensibly parallel to cleavage, are in fact, not parallel but Hancock, 1972) and notably the example of Foster and Huddleston (1986,
owe their orientation to a later strain (Geiser, 1975, p. 718-719; p. 88-92). The fractures described by Foster and Huddleston are spaced
Groshong, 1976, p. 1136; Beutner and others, 1977; Beutner, 1980, 0.1-1.0 cm apart and are several centimeters long. They appear parallel in
p. 172-173). None of the true soft-sediment foliations described previously outcrop but are anastomosing in thin section and are often filled with
has resulted in the degree of lithification characteristic of slate. alteration products of the host rock. Refraction of the fractures is noticeable
Crenulation cleavage is defined by cleavage planes, whether mica- in layers of different composition. Good evidence from offsets shows the
ceous layers or sharp breaks, which are separated by thin slices of rock extensional nature of the fractures (Foster and Huddleston, 1986, p. 92).
containing a crenulated cross lamination (Rickard, 1961). The cleavage
laminae in many crenulation cleavages are explained by pressure solution FAULT-ZONE FABRICS
(Williams, 1972, p. 19; Gray, 1977a, p. 99; 1977b, p. 234; 1979,
p. 98-99). Crenulation cleavage may be found in fissile, clay-rich Differences of opinion currently exist about the proper use of such
sedimentary rocks (Geiser and Engelder, 1983, p. 126; Nickelsen, 1986, common terms as "fault," "fault zone," "cataclasite," and "mylonite" (Tul-
p. 365). My own observations lead me to believe that the cleavage laminae lis, and others, 1982; White, 1982; Wise and others, 1984,1985a, 1985b;
may be relatively independent of the crenulation geometry because the Mawer, 1985; Raymond, 1985). A fault zone is a tabular region across
laminae do not have a consistent relationship to the fold geometry. The which the displacement parallel to the zone is appreciably greater than the
cleavage may be found on limbs or axial surfaces of the crenulations and width of the zone (modified after Wise and others, 1984) and in which the
may change positions along the axial surface. deformation is greater than outside the zone. This means that even if
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

1342 R. H. GROSHONG, JR.

bedding or foliation is continuous across the zone, it may be described as a A hydroplastic fault zone is one in which the dominant deformation
fault zone, a reasonable, approach for the discussion of rock-deformation mechanisms are grain rotation and grain-boundary sliding of unlithified
fabric, even if not always appropriate for the description of map .patterns materials (after Petit and Laville, 1987). This type of fault zone is found in
(Mawer, 1985). A fault is a surface along which displacement has .¡taken unlithified sediments under low effective confining pressures, conditions
place. ' - which allow the grains to move past one another without breaking.
A fault surface is typically grooved (Dzulynski and Kotlarczyk, 1965,
p. 151) on scales ranging from meters (Nevin, 1949, p. 130; corrugations Conjugate Fault and Riedel Shear Geometry
of Hancock and Barka, 1987, p. 579) to hand sample (molded grooves of
Willis, 1923, p. 61) to hand lens (Means, 1987, p. 588) with the groove The geometry of conjugate faults, Riedel shears, and related features
axes being parallel to the displacement direction. Willis (1923) thought the has proved to be a very important unifying concept in understanding fault
grooves were the result of asperity plowing, but Means (1987) pointed zones from the microscopic scale to the map scale. First the conjugate fault
out that parallel grooves are parallel sided, fit perfectly into ridges on the geometry is described, then the Riedel geometry.
opposite side, and appear to be unrelated to any asperity. A smoothed and According to Handin (1969), the recognition that faults occur on
grooved fault surface is a planar feature named a "slickenside" by Cony- planes of high shear stress was due to Coulomb (1776) who determined
beare and Phillips (1822) according to Dennis (1967). The grooves or that faults do not occur on the planes of maximum shear stress (45° to aj)
striations on the slickenside are termed "slickenlines" by Fleuty (1975) but rather occur at an angle on the order of 30° to o\. A fault initiated at
and give the line of slip of the fault. Slip lineations on fault surfaces are also this angle is often called a "Coulomb fault." Daubree's (1879) uniaxial
produced by crystal growth fibers. Although the original definition of the compressive tests on wax may have been the first analog experiments to
term "slickenside" is for polished or smoothed surfaces, not crystal fibers show the conjugate geometry in a geological context. In his classic 1905
(Fleuty, 1975), current usage includes both, and this practice will be paper, Anderson used the concept of a pair of conjugate Coulomb faults at
followed here. 30° to o\ that intersect parallel to 02 to explain normal, thrust, and
Fault-zone materials include mylonite, cataclasite, gouge, and brec- strike-slip fault orientations. The conjugate (Andersonian) fault geometry
cia. Mylonite was originally defined by Lapworth (1885, p. 559) for rocks is seen, although not regularly, in pure shear experiments on rocks
in fault zones that were "crushed, dragged and ground out into a finely (Adams, 1910; Paterson, 1958, PI. 1; Griggs and Handin, 1960, Pis. 7-8;
laminated schist composed of shattered fragments of the original crystals of Hadizadeh and Rutter, 1983, p. 502) and sand (Hubbert, 1951, PI. 1;
the rock . . . " ; however, the type example (Higgins, 1971, p. 18) shows Horsfield, 1980, p. 307); is common in clay (H. Cloos, 1930; E. Cloos,
clear evidence of crystal-plastic deformation textures, resulting in a persist- 1955, PI. 1) and is observed on the small scale in the field (Cloos, 1947b,
ent ambiguity in the use of the term. The definition also included a state- p. 900; Stearns, 1972, p. 163; Lockwood and Moore, 1979, p. 6045;
ment that the rock must be foliated, leaving unfoliated fault-zone rocks Ramsay, 1980a, p. 85).
without a name. Bell and Etheridge (1973) pointed out that "brittle de- Perhaps the first published simple-shear fault-zone experiment was
formation is unnecessary for the formation of a typical mylonite" (p. 337) the clay-model study of Cloos (1928). Riedel (1929, p. 361-362) repro-
and defined the term as "a foliated rock, commonly lineated and contain- duced the experiment, obtaining fault zones defined by en echelon tensile
ing megacrysts, which occurs in narrow, planar zones of intense deforma- cracks and en echelon faults and correctly interpreting oi as being 45° to
tion. It is often finer grained than the surrounding rocks into which it the trend of the fault zone. Although Riedel obtained only the conjugate
grades" (p. 347). According to Waters and Campbell (1935), the term direction having the same sense of displacement as the zone as a whole,
"cataclasite" was introduced by Grubenmann and Niggli (1924) for an similar experiments commonly result in en echelon sets of both conjugate
aphanitic, structureless rock that differs from mylonite (in the sense of directions (Cloos, 1955, PI. 3). Hills (1963) stated that the faults in this
Lapworth) by an absence of foliation. Recent classifications (Sibson, 1977; experiment are known as "Riedel shears." Experimentally, this geometry is
Wise and others, 1984; Tullis and others, 1982) use cataclasite for a characteristic of a fault zone that is forced by the boundary conditions to
nonfoliated rock formed by brittle mechanisms, and "mylonite" as a fo- have an orientation parallel to the displacement direction, as in the Cloos-
liated or generally foliated rock formed by crystal-plastic mechanisms. Riedel experiment or in a closed shear box (Morganstern and Tchalenko,
Both soft-sediment fault zones (Mandl and others, 1977) and brittle fault 1967a).
zones (Proctor and others, 1970, Fig. 4; Gay and Ortlepp, 1979, p. 53; The terminology of Riedel shear zones (Fig. 2) is due mainly to
Chester and others, 1985) may be foliated, however. Cataclasite is here Skempton (1966, p. 330). The fault zone trend is the D direction. Faults
defined as a cohesive rock formed mainly by brittle fracturing and usually parallel to this direction form after the Riedel (R and R') shears. Dis-
showing evidence of grain rotations and grain-size reduction. A cataclasite placement may be concentrated on one or two D shears that are then
might be formed by local crushing without significant fault oftset (cataclas- called "principal displacement shears." The D shear is equivalent to the C
tic flow in the sense of Stearns, 1969) and so is not necessarily restricted to direction in the terminology for crystal-plastic fault zones of Berthe and
fault zones. Mylonite is a rock formed mainly by crystal-plastic deforma- others (1979) and the Y direction of Bartlett and others (1981). The Riedel
tion mechanisms and shows evidence of internal rotation and grain-size shears are conjugate to the <n axis (45° to D): the R shear is at ~ 15° to the
reduction. Cataclasite or mylonite may be foliated or unfoliated, although trend of D and has the same sense of displacement, the R' shear is its
on the thin-section scale, many cataclasites are unfoliated and most mylo- conjugate at -75° to the trend of D and having the opposite sense of
nites are foliated. displacement. A tensile fracture 45° to D and parallel to the o\ axis is
The noncohesive equivalents of cataclasite are gouge and fault brec- termed "T." "Thrust-shears," labeled P, have the same sense of slip as R
cia (Higgins, 1971). A fault breccia is composed of angular or rounded shears and form at 10°-30° to D. In the normal fault zones illustrated by
fragments formed in the fault zone and consists of more than 30% frag- Skempton (1966), the P shears had a thrust orientation, hence the name,
ments large enough to be seen by the naked eye. Gouge is a paste-like rock but the designation P is from his interpretation that the material was in the
material formed in the fault zone in which less than 30% of the fragments passive Rankine state (incipient failure with horizontal stress greater than
are large enough to be seen by the naked eye. Cataclasite series (Sibson, vertical stress: Jaeger and Cook, 1979, p. 415). In their rock-model exper-
1977) represents all cohesive and noncohesive rocks fitting the definitions iments, Bartlett and others (1981) observed the aforementioned features
of cataclasite, gouge, and breccia. and a new direction, termed "X," that is perpendicular to R and has the
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

LOW-TEMPERATURE DEFORMATION MECHANISMS 1343

same sense of displacement as R'. A Riedel shear zone thus contains more Shear lenses (Skempton, 1966, p. 331) are small regions bounded by
than just R and R' shears. A Riedel shear zone has the same over-all slip surfaces. Rhombic zones are bordered by R and D shears; trapezoidal
geometry as the first- and second-order shears of Moody and Hill (1956) zones, by R, D, and P shears. Slickenlines on shear lenses may be oblique
but the opposite sequence of formation; the smaller faults form first. to the direction of displacement of the fault zone as determined from the
Foliations are produced by inequant minerals within the fault zone slickenlines on the principal displacement shears. Shear lenses might also
(notably clays) that rotate to positions nearly parallel to the individual be caused by the intersection of R-foliations with a bedding foliation (Piatt
faults (Weymouth and Williamson, 1953; Morganstern and Tchalenko, and Vissers, 1980, p. 402-403) or by intersecting R and R' foliations (Piatt
1967b, p. 151; Tchalenko, 1968, p. 168; Piatt and Vissers, 1980; Maltman, and Vissers, 1980, p. 404). Substantial extension by this mechanism may
1987), or rotate to approximate the finite strain orientation (Morganstern result in boudinage of relatively stiff layers (Moore and Allwardt, 1980,
and Tchalenko, 1967a, p. 316), or kink (Morganstern and Tchalenko, p. 4746).
1967a, Figs. 6-17; Tchalenko, 1968, p. 169). Kinking may be related to The Anderson (1905) and Riedel (1929) models of faulting are two
the anisotropy of the mineral foliation (Tchalenko, 1968, p. 163) or may dimensional, applying to plane strain. As a result of very careful analysis of
represent kinking of a set of parallel faults (Morganstern and Tchalenko, the results of experimental clay deformation, Oertel (1962, p. 29-30;
1967a, p. 319) with passive rotation of the mineral fabric. Solution cleav- 1965, p. 355) recognized that four conjugate fault directions had formed
age may also produce a foliation within the fault zone (Alvarez and others, simultaneously. The faults formed with a geometry that can be visualized
1978; Ghisetti, 1987, p. 691-692). A mineral shape foliation oblique to as two conjugate pairs for which the dihedral angle of each pair is bisected
the trend of the fault zone has been termed the "S plane" in crystal-plasticby a i and the line of intersection of each pair makes a small but non-zero
fault zones (Berthe and others, 1979). Rutter and others (1986) believe the angle withCT2such that the resulting pattern is symmetric across the a \ - o i
foliation in a Riedel shear zone to be parallel to the P shears and call it a "P
plane. Three or more conjugate faults formed simultaneously can be
foliation." Fault zones might show foliations parallel to cleavage; rotated termed an "Oertel conjugate geometry" (suggested by Z. Reches, 1977,
bedding; and R, R', P, and D shears, all being approximately contempo- personal commun.) to distinguish it from the Andersonian pair of conju-
raneous, and none of which necessarily exists in the rock adjacent to the gates. Aydin (1977) and Reches (1978, p. 113-114) recognized the Oertel
fault zone. Fault zones are seen to grow as Riedel shear zones from the geometry in the field and interpreted it to be the result of a three-
microscopic to the map scale (Morganstern and Tchalenko, 1967a; Tcha- dimensional strain accomplished by faulting. Field examples of the con-
lenko, 1970), and so multiple Riedel foliations are possible within a singletemporaneous formation of three or four sets of conjugate faults (as
fault zone. opposed to slip on pre-existing faults) have been interpreted using this
The origin of the P shear, which is not directly related to the stress concept (Bruhn and Pavlis, 1981, p. 288-289; Aydin and Reches, 1982,
field causing R, R \ and T, has been explained by Gamond (1983) as the p. 109-110; Underhill and Woodcock, 1987) and have been produced in
result of stress reorientation between en echelon R shears. Using the theory true triaxial rock-deformation experiments (Reches and Dieterich, 1983,
of Segall and Pollard (1980), Gamond showed that the P shear has the p. 114-116). The interpretation of Reches (1978, 1983) is based on the
expected 30° angle of a Coulomb fault to the reoriented maximum com- analysis of the number of slip systems required to accommodate a general
pressive stress axis between en echelon R shears. Gamond (1983, p. 37) three-dimensional strain, analogous to the theory for the number of slip
also showed that sliding on the P shears can cause the R shears to open. systems required in crystal-plastic deformation by Taylor (1938) and
This leads to the important observations that all open or filled cracks in Bishop (1953). Oertel conjugate faults may explain the shear lens
fault zones are not necessarily parallel to T (Gamond, 1983, p. 34-35), geometry and slip directions noted by Skempton (1966).
and that dilation across the fault zone may favor the formation and open-
ing of P shears rather than the formation of T shears. The X-direction Hydroplastic Fabric
fractures in the experiments of Bartlett and others (1981, p. 266-270)
appear to form as connections between R shears and have the orientation In unlithified materials having a major component of platy minerals,
and sense of displacement appropriate for a slightly rotated conjugate to P a Riedel shear zone forms the typical fault-zone fabric (compare Petit and
that would form at the ends of the R shears (compare Gamond, 1983, Laville, 1987, p. 111). The platy minerals are aligned parallel or nearly
Fig. 15d). parallel to the R, R', P, D, or X shear directions (Morganstern and Tcha-
R shears tend to remain active or form anew during the displacement lenko, 1967b, p. 148; Maltman, 1977, p. 424; Carson and Bergland, 1986,
of a fault zone and may cut D shears (Fig. 2; Skempton, 1966, p. 330). p. 144-145). Maltman (1987, p. 80) experimentally produced R shears
Slip on the R shears causes extension of the D shears. Late extensional that curved into the D direction. Slickensided, bi-pyramidal cone-shaped
crenulation cleavage (Piatt and Vissers, 1980, p. 402-403) and small faults, perhaps analogous to fault cones sometimes produced in the triaxial
brittle faults (Piatt and Leggett, 1986, p. 192, 200) may form in the
R-shear orientation.

¡V
I I
Figure 2. Nomenclature of a Rie-
del shear zone, after Skempton (1966)
D=Displacement shear
and Bartlett and others (1981).
V R = Riedel shear
P = Thrust shear
T=Tension shear
X from Bartlett and others, 1981
Downloaded from gsabulletin.gsapubs.org on February 27, 2013
1344 R. H. GROSHONG, JR.

compression of cylindrical samples have been found to form during shale a low angle to, the fault-zone boundaries is reasonably common, marked
compaction (Guiraud and Séguret, 1987, p. 127). Kinks are also common by compositional layering (Proctor, 1970, Fig. 4; Bonilla and others, 1978,
and may be transitional into fault zones (Tchalenko, 1968, p. 163, 169; p. 354; Mitra, 1984, p. 58; Chester and others, 1985, p. 141), elongate
Lundberg and Moore, 1986, p. 22-23). Dismembered folds contribute grain alignment (House and Gray, 1982, p. 259; Segali and Pollard,
significantly to the texture of accretionary prism fault-zone fabrics (Nelson, 1983b, p. 559; Chester and others, 1985, p. 141), and by variations in
1982, p. 627-628, 632; Lucas and Moore, 1986, p. 95); some are sheath- grain size (Wojtal and Mitra, 1986, p. 682, 684). Foliations of all three
like folds with axes parallel to the transport direction (Hibbard and Karig, types occur in the fault zone described by Gay and Ortlepp (1979, p. 54).
1987, p. 852). Also characteristic of accretionary prisms is scaley foliation The most distinctive textural features are grain fracture, fragment
defined by anastomosing polished and slickensided fracture surfaces, per- separation, and rotation. Grain sizes within the matrix are reduced from
vasive on the scale of millimeters (Lundberg and Moore, 1986, p. 18-19; that of the host to sizes of 5-25 jum as seen in thin section (Engelder,
Moore and others, 1986). This fabric seems to be the same as the shear 1974a; Aydin, 1978), but fragments down to the size range of 0.06-0.1
lenses of Skempton (1966, p. 321) where extension on R shears may be fxm are seen in SEM. The smaller particles tend to be very angular, and the
very prominent (Nelson, 1982, p. 628; Cowan, 1985, p. 453) but may quartz fragments may be basal cleavage flakes, produced by high strain
represent multiple shear foliations (Maltman, 1988, p. 173). Raymond rate (Gay and Ortlepp, 1979, p. 55; Moody and Hundley-Goff, 1980,
(1984) defined mélange as a body of rock mappable at a scale of 1:24,000 p. 306; Olgaard and Brace, 1983, p. 14). Larger fragments, on the order of
or smaller and characterized both by the lack of internal continuity of 30 /um or more, may be rounded (Engelder, 1974a, p. 1515; Moody and
contacts or strata and by the inclusion of fragments and blocks of all sizes, Hundley-Goff, 1980, p. 306; House and Gray, 1982, p. 359; Mitra, 1984,
both exotic and native, embedded in a fragmented matrix of finer-grained p. 58; Wojtal and Mitra, 1986, p. 682) or angular (Anderson and others,
material. The mélange fabric appears to be typically that of a large-scale, 1980, p. 229; House and Gray, 1982, p. 259-261; Blenkinsop and Rutter,
soft-sediment (or cataclastic) Riedel shear zone having large displace- 1986, p. 671). Bonilla and others (1978) reported rounded and polished
ments. The same textures may be seen in thin sections of shale-matrix fault pebbles and rock fragments in the San Andreas fault zone. Rounding is
zones. due to abrasion during rotation in the fault matrix (Pittman, 1981,
The Riedel R and D shear-zone geometry is also seen in hydroplastic p. 2382). Moody and Hundley-Goff (1980, p. 306) found a greater degree
fault zones in equigranular materials such as sand (Petit and Laville, 1987, of rounding in wet experiments than in dry experiments. Some rounded
p. 111). A weak grain-size segregation foliation occurs parallel to the slip grains may be relatively unaltered host-rock grains in sandstones (Dunn
direction of individual faults, and elongate particles are rotated to near and others, 1973, p. 2411; Pittman, 1981, p. 2382). Cathodoluminescence
parallelism with the slip direction (Thomson, 1973, p. 528; Mandi and of a natural fault zone showed some of the matrix to be undispersed
others, 1977, p. 101, 111, 113-118; Maltman, 1988, p. 172). Very minor fragments of the adjacent larger grains (Lucas and Moore, 1986, p. 1012).
cataclasis produces rounding of the more angular grains and leads to an Cataclasite clasts are found within cataclasites (Brock and Engelder, 1977,
increase in fines within the fault zone (Mandi and others, 1977). The p. 1669; Anderson and others, 1980; House and Gray, 1982, p. 262; Mitra,
boundaries between the country rock and the fault zone are not sharp on 1984, p. 58; Wojtal and Mitra, 1986, p. 683). Gouge injections are re-
the thin-section scale, and original bedding may be obliterated in the fault ported by Engelder (1974a, p. 1516) and Brock and Engelder (1977,
zone. Fold and kink-band geometries may occur that are transitional into p. 1669). Cataclasite veins (Engelder, 1974a; Gretener, 1977; Labaume,
fault zones (Hubbert, 1951, PI. 2; Sieh, 1984, p. 7652; Van Loon and 1987) show that the material is easily deformable.
others, 1984, p. 358). Dilation is required within the fault zone to allow The grain-scale microfractures are virtually always extensional
the grains to move past one another without fracturing (Mandi and others, (Aydin, 1978, p. 920) but may slide after formation, although rotation
1977). The depth to which this can occur is controlled by the fluid pres- appears to be more common than sliding. The well-ordered deck-of-cards
sure; high fluid pressures will permit soft-sediment deformation to consid- sliding and rotation characteristic of crystal-plastic shear zones (Simpson
erable depth. and Schmid, 1983, p. 1286; Tullis and Yund, 1987, p. 607) appears to be
The porosity and permeability of soft-sediment fault zones generally extremely rare or absent in cataclastic fault zones. Vein fillings in micro-
reflect that of the material entrained in the zone. Based on the analysis of fractures have been observed by Aydin (1978, p. 920, 926) and Blenkin-
normal faults in the Gulf of Mexico that are sealing or nonsealing to sop and Rutter (1986, p. 671). Larger-scale transgranular veins occur in
hydrocarbon migration (Smith, 1980), the fault zone is a barrier to hydro- the cataclasite (Stel, 1981, p. 587,590; House and Gray, 1982, p. 262) and
carbon migration if it contains mainly shale or is a conduit if it contains may contain fibers (Mitra, 1987, p. 574), dogtooth crystals, or voids (Stel,
mainly sand (Weber and others, 1978, Fig. 11). Where a single sand body 1981, p. 590-596; Chester and Logan, 1986, p. 87). Stylolites may occur
is in contact across the fault, the fault zone is not a barrier. Weber and in the matrix as either the grain-to-grain type (Brock and Engelder, 1977,
others (1978, Fig. 6) found a zone of dilation on the hanging wall of a p. 1669) or the wavy transgranular variety (Moore and Allwardt, 1980,
small normal fault and a parallel zone of compaction on the footwall, p. 4747; House and Gray, 1982, p. 259-262; Mitra, 1984, p. 57-58;
implying that fluids might have been better able to migrate along the Chester and Logan, 1986, p. 87; Wojtal and Mitra, 1986, p. 682).
hanging-wall side of the fault zone. Mimran (1985, p. 373) described a Pressure shadows around large grains are rare and where present are
normal fault in chalk that was a conduit for fluid flow as indicated by the filled with fragments of the adjacent large grain (House and Gray, 1982,
presence of calcite veins having the isotopie composition of meteoric p. 361; Lucas and Moore, 1986, p. 96; Rutter and others, 1986, p. 6). In
water. Because of the different abilities of oil and water to migrate through contrast, pressure shadows in mylonitic fault zones are relatively common
water-wet rocks, a fault might be a barrier to oil migration, but not to and are elongate and filled with recrystallized grains (Berthé and others,
water migration (Smith, 1980). 1979, p. 34; Watts and Williams, 1980, p. 327; Simpson and Schmid,
1983, p. 1282-1283).
Cataclastic Fabric First described by Heald (1956, PI. 2), very thin fault zones having
widths of a few millimeters or less but lengths of meters or more contain
Cataclastic textures are similar whether or not the material has pri- gouge or cataclasite like that previously described. They are usually but
mary cohesion. Jaeger (1959) produced cohesionless powder (gouge) in not always lighter in color than the host rock and show offsets in the range
dry friction experiments and a compacted dense material (cataclasite) in of less than a grain diameter to centimeters (Aydin, 1978, p. 917; Aydin
otherwise identical water-wet experiments, suggesting that cohesion is con- and Johnson, 1978, p. 934; Pittman, 1981, p. 2382; Jamison and Stearns,
trolled by the presence of water in the fault zone. Foliation parallel to, or at 1982; Byrne, 1984, p. 36). Shiny, slickensided surfaces in rock that appear
Downloaded from gsabulletin.gsapubs.org on February 27, 2013
LOW-TEMPERATURE DEFORMATION MECHANISMS 1345

to have no associated gouge may be called "fault mirrors" (for example, p. 309; Spray, 1987). The maximum temperatures are about 1150 °C over
Petit and others, 1983) and are probably fault zones of this type that have the background of 100-125 °C (Teufel and Logan, 1978). Moody and
just enough gouge adhering to the surface to smooth it (Nelson, 1985, Hundley-Goff (1980) produced glass only in their dry experiments, not in
p. 43-47). water-wet experiments. Pseudotachylite is a glassy appearing rock, dark
The Riedel shear geometry also occurs within cataclastic fault zones. colored and of veinlike or pseudointrusive character which bears a strong
Jackson and Dunn (1974, p. 243), Engelder and others (1975), and Byer- resemblance to tachylite (after Shand, 1916; and Waters and Campbell,
lee and others (1978, p. 168) produced R shears within experimental 1935, but removing reference to fusion). It usually contains angular clasts
gouge zones. R and R' shears (Friedman and Higgs, 1981, p. 14; Logan of the country rock (Shand, 1916, PL 19). Pseudotachylite is generally
and others, 1981, p. 131) as well as D shears (Logan and others, 1981) glass or devitrified glass (Shand, 1916; Sibson, 1980, p. 168-169; Mad-
have been produced. Chester and Logan (1986) interpreted the faults in dock, 1983, p. 106), but some examples may be ultrafine-grained catacla-
the damage zone adjacent to the Punchbowl fault zone as R and R' shears. site (Waters and Campbell, 1935; Wenk, 1978, p. 508-509) or both
In a very important contribution, Gay and Ortlepp (1979) found en (Philpotts, 1964, p. 1015, 1017,1021).
echelon conjugate shears that were clearly produced by a seismic event Porosity is decreased by cataclasis in high-porosity rocks but in-
(p. 53) in a mining-induced fault zone, with the R shears being larger and creased in low-porosity rocks. Aydin (1978) found that a host-rock poros-
the R' shears forming mainly in the areas of R shear overlap. The resulting ity averaging 24%-25% was reduced to 6%-10% in natural fault zones.
shear zone appears to be made up of incompletely connected R, R', and D Pittman (1981) found the host-rock porosity of 30% and permeability of
shears. 51,394 md reduced to 3% and 0.05 md in cores consisting mainly of small
Increasing fragmentation toward a large-displacement fault zone has natural fault zones. In experiments on clay-rich and clay-absent gouges,
been documented by an increase in the number of small faults and an Morrow and others (1984) found a wide range of small permeabilities
increase in the number of fractured grains (Engelder, 1974a; Brock and from 10"22 to 10~18 m 2 (where 1 darcy = 0.987 * 1 0 1 2 m2). Nelson
Engelder, 1977, p. 1669; House and Gray, 1982, p. 258; Chester and (1985, p. 432) found that polished-gouge slickensides may show signifi-
Logan, 1986, p. 84, 86; Wojtal and Mitra, 1986). The principal cant permeability due to small mismatches in the surface. He reported
displacement fault is not necessarily in the center of the damage zone permeabilities from a core containing a slickensided surface as 18-211 md
(Wallace and Morris, 1986, p. 120). oblique to the slickenside and 1,532 md parallel to the slickenside. Under-
Fault breccias appear to be of two types: vein derived and fault ground mining experience reported by Wallace and Morris (1986,
derived. In vein-derived breccias, the host-rock fragments are bounded by p. 120-123) is that most fault zones are wet, some produce water, some
veins (Carson and others, 1982, p. 286; Chester and Logan, 1986, p. 90; act as barriers to fluid migration, and that the greatest water production is
Gaviglio, 1986, p. 249; Mitra, 1987, p. 583, 585; Ramsay and Huber, from the moderately fractured damage zone adjacent to the largest fault
1987). The veinfillingmay be derived locally by pressure solution, may be zones. Dilatancy prior to (Brace and others, 1966) and during (Teufel,
introduced during deformation by high-pressure fluids (Rye and Bradbury, 1981) faulting of low-porosity rocks may increase the porosity by up to 8%
1988), or may be the result of later hydraulic fracturing or opening of the (Teufel, 1981, p. 143).
damage zones adjacent to the fault zone. Chester and Logan (1986) found
that up to 5%-10% of the clasts in the Punchbowl fault zone are vein Pressure-Solution-Dominated Fabric
fragments, indicating many fracturing andfillingevents. Veins appear to be
common in crystal plastic fault zones as well (White and White, 1983, A pressure-solution-dominated fault zone, as defined here, must con-
p. 583; Sibson, 1986a, 1987, PI. 2; Stel, 1986, p. 295) and usually show tain abundant evidence of removal of material along stylolites, slickolites,
evidence of multiple events of vein formation and deformation. Fault- or solution cleavage as seen in examples by Alvarez and others (1978),
derived breccia blocks are bounded by thin shear zones (Heald, 1956, PL Rispoli (1981, p. T31), and Marshak and others (1982, p. 1018). Fault
2; Brock and Engelder, 1977, p. 1669; Carson and others, 1982, p. 283; zones dominated by pressure solution may also contain crystal growth-
House and Gray, 1982, p. 260; Dokka, 1986, p. 81; Lucas and Moore, fibers along slickensides (Elliott, 1976, PL 12), fiber-filled veins (Spang
1986, p. 99-100) or injected cataclasite (Labaume, 1987, p. 154). Farmin and Groshong, 1981, p. 330), mosaic-filled veins (Marshak and others,
(1941, p. 148) stated that in the western United States breccia is more 1982, p. 1018), Riedel shears (Casas and Sabat, 1987, p. 649), and cata-
common on normal faults than on reverse faults. Phillips (1972) reported clastic material (Ghisetti, 1987, p. 962). Slickenside fibers and vein fillings
significant vein-fill breccia on normal faults, whereas Gretener (1977), in might be sourced from outside the fault zone and so are not sufficient
contrast, cited a number of large-displacement thrust faults that are only evidence to define deformation by pressure solution.
centimeters thick. The origin and occurrence of fault breccia merits further The fiber orientations and country-rock screens characteristic of
research. crystal-fiber slickensides were first explained by Durney and Ramsay
The thermal alteration of organic components of the rock provides (1973, p. 87). The initial slip surface contains offsets that open as slip
data on the thermal effects in fault zones. Wilson (1971) reported that fault occurs (the offsets may have initiated as an en echelon array of T fractures,
zones with displacements of a few feet in Oklahoma oil basins had no Ramsay and Huber, 1983, p. 258). As the displacement on the slip sur-
effect on the color of palynomorphs, whereas fault zones with displace- faces becomes large, thefiberbundles growing from nearby offsets overlap
ments of more than 100 ft caused palynomorphs to change from brown to to bring fibers of different ages into contact. Screens of the wall rock may
black within 600-800 ft of the fault plane (equivalent to a temperature be preserved between the fiber bundles producing the typical laminated
change from about 100 to 175 °C; Bostick, 1974). For fault zones in the appearance (Ramsay and Huber, 1983, p. 258-261). If the slip direction
frontal ranges of the Canadian Rocky Mountains, Bustin (1983) reported changes through time, because thefiberbundles are of different ages where
vitrinite reflectances of 0.72%-3.09% in the fault zone that drop to regional they are superimposed, they will have differentfiberdirections resulting in
values of 0.78%-1.21% a meter or less away from the fault zone. For superimposed slickensides with different slickenline orientations (Durney
heating times as short as 1 hour, this could correspond to temperatures of and Ramsay, 1973, p. 89).
430-650 °C locally compared to the regional background of 250-350 °C One type of fault zone develops from closely spaced solution cleav-
(Bustin, 1983). age. Alvarez and others (1978, p. 265) documented an increasing intensity
Glass has been produced in low-temperature friction and faulting of solution cleavage to spacings representing greater than 35% material loss
experiments (Friedman and others, 1974, p. 938; Jackson and Dunn, as a fault zone is approached, concomitant with a curving of cleavage to
1974, p. 248; Teufel and Logan, 1978; Moody and Hundley-Goff, 1980, a low angle to the (thrust) fault zone. In the fault zone, there is a chaotic
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

1346 R. H. GROSHONG, JR.

mixture of fragments that are derived from detached cleavage microlith- stone, no recrystallization, and vitrinite reflectance of 2.5%-3.5%
ons. Fault zones may consist of closely spaced solution cleavage (Bogacz, (Groshong and others, 1984a) equal to perhaps 160-190 °C (well-
1981) which may be cut by Riedel shears (Koopman, 1983; Lavecchia, standard curve, Bostick, 1974, p. 9). Thus superplastic flow may have
1985). The early development of such a fault zone may be represented by occurred in the fault zone while the remainder of the rock was only
the cleavage duplex recognized by Nickelsen (1986, p. 362,367,369). The moderately deformed without recrystallization.
cleavage within the duplex curves (Nickelsen, 1986, p. 364) to approxi-
mate the expected trajectories of the finite elongation direction (Ramsay Inference of Seismic Activity from Fabric
and Graham, 1970) but does not actually join the fault surface. One of the
associated cleavage-duplex thrusts is a zone of sheared shale containing It has been accepted since about 1900 that most earthquakes occur
euhedral crystals and cataclastic fragments that were originally precipi- along pre-existing faults (Howell, 1986) and are caused by the release of
tated in veins or fault-step pressure shadows and were broken by continu- stored elastic energy (Reid, 1910). It has long been known that the exper-
ing fault movements (Nickelsen, 1986, p. 367). imental formation of a fault in intact rock results in the release of elastic
A closed-system fault zone consisting of faults, stylolites, and veins energy, but not until the work of Jaeger (1959) was it shown that sliding
(Rispoli, 1981, p. T31) evidently began with the formation of parallel on an existing fault (or sawcut) in rock would also result in elastic energy
small faults as the slip surfaces in a slightly rotated kink band. The dis- release. This behavior is termed "stick-slip" because the energy release
placements on the small faults terminate in veins and stylolites that make occurs during periodic slip episodes characterized by stress drops that are
low angles to the kink-band boundary. Additional rotation results in con- only a fraction of the total differential stress. Brace and Byerlee (1966) and
nection along the kink band by the formation of additional small faults, Byerlee and Brace (1969) first proposed that stick-slip sliding could be the
veins, and stylolites to form a through-going fault zone. An orthogonal cause of earthquakes. Factors that influence the occurrence of stick-slip
array of veins and stylolites may have a flow law similar to grain-boundary behavior include temperature, pressure, and strain rate (Byerlee and Brace,
diffusion (Fletcher and Pollard, 1981). A block size of 1-10 cm and a 1969), thickness of gouge zone (Summers and Byerlee, 1977), and compo-
differential stress of 30 bars gives a reasonable geological strain rate. sition of the gouge (Summers and Byerlee, 1977; Logan, 1979). Nearly all
common crustal minerals show stick-slip behavior under suitable condi-
Superplastic Fabric tions, including many phyllosilicates but with the exception of vermiculite
and montmorillonite (Summers and Byerlee, 1977). In mixtures the fric-
Superplasticity is defined for metals as a quasi-viscous behavior tional behavior is controlled by the softer mineral (Logan, 1979; Logan
whereby elongation can exceed 1,000% without necking and subsequent and Rauenzahn, 1987).
failure (Boullier and Gueguen, 1975, p. 93); no specific deformation The formation of gouge and slickensides during stick-slip experiments
mechanism is implied (Schmid and others, 1977, p. 258). The interpreted (Jaeger, 1959) suggests that fault-zone textures might correlate with seis-
deformation mechanism is diffusion- or dislocation-assisted grain- mic behavior. Short carrot-shaped grooves on slickenside surfaces were
boundary sliding (Schmid and others, 1977, p. 277). The texture in rocks shown experimentally by Engelder (1974b, p. 4388) to correlate with
is characterized by square to loaf-shaped grains (Boullier and Gueguen, stick-slip events because their lengths are comparable to the slip in a single
1975; Schmid and others, 1977, p. 271; Behrmann, 1985, p. 104) in which event and they fail to occur during stable sliding. The features are also seen
triple-point grain intersections do not usually make 120° angles. The grains on natural slickensides (Engelder, 1974b, p. 4390-4391). Application of
are relatively undeformed internally and have no more than a weak crys- this criterion to felt earthquakes is problematical because the grooves are
tallographic fabric (Schmid and others, 1977; Behrmann, 1985). Grain only 0.4-2 mm long, much shorter than the slip in most earthquakes.
sizes on the order of 1-10 /urn are required for the processes to be impor- McKenzie and Brune (1972) proposed that pseudotachylite forms by melt-
tant at the lower experimental temperature ranges (Schmid and others, ing during seismic events. Pseudotachylite is rather rare (Sibson, 1986b;
1977), and this is the grain size found in field examples (Boullier and Spray, 1987), however, and is most commonly reported in association
Gueguen, 1975; Schmid and others, 1981, p. 156; Behrmann, 1985, with mylonite (Sibson, 1980; Hobbs and others, 1986) or in impact craters
p. 104). Superplastic deformation in limestone and quartzite is normally (Reimold and others, 1987). No texture in cataclastic fault zones has yet
associated with crystal-plastic strain at temperatures above 400 °C for been shown to have a unique correlation to felt earthquakes.
which the small grain size is achieved by syntectonic recrystallization in a Riedel shears form within the gouge in stick-slip experiments (Fried-
mylonite (Schmid and others, 1977; Etheridge and Wilkie, 1979), but man and Higgs, 1981; Logan and others, 1981; Hiraga and Shimamoto,
temperatures may be as low as 250-400 °C (Behrmann, 1985). 1987). Riedel shears also formed in intact rock during an earthquake in
Small (1-6 /urn), nearly equant, interlocking grains are relatively what might be the only known earthquake source region ever mapped
common as the matrix in low-temperature fault zones that otherwise ap- (Gay and Ortlepp, 1979). Significantly, the double-couple nature of earth-
pear to be cataclastic (Phillips, 1982, p. 112; personal observations). This quake source mechanisms implies slip along more than one plane during
texture could be the result of either the recrystallization of gouge (Jaeger, an earthquake (Howell, 1986). Perhaps the presence of Riedel shears
1959) or superplastic flow; criteria for making the distinction have yet to within the fault zone will prove to be an index of seismic activity.
be developed. The presence of reactive fluids should enhance either proc-
ess (Etheridge and Wilkie, 1979). The interlocking 4- to 6-/um grain-size Sense-of-Shear Criteria
texture in a quartz-matrix fault zone formed at 250-350 °C is interpreted
by Phillips (1982) as being the result of superplastic flow. The Lochseiten Sense-of-shear criteria are divided here on the basis of scale into
mylonite on the Glarus thrust of eastern Switzerland may be another fault-zone features and fault-surface features. Fault-zone features of value
example. The Lochseiten is a 1- to 2-m-thick foliated limestone consisting include "drag" folds, gash or feather fractures, cleavage, and foliation. A
of equant, 1- to 3-/um diameter grains (Schmid, 1975, p. 262). The fault "drag" fold (Nevin, 1949, p. 136) is a fold adjacent to a fault and is convex
zone is characterized by either mylonite or by extreme cataclasis and toward the direction of motion of the block in which it occurs. The drag
brecciation in a zone a few tens of meters thick (Schmid, 1975, p. 261). probably occurs before faulting, not as a result of slip along the fault
Foliation-parallel stylolites and veins indicate pressure solution in the my- (Nevin, 1949, p. 142; Dennis, 1987). Extension fractures at 45° to the fault
lonites (Schmid, 1975, p. 266). In the same area, the hanging wall of the zone (T direction, Fig. 2) werefirstnamed "feather joints" by Cloos (1922;
Glarus thrust is characterized by 6%-9% crystal-plastic strain in the lime- also Cloos, 1932), according to Dennis (1967). The same features were
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

LOW-TEMPERATURE DEFORMATION MECHANISMS 1347

termed "gash fractures" by Gilbert (1928) and "pinnate tension joints" by A number of displacement criteria can be related to Riedel shears. An
Hills (1953). The sense of slip on the fault is as given by Figure 2. Solution R shear offsetting a D shear causes a low-angle congruous step for which
cleavage, stylolites, and mineral-shape foliations within a fault zone are the sense of offset on the R shear is the same as for the D shear (Angelier
usually approximately parallel to the extension strain direction; the acute and others, 1985; Piatt and Leggett, 1986, p. 192). The typical incongru-
angle between foliation and the fault trend points in the direction of slip of ous fracture step appears to be caused by a fracture (X?) between an R
the adjacent block. shear and the D surface (Currie, 1969, p. 169-171; Angelier and others,
Thefirstcomprehensive review of fault-surface sense-of-shear criteria 1985, p. 352). This is perhaps the most likely explanation of the steps in
seems to be in the textbook of Willis (1923, p. 54-62). Willis pointed out the experiment of Paterson (1958, PL 2). Crescentic gouge marks and
that asymmetrical steps commonly occur on slickensided fault surfaces lunate fractures (Wegmann and Schaer, 1957, p. 492) are similar in cross
such that a finger run over the surface in the smooth direction is moving in section, but the former is convex in plan view in the direction of displace-
the direction of displacement of that side of the fault. This became the ment of the opposing block, and the latter is just the opposite. Pinnate
standard sense-of-displacement criterion in later textbooks. The interpreta- shears (Hills, 1963, p. 174-176) are slickensided incongruous steps mak-
tion was challenged by Paterson (1958, PL 2), who illustrated experimen- ing a high angle to the fault surface and having the orientation of R' shears.
tally produced fault steps with the opposite asymmetry, and soon afterward This is the geometry of the incongruous slickensided steps seen in the field
Tjia (1964, p. 684) provided field evidence of such reversed-symmetry by Tjia (1964, p. 684) as explained by Rod (1966, p. 1164). Shear bands
steps. Both types of steps occur in nature and, in fact, both types were of this orientation, if mistaken for features in the R direction, give the
described by Willis (1923). wrong sense of displacement (Behrmann, 1987, p. 661).
A general terminology for steps on a fault surface devised by Norris Not all sense-of-shear criteria are directly linked to secondary frac-
and Barron (1969, p. 153) defines incongruous steps as those oriented to tures. Any step in the fault surface, whether congruous or incongruous, if it
oppose the displacement and congruous steps as the opposite. Accretion creates a releasing bend, will cause a void or become the site of an
steps consist of gouge or cataclasite plastered to the fault plane, and frac- accretion step (Norris and Barron, 1969, p. 150-153). When a fault sur-
ture steps are small fractures that cut the fault surface (Norris and Barron, face is separated, accretion steps typically break at a high angle to the fault
1969, p. 136). A double bend in the surface of a fault in which a straight surface, leading to congruous fault steps. Accretion steps not found adja-
fault curves and returns to its original trend is a "restraining bend" if it cent to releasing bends may result in either congruous or incongruous steps
causes restraint as the fault slips and is a "releasing bend" if it causes (Norris and Barron, 1969). The apex of a triangular accretion step points
extension or void formation as the fault slips (Crowell, 1974, p. 191). The in the direction of movement of the fault surface to which it is attached
terms were defined for the map-scale geometry of strike-slip faults, but the (Norris and Barron, 1969, p. 159). Experiments by Engelder (1974b) and
concept is useful for any type of fault down to the thin-section scale (Spang Jackson and Dunn (1974, p. 246) showed that accretion steps on one side
and Groshong, 1981; Gamond, 1987). A releasing bend causes dilation as of the fault correspond to grooves on the other side. Blocks plucked from
the fault slips. The dilation may be manifested by tensile fractures at a high the surface leave depressions in the fault surface usually having a steep
angle to the displacement direction or as a zone of many conjugate faults trailing edge and a low-angle leading edge that may be smoothed or
(Sibson, 1987, p. 701-702). Releasing bends form a major site for vein- polished by subsequent fault movement. This is the type example of the
filling ore deposits, even if the main fault trend is barren (Sibson, 1987). smooth-direction test for sense of displacement (Willis, 1923, p. 60-61). A
This type of ore shoot is perpendicular to the fault slip vector. trail was first defined by Willis (1923, p. 57, by analogy to glacial spurs:
Pressure solution is common on the resisting face of a restraining Chamberlain, 1888, p. 245) as a ridge formed in the lee of a projection on
bend (Laubscher, 1979, p. 472; Spang and Groshong, 1981, p. 329; Mar- the fault surface. The ridge may consist of either wall rock or gouge; as the
shak and others, 1982, p. 1018). Pressure solution at restraining bends is ridge extends away from the projection, it decreases in height and may
frequently contemporaneous with fiber growth at releasing bends (Mar- diminish in width. Rutter and others (1986, p. 6) showed a good example
shak and others, 1982, p. 1018). The collapse and fragmentation of a of a gouge trail behind a resistant grain where the trail is formed of
restraining bend by vein-filled radiating tensile fractures is illustrated by fragments of the grain. Shear tails are elongate, asymmetric accumulations
Laubscher (1979, p. 477-478). Recognition that the feature is a restraining of fragments adjacent to large grains (Lucas and Moore, 1986,
bend allows it to be used as a sense-of-shear criterion. Low-angle restrain- p. 100-101). The asymmetry of the accumulations, similar to gouge trails
ing bends may be polished, whereas low-angle releasing bends on the same extending from diagonally opposite sides of the grain, give the sense of
fault are rough (Angelier and others, 1985, p. 352). shear. The geometry is exactly the same as the asymmetric augen that are
common in crystal-plastic fault zones (Simpson and Schmid, 1983,
Fault-surface features produced by secondary fractures and faults are
p. 1281, 1283). Asymmetric gouge lips across voids intersected by the
conveniently classified according to the Riedel system (Petit and others,
fault plane provide a good displacement criterion (Angelier and others,
1983; Petit, 1987, p. 598). T criteria are in the T direction (Fig. 2).
1985, p. 353); the gouge overhangs the void in the direction that the
Microscopic T fractures, called "microscopic feather fractures," have been
opposite wall slipped.
produced experimentally by Friedman and Logan (1970, p. 3418), Dunn
and others (1973, p. 2412), Conrad and Friedman (1976, p. 189), and
Friedman and Higgs (1981). Crescentic T fractures, concave in the direc- STRESS AND STRAIN ANALYSIS OF THE AGGREGATE
tion of slip of the opposing face (Wegmann and Schaer, 1957, p. 492)
have the same geometry as the chatter marks defined by Willis (1923, The first quantitative approach to the interpretation of rock fabrics
p. 57) and Tjia (1972, p. 55). Chatter marks are produced by a rounded was macroscopic strain analysis (Sharpe, 1847; Sorby, 1853). Becker
gouging tool on the opposite face (Chamberlin, 1888; Lawn and Wilshaw, (1893) derived many of the properties of the strain ellipsoid but unfortu-
1975, p. 22), and they may be known as Hertzian fractures after Hertz nately interpreted faults as occurring on the planes of zero elongation. This
(1896), who derived the stress distribution (Lawn and Wilshaw, 1975, interpretation was refuted by Griggs (1935), but strain analysis did not
p. 19; Engelder and Scholz, 1976, p. 157). Wegmann and Schaer (1957, recover as a respectable subject until the classic study of Cloos (1947b).
p. 494) and Laubscher (1979, p. 476) illustrated tool impressions bordered On the microscopic scale, the symmetry analysis of Sander (1911, 1930;
on one side by similar markings. Occasionally the tool will remain in place Knopf and Ingerson, 1938) was the first influential interpretive frame-
(Angelier and others, 1985, p. 352). Prod marks (Tjia, 1972, p. 54) are the work. According to Sander (1930, in Turner and Weiss, 1963, p. 385),
displacement-parallel grooves produced by a tool. "symmetry principles form the soundest basis for correlating the tectonite
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

1348 R. H. GROSHONG, JR.

fabric with the physical factors—stress, strain, movement picture, and so with one another during rotation, the model might not be appropriate
on—concerned in their evolution." This approach has been termed (Siddans, 1976), but the interactions are evidently reduced by pressure
"kinematic analysis" (Friedman, 1964; Friedman and Sowers, 1970, solution or some other means, and so the model works well even in
p. 480). low-porosity rocks (Etheridge and Oertel, 1979; Oertel, 1983). Talbot's
It is important to consider what results are theoretically possible given (1970) strain-measurement technique using the orientations of deformed
the type of data available. All methods are based on observations of veins is a similar method and might be suitable for coarse, platy materials.
permanent deformation features in the rocks, in other words, strain. It is The technique of Mitra (1976) to determine the strain due to transla-
quite obvious in an experimental context that forces and displacements are tion glide in quartz is based on the change in normal strain with direction
measured, and stresses and strains are computed. It is not possible to in the strain ellipsoid. The quartz grains contain rutile needles that are
observe stress directly. Stress may be computed from strain if the relevant buckled or boudinaged depending on their orientation. A plot of final
physical properties of the rock are known, for example, the elastic con- length divided by original length of fiber versus current orientation pro-
stants or the viscosity and strain rate. The strains form the primary meas- duces the strain ellipse, a method confirmed experimentally by Mitra and
urement, and the stress directions and magnitudes are inferred. The same Tullis (1979). The change in shear strain with orientation is the basis of
is true of field data. In situations where the stress and strain tensors do not Wellman's (1962) graphical method which is especially useful for de-
coincide, either because of rock anisotropy or rotational strain, some formed fossils in which the original angular relationships are known. The
methods may be better at predicting stress axes than strain axes, or the method has proved suitable for strains as small as 7% (Groshong, 1972) as
reverse. well as for larger magnitudes.
With deformation, grain centers become more closely spaced in the
Strain Measurement shortening direction, more widely spaced in the extension direction, and
the angles between grain centers change. Ramsay (1967, p. 195-197) first
A variety of techniques have been devised for measuring strain. proposed the nearest-neighbor strain-measurement technique based on this
Compaction of porous rocks is considered first. The packing density change. The method developed by Fry (1979) and implemented by Hanna
(Kahn, 1956) is the summation of the total framework-grain lengths along and Fry (1979) has proved effective in mildly deformed sedimentary rocks
a straight traverse divided by the length of the traverse, given as a percent. (Narahara and Wiltschko, 1986). If the initial distribution of grains were
As a measure of compaction, it is about equivalent to point counting grain truly random (Poisson), then the Fry method might indicate zero strain
area (Coogan and Manus, 1975). To determine compaction, the original even if the strain is large (Fry, 1979); but for most reasonable natural
framework grain-packing density or grain area must be known. This in distributions, there is no problem (Fry, 1979; Onasch, 1986a, 1986b,
turn depends upon the original packing density (a topic first addressed 1987) because the size of the grains precludes spacings closer than one
quantitatively by Sorby in 1908) which ranges from a grain volume of diameter. The Fry method finds the deviatoric strain which can be cor-
52.36% to 74.05% for cubic or rhombohedral packing of uniform spheres rected for volume change if necessary (Onasch, 1986a).
(Graton and Fraser, 1935). Random packing has an average grain volume The calculation of displacement path and strain from fiber directions
of 60%-63% (Graton and Fraser, 1935; Coogan and Manus, 1975), which is treated by Durney and Ramsay (1973), Wickham (1973), Wickham
is a good average for more complex mixtures of equant grains (Fraser, and Anthony (1977), Ramsay and Huber (1983), Casey and others
1935); in very poorly sorted sediments, however, the grain volume may be (1983), Dieterich and Song (1984), and Beutner and Diegel (1985). Fibers
as high as 72% (Beard and Weyl, 1973). The number of grain contacts per or inclusion trails in antitaxial veins and pressure fringes provide the best
grain is a measure of compaction (Taylor, 1950). Random sections available data for determining displacement history. Mitra (1976) used the
through randomly packed spheres produce 0.63 contacts/grain and 47% fractional area of pressure fringe to measure the amount of pressure diffu-
floating grains (no contacts visible) according to Chillingarian and Wolf sion, a useful technique in a closed system.
(1975). Experimentally deposited sand ranges from 1.6 contacts/grain and
17% floating grains (Taylor, 1950) to 0.85 contacts/grain and 46% floating Stress Measurement
grains (Gaither, 1953). In a field study of Wyoming sandstones, Taylor
(1950) found that the contacts per grain went from 1.6 to 5.2 as the depth Stresses cannot be measured directly but may be inferred from a
of burial increased from 2,885 ft to 8,343 ft with a negligible number of material response that is directly controlled by the stress magnitude.
floating grains at all depths. Transgranular extension fracture is an example (Etheridge, 1983, 1984).
The theoretical framework for quantitatively evaluating the impor- Experiments indicate that transgranular extension fracture occurs at a
tance of compaction by pressure solution together with cementation was unique, tensile effective stress that is a property of the material (Paterson,
developed by Rittenhouse (1971a, 1971b) and extended by Mitra and 1978, p. 23), although failure is highly influenced by stress concentrations,
Beard (1980). Houseknecht (1987) presented a practical method for mak- and so a range of tensile strengths is to be expected. The tensile strengths
ing the evaluation from point-count data. Mechanical compaction causes of 7 rocks tabulated by Jaeger and Cook (1979, p. 190) are in the range of
more porosity loss at the early stage of compaction by pressure solution 3.5 to 400 bars, very small numbers compared to the compressive
even if the dissolved material is deposited nearby. Complete loss of poros- strengths. Indirect tensile tests (that is, bending, point load) tend to give
ity by this process has been interpreted for some rocks (Lowry, 1956; larger values (Jaeger and Cook, 1979, p. 191). The hydraulic fracture
Heald, 1956; Thomson, 1959). The compaction rather than cementation strength of a number of rocks ranges from 30 to 266 bars (Rummel, 1987,
component was shown to be more important by Sibley and Blatt (1976) p. 220).
and Wilson and Sibley (1978). The angle between conjugate faults is a function of the differential
The amount of rotation of grains free from mutual interactions de- stress at failure. Mohr (1900) proposed that fracture occurs when the shear
pends directly upon the magnitude of the strain. March (1932; see Oertel, stress on the fracture plane reaches a value that is a function of the normal
1983) developed a model for random markers deformed homogeneously stress, the Mohr envelope on a Mohr diagram. Two assumptions are
with their matrix that predicts the relationship between strain and fabric associated with the Mohr envelope: fault initiation occurs as the Mohr's
orientation. The March method has been applied successfully to experi- circle touches the envelope, and the point of tangency represents the stress
mental phyllosilicate fabrics (Means and Paterson, 1966) and natural slates on the fault plane. The latter assumption means that the dihedral angle
(Oertel, 1970; reviewed in Oertel, 1983). If the phyllosilicates interfere between conjugate faults might be used to infer the differential stress and
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

LOW-TEMPERATURE DEFORMATION MECHANISMS 1349

effective confining pressure for a curved envelope. When all of the stresses p. 870) uses the relative numbers of calcite grains containing one, two, and
are compressive and above about 1 kbar (100 MPa), the experimental three sets of twin lamellae to measure differential stress. A small differen-
Mohr envelope is approximately linear for many rocks (Handin and tial stress on the aggregate will cause only the most favorably oriented
others, 1963, p. 731,743; Brace, 1964, p. 163). The typical dihedral angle crystals to twin, and mainly single twin sets will form, but a large shear
is 55 9 -60° and is close to that predicted from the Mohr envelope in stress will cause twinning in less favorably oriented directions, resulting in
experiments on Solnhofen limestone (Heard, 1960), Repetto siltstone, and two or three sets in some grains. Tested against experimental data, the
Muddy shale (Handin and others, 1963) but differs by an average of +6° in percentage of grains having three twin sets gave the best stress estimate,
Berea sandstone and -6° in Marianna limestone (Handin and others, agreeing to 21% or better with the experimental values of differential stress.
1963). The average differences are +8° to 12° on various rock types in the The later approach of Laurent and others (1981, p. 658), based on finding
low confining pressure tests of Borg and Handin (1966, p. 262) but only the tensor that maximizes the shear stress on every twin set, underestimates
+1° in their high confining pressure tests. the differential stress in their experimental test by a factor of 3-4 or,
The Mohr envelope for rocks is parabolic in the vicinity of the origin alternatively, requires a critical resolved shear stress for calcite of 3 or 4
where the confining pressure is less than 1 kbar (Handin and others, 1963; times the experimental single-crystal value.
Brace, 1964; Corbett and others, 1987). Muehlberger (1961, p. 214) used The minimum differential stress required to cause pressure solution
the concept of a parabolic Mohr envelope to explain fractures with low has never been established but must be very small. Solution-pitted pebbles
dihedral angles as having formed with a 3 tensile. The later triaxial experi- have been found as shallow as 30-40 m (Trurnit, 1968), and sutured
ments of Brace (1964, p. 163) on dry rock, however, demonstrated that stylolites were found at 90 m burial (Schlanger, 1964). Alvarez and others
the dihedral angle is zero when 03 is tensile. There seems to be no experi- (1976) found abundant solution cleavage in folded limestones from north-
mental evidence that the Mohr envelope is continuous into the tensile field. ern Italy in which the coarse-grained calcite was untwinned.
Low-dihedral-angle conjugate faults (~50° or less) evidently represent
small compressive effective stresses. The absolute magnitude depends on Dynamic Analysis and Related Strain-Analysis Techniques
the rock type. The sequence of events in a Riedel shear zone may be a
better micromechanical model for some faults. A very important feature of Dynamic analysis, a term introduced by Turner (1953), is defined by
the Riedel model is the fact that the ultimate fault zone (D shear) is at an Friedman and Sowers (1970, p. 488) as "the use of certain fabric elements
angle of 45° to the far-field a h not 30°, and there need be no conjugate. as criteria for inferring the orientations and relative magnitudes of princi-
Deformation by crystal-plastic glide is related to the differential stress. pal stress axes in the rock at the time these fabric elements develop." The
Handin and Griggs (1951, p. 866) introduced into structural geology the fabric elements are the result of strain, and so dynamic analysis necessarily
metallurgical concept of critical resolved shear stress. Slip occurs on a glide represents inferences based on strain phenomena. In practice, the term is
plane when the shear-stress component in the glide direction reaches the used for methods in which the direction and sense of glide or slip are
critical resolved shear stress. Called "Schmid's law" (Schmid and Boas, known but in which the amount of glide or slip is unknown or not used. In
1950), Handin and Griggs (1951) used it to predict fabric changes due to other words, dynamic analysis represents the inference of stress from strain
large strain. It has been established experimentally that the critical resolved without using strain-magnitude data. Particularly for anisotropic materials,
shear stress for glide is independent of the normal stress across the glide the stress axes so derived may be more accurate than stress axes that are
plane (Turner and others, 1954a; Friedman, 1967) but is, in general, a assumed to be parallel to the strain axes, but this has yet to be demon-
function of temperature and strain rate (Turner and others, 1954a; Higgs strated. The current techniques have arisen rather independently within
and Handin, 1959, p. 276; Carter and Raleigh, 1969; Heard and others, different interpretive traditions, yet fundamentally they are very similar in
1978; Spang and Friedman, 1978). Data on critical resolved shear stress that all are designed to estimate some or all of the properties of a tensor.
are available for calcite (Griggs and others, 1951,1953; Turner and others, The deformed-crystal techniques use the crystal lattice to define an original
1954a; Griggs and others, 1960a; Friedman, 1967), dolomite (Higgs and configuration, and crystal glide systems to determine the direction, sense,
Handin, 1959; Griggs and others, 1960b; Heard and others, 1978; Spang and amount of glide. The fault-slip techniques use slickenlines and sense-of-
and Friedman, 1978), and clinopyroxene (Tullis, 1980). Tullis (1980) shear indicators to define direction and sense of slip. Twin gliding is
found that the average single-crystal critical resolved shear stress for calcite constrained by the crystal lattice to occur along a specific glide line with a
is 100 bars; for dolomite, 1 kbar; and for clinopyroxene, 1.4 kbar. The specific sense of shear. Quartz deformation lamellae are less constrained
values are subject to variation depending on the starting material but, because the lamellae are irrational planes and the amount of translation
clearly, the presence of twins provides a good measure of the minimum glide is not fixed. Fault slip is the least constrained because slip may occur
differential stress. in any direction on any plane and in any amount; for some faults, the slip
Determining the stress in a grain within a multicrystalline aggregate is line may be known but not the slip direction.
a more complex problem. Early in the experimental deformation of rocks, Dynamic analysis based on crystal glide was developed by Turner
the question arose as to whether the fabric was the result of (nearly) (1953), whofirstrecognized that it was possible to invert glide data to infer
homogeneous strain and consequently stress that changed from grain to average stress orientation. In the Turner technique, the principal stress
grain or (nearly) homogeneous stress with strain that changed from grain directions (C and T axes) most favorably oriented to cause twinning in
to grain. By comparing predicted and observed fabrics, Griggs and Miller each twin set are plotted on a stereonet for a large number of grains, and
(1951, p. 860) and Handin and Griggs (1951, p. 864) determined that the the average C and T positions are interpreted as the stress axes most
strain was homogeneous, and so the stress must vary in the differently favorably oriented to produce the observed twinning. Turner (1953) dem-
oriented grains. The deformed oolite experiments of Donath and Wood onstrated that this interpretation was consistent with the experimental
(1976), lower temperature experiments of Schmid and Paterson (500 °C, results, and Friedman (1963) first demonstrated by direct experiment that
1977), and Wood and Holm (1980) confirmed the homogeneity of the the technique works. It should be remembered that in these experiments
strain. Whether the critical resolved shear stress is reached on a glide the stress axes and strain axes are coaxial, and so it is not possible to
system in any given crystal depends on the orientation of the glide system determine which tensor is being estimated. The method was extended to
with respect to the principal stresses and upon the constraints provided by dolomite by Christie (1958) and Crampton (1958) and to plagioclase by
neighboring crystals (Paterson and Turner, 1970). Lawrence (1970). Not until the experiments of Heard and Carter (1968)
A theory based on resolved shear stress (Jamison and Spang, 1976, was the proper technique for dynamic analysis of quartz deformation
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

1350 R. H. GROSHONG, JR.

lamellae understood. Carter and Raleigh (1969) pointed out that dynamic technique has been demonstrated to be accurate to within about 1% strain
analysis techniques could be extended to kinking. in room temperature experiments favoring glide (Groshong, 1974; Gro-
Later work has produced refinements in the original Turner (1953) shong and others, 1984b). It is possible to recognize a multiple deforma-
method. Using only twin sets with high spacing indices produces better tion and, in favorable circumstances, accurately compute the strains
results (Turner and Weiss, 1963, p. 414). Nissen (1964a, p. 901) showed involved (Teufel, 1980). On the other hand, Spiers (1979) in experiments
that C and T axes computed as most favorable for glide on two planes in at 400° found the strain in individual twinned grains to be consistently
doubly twinned crystals produced sharper average axis concentrations twice that predicted from the imposed bulk strain. Spiers (1979, p. 287)
than did C and T axes for singly twinned grains. Venkitasubramanyan suggested that the grains suitably oriented for twinning deformed preferen-
(1971) and Dietrich and Song (1984) have developed other versions of tially, although his photomicrographs show all the grains to be twinned. If
this approach also with improved results. the difference is not an algebraic error (the engineering shear strain used to
The dynamic analysis of faults has developed from several different define the twin strain is twice the tensor shear strain used in the transfor-
and independent directions. For Andersonian conjugate fault pairs, slip on mation equation; for example, Jaeger and Cook, 1979, p. 38), then this
the fault planes should be parallel to the maximum shear stress direction at result suggests a change in the strain mechanism between 25 and 400 °C
the time of formation of the faults. Wallace (1951) appears to have been that should be further investigated.
the first to relate the slip direction to slip parallel to the maximum shear Inspired by Bott's (1959) equations for slip on an arbitrary fault
stress on arbitrarily oriented faults in a true triaxial stressfield.Bott (1959) plane, the first least-squares dynamic analysis for faults was published by
presented a similar analysis for determining the orientation of slip on a Carey and Brunier (1974). Since then a number of other methods have
plane of weakness in an anisotropic body. been published, the differences being mainly in the parameters minimized
A practical dynamic analysis for slip on multiple (non-Andersonian) by least squares. The Carey and Brunier (1974) method minimized the
faults was developed by Compton (1966) independently of these theoreti- angle between the maximum shear strain on the fault and the slickenline
cal treatments. Referring to Hubbert (1951), Compton noted that slip is direction, an approach also used by Etchecopar and others (1981) in an
most likely on lines at 30° to aj. In a stress state for which 02 = <73, a plot iterative solution. Another approach is to minimize the components of the
of the slip linears on a stereonet is a small circle of points arranged around calculated shear stress perpendicular to the slip line (Angelier and Goguel,
a 1 (Compton, 1966, p. 1371), confirmed by Aleksandrowski (1985, 1979; Angelier, 1979; Angelier and others, 1982). The method of Michael
p. 77). Compton suggested that for 02 <73 the distribution would be (1984) uses the idea that the magnitude of the shear stress in the active slip
elliptical, and the examples of Aleksandrowski (1985) show that, although direction should be the same on each fault and so minimizes the difference
this is correct, it is nearly impossible to recognize in practice. Compton between the computed shear stress on each slip line and a constant refer-
(1966, p. 1373) also performed an analysis very similar to a Turner ence value. Because the maximum shear stress directions are controlled by
dynamic analysis but with the compression axis at 30° to the fault. Use of the ratios of the three principal stresses (Wallace, 1951; Bott, 1959), all of
the sense-of-slip data clearly indicates theCT[direction. Arthaud (1969) the techniques allow calculation of the principal stress ratios.
developed a different graphical technique based on the stereonet intersec- Dynamic analysis techniques for fault slip directions are very similar
tions of the planes containing the slip line and the pole to the fault. This to earthquake focal-mechanism solutions. Pavoni (1980) used a focal-
relates to Compton's method as the beta pole fold axis relates to the pi mechanism-solution method to determine the compressional and exten-
pole. Both methods seem most suitable for biaxial stress tensors. Aleksan- sional quadrants and best compression and extension axes for fault slip
drowski (1985) modified Arthaud's method to improve the resolution of data from outcrop. Pfiffner and Burkhard (1987) have extended this ap-
the three principal axes and computed the stress ratio from the result. proach to petrofabric analysis. None of the fault dynamic analysis methods
Spang (1972) introduced thefirstnumerical technique for performing has been verified experimentally, but all produce reasonable results. The
a dynamic analysis based on twin glide in calcite. The Spang numerical Angelier (1979) Crete data set gave nearly the same answer when recom-
dynamic analysis technique (NDA) is produced by transforming with the puted by the later methods of Angelier and others (1982, p. 615), Michael
tensor transformation equation the Turner C and T axes for each twin set, (1984, p. 11520), and Pfiffner and Burkhard (1987). The method of
taken as axes of unit length, into the same coordinate system, averaging the Pfiffner and Burkhard (1987) has also been compared to twin-strain data
components, and finding the principal values of the resulting 3 x 3 matrix. computed by the experimentally verified strain-gauge method and found
The principal directions correspond to the centers of the C and T distribu- to give about the same axis directions. It seems safe to conclude that all the
tions produced graphically (Spang, 1972). In experimental results, the published tensor-estimation techniques give similar principal axis direc-
magnitudes of the principal NDA values correlate with the tightness of the tions. It has yet to be demonstrated whether the differences reflect merely
C-T axis distributions (Groshong, 1974). Spang and Van Der Lee (1975) the different approaches to accommodating natural variation or reflect real
extended the NDA method to quartz deformation lamellae and dolomite differences between the stress and strain tensors.
twin lamellae. A new method based on maximizing the shear stress on
twinned planes while minimizing the shear stress on untwinned planes TEXTURAL INTERPRETATION OF THE AGGREGATE
(Laurent and others, 1981) is as successful as the Spang technique in
finding the principal axis directions. The aggregate fabric in low-temperature deformation is characterized
Groshong (1972) introduced the least-squares strain gauge technique by simultaneously active deformation mechanisms and heterogeneous
for computing strain in a calcite aggregate from measurements on twin strain. The total strain may be the result of steady-state deformation by
lamellae. The method traces to the calculation by Handin and Griggs more than one mechanism or may represent a time sequence of mecha-
(1951, p. 867-868) of the strain produced by complete twinning in calcite nisms in pre-failure strain.
from which Conel (1962) derived the equation for the strain in a partially
twinned grain. The strain in each grain is related to that in the other grains Strain Partitioning
by the strain-transformation equation (Jaeger and Cook, 1979), which is
solved by least squares for the strain in the aggregate (Groshong, 1972, Strain partitioning is the subdivision of the total strain at some scale
p. 2029). The least-squares technique is a multiple linear regression that into components produced by different mechanisms (Groshong and others,
simultaneously minimizes the difference between grain-strain components. 1984a). Perhaps the first quantitative field study of low-temperature strain
The result is the magnitude and orientation of the strain tensor. The partitioning into different mechanisms at the same scale was that of Nissen
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

LOW-TEMPERATURE DEFORMATION MECHANISMS 1351

(1964b), who found that the plastic twinning strain in elliptical crinoid does not increase appreciably with total strain. The folding strain is pre-
columnals in a graywacke matrix was insufficient to account for the shape dominantly by pressure solution. A natural suite of rocks (Friedman, 1967,
change. Nissen attributed the discrepancy to an initial ellipticity of the p. 191) showed greater fracture strengths in presumably more deformed
columnals, but he mentioned that pressure solution was active, and his samples from within folds than in presumably less deformed samples from
photos show stylolitic boundaries on the long sides of the ellipses. Engelder horizontal beds away from the folds. Thus, as a result of work hardening,
and Engelder (1977) and Engelder (1979) obtained the same result in the current strength of the rock may be a function of the total strain.
another area and demonstrated that pressure solution at the crinoid
boundaries explained the discrepancy. Nickelsen (1966) discovered Flow Laws
deformed fossils in an apparently undeformed matrix, a relationship he
attributed to strain partitioning between microfolding and extension mi- Within the work-hardening regime, the aggregate flow law must be a
crofracture in the fossils and lateral compaction in the matrix. The parti- combination of flow laws for the individual mechanisms (Groshong,
tioning of strain between grain rotation and grain strain is important 1975b; Mitra, 1976). Flow laws for individual mechanisms might be
because the grain strain will be less than the total strain, and for compac- interpreted from experimentally derived deformation-mechanism maps
tion the grain strain could be zero (Borradaile, 1981). (Stacker and Ashby, 1973; Frost and Ashby, 1982) and can be combined
The strain partitioning changes as strain increases. Donath and others by using standard strength-of-materials concepts as given by, for example,
(1971) and Tobin and Donath (1971) have provided the only quantitative Jaeger and Cook (1979, p. 314-325).
experimental analysis of the deformation mechanism transitions that occur Crystal-plastic dislocation glide mechanisms follow an exponential
with increasing strain before faulting. They defined the mesoscopic defor- flow law (Cottrell, 1964; Schmid, 1982). Dislocation glide is resisted by
mation modes as extension fracture, brittle fault, "ductile" fault, and uni- impurities and the presence of grain boundaries, and so small grains de-
form flow (Donath and others, 1971, p. 1447). In their terminology, a forming by this mechanism are stronger than big grains (Robertson, 1955;
brittle fault is cohesionless and a "ductile" fault cohesive. They showed Handin and Hager, 1957; Paterson, 1958; Brace, 1961; Friedman, 1967;
that at 25 °C the modes were controlled by confining pressure (1-2 kbar Hugman and Friedman, 1979). The Hall-Petch law in metallurgy (Nicolas
range) and total strain (2%-20% range). The microscopic fabric elements and Poirier, 1976) states that the excess hardening is related to the inverse
that correlated with the macroscopic modes were (Tobin and Donath, square root of grain size. Pressure enhances this effect by the suppression of
1971, p. 1468) undeformed grains, grains with twin lamellae, grains with microfractures, whereas temperature reduces the effect (Olsson, 1974).
fractures and twin lamellae, and transgranular fractures. Tobin and Increasing temperature reduces work hardening due to dislocation tangles,
Donath (1971, p. 1472) found that as the strain increased, the relative marked by the reduction of yield strength in materials deforming by glide
proportion of these elements changed in conjunction with the change in (Griggs and others, 1951; Handin and Hager, 1958; Heard, 1963). Reduc-
macroscopic mode. Uniform flow is characterized by undeformed and tion in strain rate also significantly reduces the yield strength (Heard, 1963;
twinned grains, with the proportion being a function of the total strain. Donath and Fruth, 1971).
Grains with twins and fractures become important as microfaulting begins Grain-boundary diffusion mechanisms follow a linear flow law
and transgranular fractures characterize the region of faulting. Within the (Coble, 1963). Elliott (1973) proposed that pressure solution is analogous
regime of brittle faulting, a moderate percentage of undeformed grains to grain-boundary diffusion (Coble creep) for which the strain rate should
persist but are nearly absent in the low-temperature cohesive faulting be directly proportional to the differential stress and inversely proportional
regime. Primary undeformed grains must be distinguished from recrystal- to the temperature and the square of grain size. The effect of temperature is
lized undeformed grains in higher-temperature deformation. quite small (Rutter, 1983), however. Kerrich and others (1977, p. 248)
Another way to express the difference between mesoscopic faulting found in afieldexample that pressure solution had a greater effect on small
and flow is in terms of the strain heterogeneity, first shown by Donath and grains than on large grains, substantiating the grain-size effect.
Wood (1976, PI. 1) as contours of strain magnitudes on experimentally The fault-initiation stress increases with confining pressures but de-
deformed oolitic limestone and later by Wood and Holm (1980) as the creases with temperature (Griggs, 1936; Griggs and others, 1951; Handin
correlation coefficient on a plot of the long axis versus the short axis of and Hager, 1958; Paterson, 1958, PL 1; Griggs and Handin, 1960; Handin
oolite strain ellipses. The homogeneity of strain increases with increasing and others, 1963). A completely satisfactory fracture criterion has yet to be
temperature and confining pressure, and little effect of strain rate is seen. In developed. The fracture stress should not depend upon the choice of
all experiments, the tendency to form fault zones increased with total coordinate system and therefore should be a function of the stress invar-
strain. Gray (1981, p. 231 -232) found a somewhat analogous result in the iants (Jaeger and Cook, 1979, p. 23-24). Handin and others (1967)
field for deformation by pressure solution. The early stage of a single attempted to find a single stress-invariant (Jaeger and Cook, 1979,
deformation was by grain-scale stylolites and produced a weak slaty fabric p. 23-24) fracture criterion that would apply to compression, extension,
that is truncated by zones of more intense pressure solution that form and torsion tests by plotting the octahedral shear stress (second deviatoric
spaced cleavage. invariant) against the mean stress (first invariant) but did not obtain con-
The change in mesoscopic mode from flow to faulting with increasing sistent results. Cherry and others (1968) defined a more complex function
strain is to be expected in a work-hardening material. The grain-scale of octahedral shear stress, the third stress invariant and the third deviatoric
strain may reflect either the total strain when deformation ended or the invariant that gives more consistent results; Mogi (1971) showed that a
ductility of the rock, that is, the strain at which deformation shifts to a curve relating the octahedral shear stress to the octahedral normal stress
mesoscopic mode. It is clear in experiments, at least, that when the meso- provided an acceptable criterion; and White (1973) determined a function
scopic failure mode is faulting, the grain-scale flow mechanisms are of the three invariants. Kirby (1980, p. 6355-6356) found that a plot of
supplanted by fault slip. If the mesoscopic failure mode is solution cleav- differential stress versusCT3caused data from different types of tests to fall
age, it is not yet certain whether grain-scale deformation continues. The on a single curve.
average glide strains in three small folds from one outcrop (Groshong, Many of the experiments and field examples previously cited show
1975b, p. 1367) are nearly constant in magnitude (2.5% ± 2.0%) and direc- tensile microcracks in the vicinity of faults, and experiments often show
tion (about layer parallel), even though the interlimb limb angles of 150°, that microfracturing and dilation occur just prior to faulting (Gramberg,
90°, and 0° imply large differences in total strain. This indicates that the 1965; Brace and others, 1966; Scholz, 1968a, p. 4794; 1968b; Wawersik
crystal-plastic strain in this example is a measure of the ductility because it and Brace, 1971, p. 74-77; Edmond and Paterson, 1972; Paterson, 1978,
1352 R. H. GROSHONG, JR.

p. 112-137). This is a work-softening process that causes deformation to Deformation Fronts and Deformation Mechanism Associations
be localized in the fault zone. Griffith (1921) demonstrated that cracks
form flaws that greatly reduce the stress needed for tensile fracture propa- It has long been known that the character of rock deformation
gation and extended the concept in 1924 to plane-strain shear failure changes from sedimentary to metamorphic from the external to the inter-
related to obliquely oriented cracks. The Griffith criterion has been modi- nal parts of many mountain belts, and from the shallow levels to the
fied for closed cracks by Brace (1960) and McClintock and Walsh (1962) deeper levels (Heim, 1878, as discussed by Milnes, 1979). Van Hise (1898)
and extended to three dimensions by Murrell (1963) as discussed by defined an upper zone of fracture, an intermediate zone of fracture and
Jaeger and Cook (1979, p. 101-106). The extended theory predicts a flow, and a lower zone of flow. This type of categorization remains impor-
parabolic Mohr envelope for which the uniaxial compressive strength is 12 tant in tectonic analysis and in the interpretation of structural provenance.
times the tensile strength. The extended theory plots as a straight line A deformation front was first explicitly defined for quartzite in the
when all stresses are compressive and predicts that 02 will be important, central Appalachians by Fellows (1943, p. 1415) as occurring where the
all in reasonable accord with observations (Jaeger and Cook, 1979, quartzites change from having unoriented grains to having grains that
p. 190-192; Paterson, 1978, p. 52-70). show a preferred orientation in shape and crystal lattice orientation. Mitra
A problem with theories of failure based on tensile microcracks is (1987, p. 588) showed that in the same area the tectonite front for carbon-
how the cracks link together to form a fault. Numerous experiments have ates is more external in the mountain belt than the tectonite front for
shown that tensile cracks form parallel to ai or extend parallel to o\ from siliciclastic rocks. Also in the same area, Cloos (1947b, p. 911) defined a
oblique cracks (Brace and Bombolakis, 1963; Gramberg, 1965, p. 48; front for cleavage in limestone as occurring where the oolite extension first
Hoek and Bieniawski, 1965, p. 148,150; Lajtai, 1971, p. 142,146-147). reaches 20% and the cleavage front for shale as visible at a more external
Peng and Johnson (1972) proposed that a fault develops from the tensile position. (In this area, the cleavage in limestone is the result of the grain
cracks by buckling of the thin beams between the cracks after the cracks shape fabric, not pressure solution.) A deformation front could be defined
have become long enough. The concept is appealing because brittle buck- as the first occurrence of any deformation mechanism.
ling is an unstable phenomenon, and the granulation of the beams would Blake and others (1967, p. C7) introduced the concept of naming the
form gouge almost instantly. rocks between fronts in a classification currently used for subduction-zone
Faulting in porous materials leads to porosity collapse and work rocks. They defined "textural zones of progressive metamorphism" (in-
hardening in the fault zone. This wasfirstsuggested by Aydin and Johnson cluding deformation) for graywacke: zone 1 appears unmetamorphosed
(1978) based on their observation that increased displacement across fault and shows no evidence of cataclasis (or mylonitization), either in outcrop
zones in porous sandstone was accompanied by the formation of new, or hand lens; zone 2 shows well-developed platy cleavage in outcrop and is
small-displacement faults in intact rock. The hardening occurs because of clearly mylonitic (they used the term "cataclastic" in the sense of Lap-
cataclastic porosity loss (Aydin and Johnson, 1978; Underhill and Wood- worth) under the hand lens; zone 3 is completely recrystallized to quartz-
cock, 1987). Work hardening due to porosity loss by grain rotation and mica schist.
dewatering in clay-rich sediments is interpreted by Moore and Byrne The term "deformation regime" appears to have been introduced by
(1987) to be the cause of fault-zone widening during subduction and to Kerrich and Allison (1978) without explicit definition but in the context of
lead to the formation of mélange. ranges of environmental parameters in which deformation occurs by dif-
After a fault has formed, further displacements may be by frictional ferent mechanisms. Schmid (1982, p. 97) stated that deformation regimes
sliding. Bedding-plane slip is also a mechanism of major importance for are defined by empirical flow laws and diagnostic microstructural im-
which slip on a pre-existing surface is significant. Frictional sliding is a prints. The regimes identified by Schmid (1982) are cataclastic flow, low-
function of the normal and shear stress and can be represented by a linear temperature plasticity, power-law creep, and grain-size sensitive creep.
Mohr envelope. The coefficient of friction is the slope of the line. Byerlee Kerrich and Allison (1978) treated pressure solution as a separate regime.
(1978) divided frictional sliding into three regions: low pressure having Mechanisms in the regimes just identified can, however, occur simultane-
normal stresses of up to 50 bars, intermediate pressure up to 1000 bars, ously in the same rock. For example, low-temperature plasticity is inti-
and high pressure up to 17 kbar or more. In the low-pressure region, mately related to cataclasis under a range of laboratory conditions
friction is dominated by the roughness of the sliding surfaces, and conse- (Donath and others, 1971), and both types of features are found in the
quently the lithology is important, with the coefficient of friction ranging same samples in the field (Spang and Brown, 1981; Mitra, 1987, p. 583).
from 0.3 to 10 (Logan and others, 1972). In the intermediate- and high- Pressure solution and low-temperature plasticity also occur together
pressure range, the maximum friction coefficient is relatively independent (Groshong, 1975b; Kerrich and others, 1977).
of rock type. In the intermediate range, the value is 0.85 and in the high- A deformation-mechanism association is here defined as the deforma-
pressure range is 0.6 plus a constant of 0.5 divided by the normal stress. In tion mechanisms formed under the same environmental conditions in the
the high-pressure regime, much lower coefficients of friction are found for same rock type. Environmental conditions are temperature, confining
vermiculite, montmorillonite, and illite (Byerlee, 1978, p. 624). pressure, differential stress, strain rate, pore fluid type, and pore pressure. A
Based on the results of Donath (1970), Donath and others (1971), rock deformation experiment produces a single deformation-mechanism
and Tobin and Donath (1971), it might be possible to quantitatively infer association in the sample. Five associations are defined, based on the
stress and strain as given on ductility-depth curves (Handin and others, nature of the deformation of the framework minerals. In Association I,
1963) from the strain partitioning, but this has not yet been seriously hydroplastic deformation, the framework minerals are undeformed; the
attempted. When the ductile limit has been reached, the deformation is mechanism of deformation is cohesionless particulate flow in the termi-
usually dominated by a single mechanism, either pressure solution or fault nology of Borradaile (1981). Large strain in soft grains, such as unlithified
slip for which a steady-state flow law might be appropriate. pellets in limestone or glauconite in sandstone, is allowed: the crystals
LOW-TEMPERATURE DEFORMATION MECHANISMS 1353

within the grains are not deformed by the strain of the grain. The Associa- assistance. I greatly appreciate the hard work and valuable suggestions
tion I-II boundary is the crystal deformation front. In Association II, from the reviewers Terry Engelder, Carol Simpson, Dave Wiltschko, Joe
brittle-framework deformation, crystal deformation occurs by truncation Benson, Mike Lesher, and Greg Guthrie. Work supported by National
due to fracture or faulting (IIF) or pressure solution (IIS) or both (IIFS). Science Foundation Grant EAR-8402915 has contributed directly to the
Strain due to crystal-plastic glide is in the range of 0%-2%. The crystal- interpretations.
plastic component is characterized by slight undulatory extinction and rare
deformation lamellae. The Association II-III boundary is the crystal-
plastic front. Association III, semi-brittle-framework deformation, is de- R E F E R E N C E S CITED
fined by obvious crystal-plastic glide that is insufficient to produce a Abdel-Gawad, M., Bulau, J., and Tittmann, B., 1987, Quantitative characterization of microcracks at elevated pressure:
grain-orientation fabric, a range of about 2%-15% strain (for example, Journal of Geophysical Research, v. 92, p. 12,911-12,916.
Adams, F. D., 1910, An experimental investigation into the action of differential pressure on certain minerals and rocks
Fig. 1). The crystal plastic component may be characterized by obvious employing the process suggested by Professor Kick: Journal of Geology, v, 18, p. 489-525.
undulatory extinction and/or abundant deformation lamellae. Mineral Adams, F. D., and Coker, E. G., 1910, An experimental investigation into the flow of rocks: American Journal of Science,
Fourth Series, v. 29, p. 465-487.
truncation may occur by faulting and possibly by tension or extension Adams, F. D., and Nicolson, J. T., 1901, An experimental investigation into the flow of marble: Royal Society of London
Philosophical Transactions, ser. A, v. 195, p. 363-401.
fracture (IIIF), pressure solution (IIIS), or both (IIIFS). The Association Adams, S. F., 1920, A microscopic study of vein quartz: Economic Geology, v. 15, p. 623-664.
III-IV boundary is the tectonite front. In Association IV, tectonites, a Aleksandrowski, P., 1985, Graphic determination of principal stress directions for slickenside populations: An attempt to
modify Arthaud's method: Journal of Structural Geology, v. 7, p. 73-82.
deformation-induced crystallographic fabric is present. Association IV Alvarez, W., Engelder, T., and Lowrie, W., 1976, Formation of spaced cleavage and folds in brittle limestone by
dissolution: Geology, v. 4, p. 698-701.
may be subdivided into (IVG), glide tectonites, in which the crystallo- Alvarez, W„ Engelder, T., and Geiser, P. A., 1978, Classification of solution cleavage in pelagic limestones: Geology, v. 6,
graphic fabric is caused by glide mechanisms and the original mineral p. 263-266.
Anderson, E. M., 1905, The dynamics of faulting: Edinburgh Geological Society, v. 8, p. 387-402.
boundaries are preserved; and (IVR), recrystallization tectonites, in which a 1942, The dynamics of faulting and dyke formation with applications to Britain (1st edition): London, England,
Oliver and Boyd, 191 p.; revised 2nd edition, 1951,206 p.
crystallographic fabric is present but the grain boundaries are no longer Anderson, J. J., and Everett, J. R., 1965, Mudcrack formation studied by time-lapse photography [abs.]: Geological
original. Association V is annealed, having no more than a weak crystallo- Society of America Special Paper 82, p. 4-5.
Anderson, J. L., Osborne, R. H., and Palmer, D. R., 1980, Pedogenesis of cataclastic rocks within the San Andreas fault
graphic fabric and the original grain boundaries obliterated. Associations zone of southern California, U.S.A.: Tectonophysics, v. 67, p. 221-249.
Anderson, T. B,, 1968, The geometry of a natural orthorhombic system of kink bands: Geological Survey of Canada Paper
I—III fall in the low-temperature regime in which a sedimentary rock 68-52, p. 200-220.
continues to maintain its sedimentary texture. Deformation-mechanism Angelier, J., 1979, Determination of the mean principal directions of stresses for a given fault population: Tectonophysics,
v. 56, p. T17-T26.
associations are related to rock type, and so in any given environment of Angelier, J., and Goguel, J-, 1979, Sur une methode simple de determination des axes principaux des contraintes pour une
population de failles: Comptes Rendus Académie de Science Paris, sér. D, v. 288, p. 307-310.
deformation, the boundary between two zones will not in general occur at Angelier, J., Tarantola, A-, Valette, B., and Manoussis, S., 1982, Inversion of field data in fault tectonics to obtain the
the same location for different lithologies. regional stress—I. Single phase fault populations: A new method of computing the stress tensor: Royal Astronomi-
cal Society Geophysical Journal, v. 69, p. 607-621.
Angelier, J., Colletta, B., and Anderson, R. E., 1985, Neogene paleostress changes in the Basin and Range: A case study at
The primary control on the location of the tectonite front appears to Hoover Dam, Nevada-Arizona: Geological Society of America Bulletin, v. 96, p. 347-361.
be temperature. Groshong and others (1984a) found this transition to be Ardell, A. J„ Christie, J. M., Kirby, S. H., and McCormick, J. W., 1976, Electron microscopy of deformation structures in
quartz, in Proceedings of the Electron Microscopy and Analysis Group: Bristol, England, Institute of Physics,
rather sharp and to occur at the vitrinite reflectance (R0) of 3.5% in a University of Bristol.
Arthaud, F„ 1969, Détermination graphique des directions de raccourcissement, d'allongement et intermédiaire d'une
relatively clean, coarse-grained limestone in the eastern Helvetic Alps. This population de failles: Société Géologique de France Bulletin, v. 11, p. 729-737.
corresponds to a temperature of about 175 °C based on the well-standard Arthaud, F., and Mattauer, M., 1969, Examples de stylolites d'origine tectonique dans le Languedoc, leur relation avec la
tectonique cassante: Société Géologique de France Bulletin, v. 11, p. 738-744.
reflectance curve of Bostick (1974, p. 9). Low-temperature metamorphic 1972, Sur l'origine tectonique de certains joints stylolitiques parallèles a la stratification; leur relation avec une
phase de distension (example du Languedoc): Société Géologique de France Bulletin, v. 14, p. 12-17.
stages are subdivided by Tissot and Welte (1978) into diagenesis, having Aydin, A., 1977, Faulting in sandstone, Utah [Ph.D. dissert.]: Stanford, California, Stanford University, 246 p.
an upper limit of Rq = 0.5; catagenesis, upper limit = 2.0; and metagene- 1978, Small faults formed as deformation bands in sandstone: Pure and Applied Geophysics, v. 116, p. 913-930.
Aydin, A., and Johnson, A. M., 1978, Developments of faults as zones of deformation bands and as slip surfaces in
sis, upper limit =» 4.0, above which greenschist metamorphism begins. This sandstone: Pure and Applied Geophysics, v. 116, p. 932-942.
Aydin, A., and Reches, Z., 1982, Number and orientation of fault sets in the field and in experiments: Geology, v. 10,
scale can be correlated to coal rank (Teichmiiller and Teichmtiller, 1981), p. 107-112.
spore and pollen coloration (Gray and Boucot, 1975), conodont color Bahat, D., 1979, Theoretical considerations on mechanical parameters of joint surfaces based on studies on ceramics:
Geological Magazine, v. 166, p. 81-92.
(Epstein and others, 1977; Rejebian and others, 1987), chitinozoan, grap- !986a, Criteria for the differentiation of en echelons and hackles in fractured rocks: Tectonophysics, v. 121,
p. 197-206.
tolite, and scolecodont reflectance (Bertrand and Heroux, 1987), illite 1986b, Joints and en echelon cracks in middle Eocene chalks near Beer Sheva, Israel: Journal of Structural
crystallinity (Kiibler, 1968), and certain mineral reactions (Kiibler, 1968). Geology, v. 8, p. 181-190.
1987a, Correlation between fracture surface morphology and orientation of cross-fold joints in Eocene chalks
Temperature and time both clearly affect these indicators of metamor- around Beer Sheva: Tectonophysics, v. 136, p. 323-333.
1987b, Jointing and fracture interactions in middle Eocene chalks near Beer Sheva, Israel: Tectonophysics, v. 136,
phism (Bostick, 1974; Dow, 1978). Confining pressure alone does not p. 299-321.
have a significant effect (Bostick, 1974; Epstein and others, 1977), al- Bahat, D., and Engelder, T., 1984, Surface morphology on cross-fold joints of the Appalachian Plateau, New York and
Pennsylvania: Tectonophysics, v. 104, p. 299-313.
though the presence of water affects conodont coloration (Rejebian and Bailey, S. W., Bell, R. A., and Peng, C. J., 1958, Plastic deformation of quartz in nature: Geological Society of America
Bulletin, v. 69. p. 1443-1466.
others, 1987), and strain alone can alter the illite crystallinity (Flehmig and Bailly, P. A., 1954, Presence de microstylolites dans des pegmatites et des lentilles de quartz: Société Géologique de France
Langheinrich, 1974). Considerably more work should be done to deter- Bulletin, v. 3, p. 299-301.
Baker, P. A., Kastner, M., Byerlee, J. D., and Lockner, D. A., 1980, Pressure solution and hydrothermal recrystallization
mine the relationship between deformation-mechanism associations and of carbonate sediments—An experimental study: Marine Geology, v. 38, p. 185-203.
Balk, R., 1937, Structural behavior of igneous rocks: Geological Society of America Memoir 5, 177 p.
low-temperature metamorphism. Barrett, P. J., 1964, Residual seams and cementation in Oligocene shell calcarenites, Te Kuiti Group: Journal of
Sedimentary Petrology, v. 34, p. 523-531.
Bartlett, W. L., Friedman, M., and Logan, J. M., 1981, Experimental folding of rocks under confining pressure, Part IX.
Wrench faults in limestone layers: Tectonophysics, v. 79, p. 255-277.
ACKNOWLEDGMENTS Bates, R. L., and Jackson, J. A., 1980, Glossary of geology, (2nd edition): Falls Church, Virginia, American Geological
Institute, 751 p.
Bathurst, R.G.C., 1958, Diagenetic fabrics in some British Dinantian limestones: Liverpool and Manchester Geological
I would like to thank the GSA editors William A. Thomas and Journal, v. 2, p. 11-36.
Robert D. Hatcher, Jr., for inviting me to write this paper and Richard P. Batzle, M. L., Simmons, G-, and Siegfried, R. W., 1980, Microcrack closure in rocks under stress: Direct observation:
Journal of Geophysical Research, v. 85, p. 7072-7090.
Nickelsen for providing the original inspiration. The invaluable Interna- Beach, A., 1974, A geochemical investigation of pressure solution and the formation of veins in a deformed greywacke:
Contributions to Mineralogy and Petrology, v. 46, p. 61-68.
tional Tectonic Dictionary by John Dennis and John himself were of great 1975, The geometry of en-echelon vein arrays: Tectonophysics, v. 28, p. 245-263.
1354 R. H. GROSHONG, JR.

1977, Vein arrays, hydraulic fractures and pressure-solution structures in a deformed flysch sequence, S.W. Bruhn, R. L., and Pavlis, T. L., 1981, Late Cenozoic deformation in the Matanuska Valley, Alaska: Three-dimensional
England: Tectonophysics, v. 40, p. 201-225. strain in a forearc region: Geological Society of America Bulletin, v. 92, p. 282-293.
1979, Pressure solution as a metamorphic process in deformed terrigenous sedimentary rocks: Lithos, v. 12, Bûcher, W. H., 1920, The mechanical interpretation of joints: Journal of Geology, v. 28, p. 707-730.
p. 51-58. Burg, J. P., and Harris, L. B., 1982, Tension fractures and boudinage oblique to the maximum extension direction: An
Beard, D. C., and Weyl, P. K., 1973, Influence of texture on porosity and permeability of unconsolidated sand: American analogy with Lüders' bands: Tectonophysics, v. 83, p. 347-363.
Association of Petroleum Geologists Bulletin, v. 57, p. 349-369. Burg, J. P., and Inglesias Ponce de Leon, M., 1985, Pressure-solution structures in a granite: Journal of Structural Geology,
Becker, G. F., 1893, Finite homogeneous strain, flow and rupture of rocks: Geological Society of America Bulletin, v. 4, v. 7, p. 431-436.
p. 13-90. Burgers, W. G., and Burgers, J. M., 1935, First report on viscosity and plasticity: Koninklijk Nederlandse Akademie van
Beere, W., 1978, Stresses and deformation at grain boundaries: Royal Society of London Philosophical Transactions, Wetenschappen Verhandelingen Series 1, v. 15, p. 1-256.
ser. A, v. 288, p. 177-196. Bustin, R. M., 1983, Heating during thrust faulting in the Rocky Mountains: friction or fiction?: Tectonophysics, v. 95,
Behrmann, J. H., 1985, Crystal plasticity and superplasticity in quartzite: A natural example: Tectonophysics, v. 115, p. 309-328.
p. 101-129. Buxton, T. M., and Sibley, D. F., 1981, Pressure solution features in a shallow buried limestone: Journal of Sedimentary
— ~ — 1 9 8 7 , A precautionary note on shear bands as kinematic indicators: Journal of Structural Geology, v. 9, Petrology, v. 51, p. 19-26.
p. 659-666. Byerlee, J. D., 1978, Friction of rocks: Pure and Applied Geophysics, v. 116, p. 615-626.
Bell, J. F., 1941, Morphology of mechanical twinning in crystals: American Mineralogist, v. 26, p. 247-261. Byerlee, J. D., and Brace, W. F., 1969, High-pressure mechanical instability in rocks: Science, v. 164, p. 713-715.
Bell, T. H., 1978, The development of slaty cleavage across the Nackara Arc of the Adelaide geosyncline: Tectonophysics, Byerlee, J., Mjachkin, V., Summers, R., and Voevoda, O., 1978, Structures developed in fault gouge during stable sliding
v. 51, p. 171-201. and stick-slip: Tectonophysics, v. 44, p. 161-171.
Bell, T. H., and Etheridge, M. A., 1973, Microstructure of mylonites and their descriptive terminology: Lithos, v. 6, Byrne, T., 1984, Structural geology of mélange terranes in the Ghost Rocks Formation, Kodiak Islands, Alaska, in
p.337-348. Raymond, L. A., ed., Mélanges, their nature, origin, and significance: Geological Society of America Special Paper
Berthe, D., Choukroune, P., and Jeqouzo, P., 1979, Orthogneiss, mylonite, and non-coaxial deformation of granites: The 198, p. 21-52.
example of the South Armorican shear zone: Journal of Structural Geology, v. 1, p. 31-42. Carannante, G., and Guzzetta, G., 1972, Stiloliti e sliccoiiti come meccanismo di deformayione delle masse rocciose:
Bertrand, R., and Heroux, Y., 1987, Chitinozoan, graptolite, and scolecodont reflectance as an alternative to vitrinite and Bollettino della Società dei Naturalisti in Napoli, v. 81, p. 157-170.
pyrobitumen reflectance in Ordovician and Silurian strata, Anticosti Island, Quebec, Canada: American Associa- Carey, E., and Brünier, B., 1974, Analyse théorique et numérique d'un Modèle mécanique élémentaire appliqué à l'étude
tion of Petroleum Geologists Bulletin, v. 71, p. 951-957. d'une population de failles: Comptes Rendus Académie de Science Paris, sér. D, v. 279, p. 891-894.
Beutner, E. C., 1978, Slaty cleavage and related strain in Martinsburg Slate, Delaware Water Gap, New Jersey: American Carlson, S. R., and Wang, H. F., 1986, Microcrack porosity and in situ stress in Illinois borehold UPH 3: Journal of
Journal of Science, v. 278, p. 1-23. Geophysical Research, v. 91, p. 10421-10428.
1980, Slaty cleavage unrelated to tectonic dewatering: The Siamo and Michigamme slates revisited: Geological Carson, B., and Berglund, P. L., 1986, Sediment deformation and dewatering under horizontal compression: Experimental
Society of America Bulletin, Part I, v. 91, p. 171-178. results: Geological Society of America Memoir 166, p. 135-150.
Beutner, E. C., and Charles, E. G., 1985, Large volume loss during cleavage formation, Hamburg sequence, Pennsylvania: Carson, B., von Huene, R., and Arthur, M., 1982, Small-scale deformation structures and physical properties related to
Geology, v. 13, p. 803-805. convergence in Japan Trench slope sediments: Tectonics, v. 1, p. 277-302.
Beutner, E. C., and Diegel, F. A., 1985, Determination of fold kinematics from syntectonic fibers in pressure shadows, Carter, N. L., 1968, Row of rock-forming crystals and aggregates, in Riecker, R. E., ed., Rock mechanics seminar,
Martinsburg Slate, New Jersey: American Journal of Science, v. 285, p. 16-50. Volume 2: U.S. Air Force Cambridge Research Laboratories, Clearinghouse for Federal Science and Technology
Beutner, E. C., Jancin, M. D., and Simon, R. W., 1977, Dewatering origin of cleavage in light of deformed calcite veins Information AD669 376, p. 509-594.
and clastic dikes in Martinsburg slate, Delaware Water Gap, New Jersey: Geology, v. 5, p. 118-122. 1971, Static deformation of silica and silicates: Journal of Geophysical Research, v. 76, p. 5514-5540.
Bielenstein, H. U., and Charlesworth, H.A.K., 1965, Precambrian sandstone sills near Jasper, Alberta: Bulletin of 1976. Steady-state flow of rocks: Reviews of Geophysics and Space Science, v. 14, p. 301-360.
Canadian Petroleum Geology, v. 13, p. 405-408. Carter, N. L., and Raleigh, C. B., 1969, Principal stress directions from plastic flow in crystals: Geological Society of
Bishop, J.F.W., 1953, A theoretical examination of the plastic deformation of crystals by glide: The Philosophical America Bulletin, v. 80, p. 1231-1264.
Magazine, v. 44, p. 51-64. Carter, N. L., Christie, J. M., and Griggs, D. T., 1964, Experimental deformation and recrystallization of quartz: Journal of
Blake, D. B., and Roy, C. J., 1949, Unusual stylolites: American Journal of Science, v. 247, p. 779-790. Geology, v. 72, p. 687-733.
Blake, M. C., Jr., Irwin, W. P., and Coleman, R, C., 1967, Upside-down metamorphic zonation, blueschist fades, along a Casas, J. M., and Sabat, F., 1987, An example of three-dimensional analysis of thrust-related tectonites: Journal of
regional thrust in California and Oregon: U.S. Geological Survey Professional Paper 575-C, p. C1-C9. Structural Geology, v. 9, p. 647-657.
Blenkinsop, T. G., and Rutter, E. H., 1986, Cataclastic deformation of quartzite in the Moine thrust zone: Journal of Casey, M., Dietrich, D., and Ramsay, J. G., 1983, Methods for determining history for chocolate tablet boudinage with
Structural Geology, v. 8, p. 669-681. fibrous crystals: Tectonophysics, v. 92, p. 211-239.
Bogacz, W., 1981, The role of cleavage in the shear folding and faulting processes in the vicinity of the Boguszowice thrust Chamberlin, T. G , 1888, Rock scorings of the great ice invasions: U.S. Geological Survey Seventh Annual Report,
(Upper Silesian Coal Basin): Bulletin de I'Academie Polonaise des Sciences, Serie des Sciences de la Terre, v. 29, p. 147-248.
p. 251-259. Cherry, J. T., Larson, D. B., and Rapp, E. G., 1968, A unique description of the failure of a brittle material: International
Bohm, A., 1883, Uber Gesteine des Wechsels: Tschermaks Mineralogische und Petrographische Mitteilungen, new series, Journal of Rock Mechanics and Mining Science, v. 5, p. 455-463.
v. 5, p. 197-214. Chester, F. M., and Logan, J. M., 1986, Implications for mechanical properties of brittle faults from observations of the
Bonham, L. C., 1957, Structural petrology of the Pico anticline, Los Angeles County, California: Journal of Sedimentary Punchbowl fault zone, California: Pure and Applied Geophysics, v. 124, p. 79-106.
Petrology, v, 27, p. 251-264. Chester, F. M., Friedman, M., and Logan, J. M., 1985, Foliated cataclasites: Tectonophysics, v. 111, p. 139-146.
Bonilla, M. G., Alt, J. N., and Hodgen, L. D., 1978, Trenches across the 1906 trace of the San Andreas fault in northern Chilingarian, I.G.V., and Wolf, K. H,, 1975, Compaction of coarse-grained sediments, I. Developments in
San Mateo County, California: U.S. Geological Survey Journal of Research, p. 347-358. sedimentology—18A: New York, Elsevier, 552 p.
Borg, I. Y., and Handin, J., 1966, Experimental deformation of crystalline rocks: Tectonophysics, v. 3, p. 251-367. Choquette, P. W., and James, N. P., 1987, Diagenesis in limestones—3. The deep burial environment: Geoscience
Borg, I. Y., and Heard, H. C., 1970, Experimental deformation of plagioclases, in Paulitsch, P., ed., Experimental and Canada, v. 14, p. 3-35.
natural rock deformation: New York, Springer-Verlag, p. 375-403. Choukroune, P., 1969, Un exemple d'analyse microtectonique d'une série calcaire affectée de plis isopaques ("con-
Borg, I. Y„ and Maxwell, J. C., 1956, Interpretation of fabrics of experimentally deformed sands: American Journal of centriques"): Tectonophysics, v. 7, p, 57-70.
Science, v. 254, p. 71-81. 1971, Contribution à l'étude des mechanismes de la déformtion avec schistosité grâce aux cristallisations synciné-
Borg, 1. Y., Friedman, M., Handin, J., and Higgs, D. V., 1960, Experimental deformation of St. Peter Sand: A study of matiques dan les "zones abriteés" ("pressure shadows"): Société Géologique de France Bulletin, v. 13, p. 257-271.
cataclastic flow, in Griggs, D., and Handin, J., eds., Rock deformation (A symposium): Geological Society of Choukroune, P., and Séguret, M., 1968, Exemple de relations entre joints de cisaillement, fentes de tension, plis et
America Memoir 79, p. 133-191. schistosité: Revue Géography Physique et de Géologie Dynamique, v. 10, p. 239-246.
Borradaile, G. J., 1977, On cleavage and strain: Results of a study in West Germany using technically deformed sand Christie, J. M., 1958, Dynamic interpretation of the fabric of dolomite from the Moine thrust zone in northwest Scotland:
dykes: Geological Society of London Journal, v. 133, p. 146-164. American Journal of Science, v. 256, p. 159-170.
1981, Particulate flow of rock and the formation of cleavage: Tectonophysics, v. 72, p. 305-321. Christie, J. M., and Ardell, A. J., 1974, Substructures of deformation lamellae in quartz: Geology, v. 2, p. 405-408.
Borradaile, G. J., Bayly, M. B., and Powell, C. McA., eds., 1982, Atlas of deformationa! and metamorphic rock fabrics: Christie, J. M., and Ord, A,, 1980, Flow stresses from microstructures of mylonites: Examples and current assessment:
Berlin, Springer-Verlag, 551 p. Journal of Geophysical Research, v. 85, p. 6253-6262.
Bostick, N. H., 1974, Phytoclasts as indicators of thermal metamorphism, Franciscan assemblage and Great Valley Christie, J. M., Griggs, D. T., and Carter, N. L., 1964, Experimental evidence of basal slip in quartz: Journal of Geology,
Sequence (upper Mesozoic), California: Geological Society of America Special Paper 153, p. 1-17. v. 72, p. 734-756.
Bott, M.H.P., 1959, The mechanics of oblique slip faulting: Geological Magazine, v. 46, p. 109-117. Clark, B. R., 1970, Mechanical formation of preferred orientation in clays: American Journal of Science, v. 269,
Boullier, A. M., and Gueguen, Y., 1975, SP-mylonites: Origin of some mylonites by superplastic flow: Contributions to p. 250-266.
Mineralogy and Petrology, v. 50, p. 93-104. Clendenen, W. S., Kligfield, R., Hirt, A. M., and Lowrie, W., 1988, Strain studies of cleavage development of the
Brace, W. F., 1960, An extension of the Griffith theory of fracture of rocks: Journal of Geophysical Research, v. 65, Chelmsford Formation, Sudbury Basin, Ontario: Tectonophysics, v. 145, p. 191-211.
p. 3477-3480. Clifton, H. E., 1965, Tectonic polish of pebbles: Journal of Sedimentary Petrology, v. 35, p. 867-873.
1961, Dependence of fracture strength of rocks on grain size: Pennsylvania State University Mineral Industries Cloos, E., 1932, Feather joints as indicators of the direction of movements on faults, thrusts, joints and magmatic contacts:
Experiment Station Bulletin, v. 76, p. 99-103. Proceedings of the National Academy of Sciences of the United States of America, v. 18, p. 387-395.
1964, Brittle fracture of rocks, in Judd, W. R., ed., State of stress in the Earth's crust: New York, American Elsevier 1946, Lineation, a critical review and annotated bibliography; Reprinted with supplement, 1953,1957: Geological
Publishing Company, Inc., p. 111-174. Society of America Memoir 18,122 p.
Brace, W. F., and Bombolakis, E. G., 1963, A note on brittle crack growth in compression: Journal of Geophysical 1947a, Boudinage: American Geophysical Union Transactions, v. 28, p. 626-632.
Research, v. 68, p. 3709-3713. 1947b, Oolite deformation in the South Mountain fold, Maryland: Geological Society of America Bulletin, v. 58,
Brace, W. F., and Byerlee, J. D., 1966, Stick-slip as a mechanism for earthquakes: Science, v. 153, p. 990-992. p. 843-918.
Brace, W. F., Paulding, B. W., Jr., and Scholz, C., 1966, Dilatancy in the fracture of the crystalline rocks: Journal of 1955, Experimental analysis of fracture patterns: Geological Society of America Bulletin, v. 66, p. 241-256.
Geophysical Research, v. 71, p. 3939-3953. Cloos, H., 1922, Über Ausbau und Anwendung der granit-tektonischen Methode: Preussische Geologische Landesanstalt
Bradley, W. C., 1963, Large-scale exfoliation in massive sandstones of the Colorado plateau: Geological Society of Abhandlungen, v. 89, p. 1-18.
America Bulletin, v. 74, p. 519-528. 1928, Experimente zur Inneren Tektonik: Centralblatt fur Mineralogie, Geologie und Paläontologie Abteilung B,
Bretz, J. H., 1940, Solution cavities in the Joliet Limestone of northeastern Illinois: Journal of Geology, v. 48, p. 337-384. Geologie und Paläontologie, p. 609-621.
1950, Origin of the filled sink-structures and circle deposits of Missouri: Geological Society of America Bulletin, 1930, Zur Experimentellen Tektonik: Die Naturwissenschaften, Jhq. 18, Hft. 34, p. 741-747.
v. 61, p. 789-834. Coble, R. L., 1963, Model for boundary diffusion controlled creep in polycrystalline materials: Journal of Applied Physics,
Brewer, R., 1964, Fabric and mineral analysis of soils: New York, John Wiley & Sons, Inc., 470 p. v. 34, p. 1679-1682.
Brock, W. G., and Engelder, T., 1977, Deformation associated with the movement of the Muddy Mountain overthrust in Collins, D. A., and de Paor, D. G., 1986, A determination of bulk rotational deformation resulting from displacements in
the Buffington window, southeastern Nevada: Geological Society of America Bulletin, v. 88, p. 1667-1677. discrete shear zones in the Hercynian fold belt of South Ireland: Journal of Structural Geology, v. 8, p. 101-109.
Brodzikowski, K., Gotowala, R., Haluszczak, A., Krzyszkowski, D., and Van Loon, A. J., 1987, Soft-sediment deforma- Collins, K., and McGowan, A., 1974, The form and function of microfabric features in a variety of natural soils:
tions from glaciodeltaic glaciolacustrine, and fluviolacustrine sediments in the Kleszczdw Graben, in Jones, M. E., Geotechnique, v. 24, p. 223-254.
and Preston, R.M.F., eds., Deformation of sediments and sedimentary rocks: Geological Society Special Publica- Compton, R. R., 1966, Analyses of Pliocene-Pleistocene deformation and stresses in northern Santa Lucia Range,
tion No. 29: Oxford, England, Blackwell Scientific Publications, p. 255-265. California: Geological Society of America Bulletin, v. 77, p. 1361-1380.
Brown, W. L., and Macaudiere, J., 1984, Microfracturing in relation to atomic structure of plagioclase from a deformed Conel, J. E., 1962, Studies of the development of fabrics in some naturally deformed limestones [Ph.D. dissert.}: California
meta-anorthosite: Journal of Structural Geology, v. 6, p. 579-586. Institute of Technology, Pasadena, California, 257 p.
LOW-TEMPERATURE DEFORMATION MECHANISMS 1355

Conrad, R. E., II, and Friedman, M., 1976, Microscopic feather fractures in the faulting process: Tectonophysics, v. 33, U.S.A.: Journal of Structural Geology, v. 7, p. 459-476.
p. 187-198. 1987, Joints and shear fractures in rock, in Atkinson, B. K., ed., Fracture mechanics of rock: London, England,
Conybeare, C.E.B., 1949, Stylolites in Precambrian quartzite: Journal of Geology, v. 57, p. 83-85. Academic Press, p. 27-69.
Conybeare, W. D., and Phillips, W., 1822, Outlines of the geology of England and Wales: London, England, William Engelder, J. T., and Engelder, R., 1977, Fossil distortion and décollement tectonics of the Appalachian Plateau: Geology,
Phillips, 470 p. v. 5, p. 457-460.
Coogan, A. H., and Manus, R. W., 1975, Compaction and diagenesis of carbonate sands, in Chilingarian, G. V., and Wolf, Engelder, J. T., and Geiser, P., 1979, The relationship between pencil cleavage and lateral shortening within the Devonian
K. H., eds., Compaction of coarse-grained sediments, I. Developments in sedimentology—18A: New York, section of the Appalachian Plateau, New York: Geology, v. 7, p. 460-464.
Elsevier, p. 79-166. 1980, On the use of regional joint sets as trajectories of paleostress fields during the development of the Appala-
Corbett, K., Friedman, M., and Spang, J., 1987, Fracture development and mechanical stratigraphy of Austin Chalk, chian Plateau, New York: Journal of Geophysical Research, v. 85, p. 6319-6341.
Texas: American Association of Petroleum Geologists Bulletin, v. 71, p. 17-28. Engelder, J. T., and Marshak, S., 1985, Disjunctive cleavage formed at shallow depths in sedimentary rocks: Journal of
Cottrell, A. H., 1964, The mechanical properties of matter: New York, John Wiley and Sons, Inc., 430 p. Structural Geology, v. 7, p. 327-343.
Coulomb, C. A., 1776, Sur une application des règles de maximis et minimis a quelques problèmes de statique relatifs à Engelder, J. T., and Oertel, G., 1985, Correlation between abnormal pore pressure and tectonic jointing in the Devonian
l'architecture: Academic des Sciences Mémoires de Math et de Physique, v. 7, p. 343-382. Catskill Delta: Geology, v. 13, p. 863-866.
Cowan, D. S., 1985, Structural styles in Mesozoic and Cenozoic melanges in the western Cordillera of North America: Engelder, J. T., and Scholz, C. H„ 1976, The role of asperity indentation and ploughing in rock friction—II: International
Geological Society of America Bulletin, v. 96, p. 451-462. Journal of Rock Mechanics and Mining Science, v. 13, p. 155-163.
Cox, M. A., and Whitford-Stark, J. L., 1987, Stylolites in the Caballos Novaculite, west Texas: Geology, v. 15, 1981, Fluid flow along very smooth joints at effective pressures up to 200 Megapascals: American Geophysical
p. 439-442. ' Union Monograph 24, p. 147-152.
Cox, S. F., 1987, Antitaxial crack-seal vein microstructures and their relationship to displacement paths: Journal of Engelder, J. T., Logan, J. M., and Handin, J., 1975, The sliding characteristics of sandstone on quartz fault gouge: Pure
Structural Geology, v. 9, p. 779-787. and Applied Geophysics, v. 123, p. 69-86.
Cox, S. F., and Etheridge, M. A., 1983, Crack-seal fibre growth mechanisms and their significance in the development of Epstein, A. G., Epstein, J. B., and Harris, L. D., 1977, Conodont color alteration—An index to organic metamorphism:
oriented layer silicate microstructures: Tectonophysics, v. 92, p. 147-170. U.S. Geological Survey Professional Paper 995, 27 p.
Crampton, C. B., 1958, Structural petrology of Cambro-Ordovician limestones of northwest Highlands of Scotland: Etchecopar, A., Vasseur, G., and Daignieres, M., 1981, An inverse problem in microtectonics for the determination of
American Journal of Science, v. 256, p. 145-158. stress tensors from fault striation analysis: Journal of Structural Geology, v. 3, p. 51 -65.
Crook, K. A., 1964, Cleavage in weakly deformed mudstones: American Journal of Science, v. 262, p. 523-531. Etheridge, M. A., 1983, Differential stress magnitudes during regional deformation and metamorphism: Upper bound
Crowell, J. G , 1974, Origin of Late Cenozoic basins in southern California: Society of Economic Paleontologists and imposed by tensile fracturing: Geology, v. 11, p. 231-234.
Mineralogists Special Publication No. 22, p. 190-204. 1984, Reply to comment on "Differential stress magnitudes during regional deformation and metamorphism:
Currie, J. B,, 1969, Written comments on "Structural analysis of features on natural and artificial faults" by D. K. Norris Upper bound imposed by tensile fracturing": Geology, v. 12, p. 56-57.
and K. Barron: Canada Geological Survey, Paper 68-52, p. 168-172. Etheridge, M. A., and Oertel, G., 1979, Strain measurements from phyllostlicate preferred orientation—A precautionary
Curtis, C. D., Lipshie, S. R., Oertel, G., and Pearson, M. J., 1980, Clay orientation in some Upper Carboniferous note: Tectonophysics, v. 60, p. 107-120.
mudrocks, its relationship to quartz content and some inferences about fissility, porosity, and compactional history: Etheridge, M. A., and Wilkie, J. C., 1979, Grain size reduction, grain boundary sliding and the flow strength of mylonites:
Sedimentology, v. 27, p. 333-339. Tectonophysics, v, 58, p. 159-178.
Dale, T. N., and others, 1914, Slate in the United States: U.S. Geological Survey Bulletin 586, 220 p. 1981, An assessment of dynamically recrystallized grain size as a palaeopiezometer in quartz-bearing mylonite
Daubrée, A., 1879, Études synthétiques de géologie expérimentale: Paris, Dunod, 828 p. zones: Tectonophysics, v. 78, p. 475-508.
Davies, W., 1982, Fine structure of slate, in Borradatte, G. J., Bayly, M. B., and Powell, C. McA., eds., Atlas of Evans, K. F., and Engelder, J. T., 1986, A study of stress in Devonian shale: Part 1—3D stress mapping using a wireline
deformational and metamorphic rock fabrics: New York, Springer-Verlag, p. 64-65. microfrac system: Annual Technical Conference and Exhibition of the Society of Petroleum Engineers, 61st, New
de Boer, R. B., 1977a, On the thermodynamics of pressure solution—interaction between chemical and mechanical forces: Orleans, 1986, SPE 15609,12 p.
Geochimica et Cosmochimica Acta, v. 41, p. 249-256. Fairbairn, H. W., 1941, Deformation lamellae in quartz from the Ajibik Formation, Michigan: Geological Society of
- — — 1977b, Pressure solution: Theory and experiments: Tectonophysics, v. 39, p. 287-301. America Bulletin, v. 52, p. 1265-1278.
de Boer.R. B., Nagtegaal, P.J.C., and Duyvis, E. M., 1977, Pressure solution experiments on quartz sand: Geochimica et Fairbairn, H. W., and Hawkes, H. E., Jr., 1941, Dolomite orientation in deformed rocks: American Journal of Science,
Cosmochimica Acta, v. 41, p. 257-264. v. 239, p. 617-632.
Deelman, J. G , 1975, "Pressure solution" or indentation?: Geology, v. 3, p. 23-24. Farmin, R., 1941, Host-rock inflation by veins and dikes at Grass Valley, California: Economic Geology, v. 36,
DeGraff, M. J., and Aydin, A., 1987, Surface morphology of columnar joints and its significance to mechanics and p. 143-174.
direction of joint growth: Geological Society of America Bulletin, v. 99, p. 605-617. Fellows, R. E., 1943, Recrystallization and flowage in Appalachian quartzite: Geological Society of America Bulletin,
Dennis, J. G., 1967, International tectonic dictionary: American Association of Petroleum Geologists Memoir 7, 196 p. v. 54, p. 1399-1432.
1987, Structural geology, an introduction: Dubuque, Iowa, Wm. C. Brown Publishers, 448 p. Fernandez, A., 1987, Preferred orientation developed by rigid markers in two-dimensional simple shear strain: A theoreti-
de Sitter, L. U., 1956, Structural geology: New York, McGraw-Hill Book Company, Inc., 552 p. cal and experimental study: Tectonophysics, v. 136, p. 151-158.
Dietrich, D., and Song, H., 1984, Calcite fabrics in a natural shear environment, the Helvetic Nappes of western Flehmig, W., and Langheinrich, G., 1974, Relation between tectonic deformation and illite crystallinity: Neues Jahrbuch
Switzerland: Journal of Structural Geology, v. 6, p. 19-32. für Geologie und Paläontologie Abhandlungen, v. 146, p. 325-346.
Dietz, R. S., 1959, Shatter cones in cryptoexplosion structures (meteorite impact?): Journal of Geology, v. 67, p. 496-505. Fletcher, R. C., and Hofmann, A. W., 1974, Simple models of diffusion and combined diffusion-infiltration metasomatism:
Diller, J. S,, 1890, Sandstone dikes: Geological Society of America Bulletin, v. 1, p. 411-442. Carnegie Institution of Washington Publication 634, p. 243-259.
Dionne, J.-G, and Shilts, W. W., 1974, A Pleistocene clastic dike, Upper Chaudiere Valley, Quebec: Canadian Journal of Fletcher, R. C., and Pollard, D. D., 1981, Anticrack model for pressure solution surfaces: Geology, v. 9, p. 419-424.
Earth Sciences, v. 11, p. 1594-1605. Fleuty, M. J., 1975, Slickensides and slickenlines: Geological Magazine, v. 112, p. 319-322.
Dokka, R. K., 1986, Patterns and modes of early Miocene crustal extensions, central Mojave Desert, California: Geologi- Folk, R. L., 1965, Some aspects of recrystallization in ancient limestones, in Pray, L. G , and Murray, R. C., eds.,
cal Society of America Special Paper 208, p. 75-95. Dolomitization and limestone diagenesis, a symposium: Society of Economic Paleontologists and Mineralogists
Donath, F. A., 1970, Some information squeezed out of rock: American Scientist, v. 58, p. 54-72. Special Publication 13, p. 14-48.
Donaih, F. A., and Fruth, L. S., Jr., 1971, Dependence of strain-rate effects on deformation mechanisms and rock types: 1974, The natural history of crystalline calcium carbonate: Effect of magnesium content and salinity: Journal of
Journal of Geology, v. 79, p. 347-371. Sedimentary Petrology, v. 44, p. 40-53.
Donath, F. A., and Wood, D. S., 1976, Experimental evaluation of the deformation path concept: Royal Society of Foster, M. E., and Hudleston, P. J., 1986, "Fracture cleavage" in the Duluth Complex, northeastern Minnesota: Geologi-
London Philosophical Transactions, ser. A, v. 283, p. 187-201. cal Society of America Bulletin, v. 97, p. 85-96.
Donaih. F. A., Fail!, R. T., and Tobin, D. G., 1971, Deformation mode fields in experimentally deformed rock: Geological Foster, R. H., and De, P. K., 1971, Optical and electron microscopic investigation of shear induced structures in lightly
Society of America Bulletin, v. 82, p. 1441-1462. consolidated (soft) and heavily consolidated (hard) kaolinite: Clays and Clay Minerals, v. 19, p. 31-47.
Dow, W. G., 1978, Petroleum source beds on continental slope and rises: American Association of Petroleum Geologists Fraser, H. J., 1935, Experimental study of porosity and permeability of clastic sediments: Journal of Geology, v. 43,
Bulletin, v. 62, p. 1584-1606. p. 910-1010.
Droxler, A., arid Schaer, J.-P., 1979, Déformation cataclastique plastique lors du plissement sous faible couverture, de Friedman, M., 1963, Petrofabric analysis of experimentally deformed calcite-cemented sandstones: Journal of Geology,
strates calcaires: Eclogae Geologicae Helvetiae, v. 72, p. 551-570. v. 71, p. 12-37.
Dula, W. F., Jr., 1981, Correlation between deformation lamellae, microfractures, macrofractures, and in situ stress 1964, Petrofabric techniques for determination of principal stress direction in rocks, in Judd, W. R., ed., State of
measurements, White River Uplift, Colorado: Geological Society of America Bulletin, Part I, v. 92, p. 37-46. stress in the Earth's crust: New York, American Elsevier Publishing Company, Inc., p. 451-550.
Dunham, R. J., 1962, Classification of carbonate rocks according to ^positional texture, in Ham, W. E., ed., Classifica- 1967, Description of rocks and rock masses with a view toward their physical and mechanical behavior—A
tion of carbonate rocks: American Association of Petroleum Geologists Memoir 1, p. 108-121. general report: International Congress on Rock Mechanics, 1st, Proceedings, v. 3, p. 182-197.
Dunn, D. £., LaFountain, L. J., and Jackson, R. E., 1973, Porosity dependence and mechanism of brittle fracture in 1969, Structural analysis of fractures in cores from Saticoy Field, Ventura County, California: American Associa-
sandstone: Journal of Geophysical Research, v. 78, p. 2403-2417. tion of Petroleum Geologists Bulletin, v. 53, p. 367-389.
Dunnington, H. V., 1967, Aspects of diagenesis and shape change in stylolitic limestone reservoirs: World Petroleum 1972, Residual elastic strain in rocks: Tectonophysics, v. 15, p. 297-330.
Congress, 7th, Mexico, Proceedings, v. 2, p. 339-352. Friedman, M., and Higgs, N. G., 1981, Calcite fabrics in experimental shear zones: American Geophysical Union
Durney, D. W., 1972, Solution-transfer, an important geological deformation mechanism: Nature, v. 235, p. 315-317. Monograph 24, p. 11-27.
1974, The influence of stress concentrations on the lateral propagation of pressure solution zones and surfaces Friedman, M., and Logan, J. M., 1970, Microscopic feather fractures: Geological Society of America Bulletin, v. 81,
[abs.]: Geological Society of Australia, Tectonics and Structural Newsletter, no. 3» p. 19, p. 3417-3420.
1978, Early theories and hypotheses on pressure-soiution-redeposition: Geology, v. 6, p. 369-372. 1973, Liiders' bands in experimentally deformed sandstone and limestone: Geological Society of America Bulletin.
Durney, D. W„ and Ramsay, J, G., 1973, Incremental strains measured by syntectoniccrystal growths, in DeJong, K. A., v. 84, p. 1465-1476.
and Scholten, R., eds., Gravity and tectonics: New York, Wiley-lnterscience, p, 67-96. Friedman, M., and Sowers, G. M., 1970, Petrofabrics: A critical review: Canadian Journal of Earth Sciences, v. 7,
Dzulynski, S., and Kotlarczyk, J., 1965, Tectoglyphs on stickensided surfaces: Bulletin de l'Académie Polonaise des p. 447-497.
Sciences, Serie des sciences géologie et géographie, v. 13, p. 149-154. Friedman, M., Logan, J. M., and Rigert, J. A., 1974, Glass-indurated quartz gouge in sliding-friction experiments on
Edmond, J. M., and Paterson, M. S., 1972, Volume changes during the deformation of rocks at high pressures: Interna- sandstone: Geological Society of America Bulletin, v. 85, p. 937-942.
tional Journal of Rock Mechanics and Mining Science, v. 9, p. 161-182. Frost, H. J., and Ashby, M. F., 1982, Deformation mechanism maps: The plasticity and creep of metals and ceramics:
Elliott, D., 1973, Diffusion flow laws in metamorphic rocks: Geological Society of America Bulletin, v. 84, p. 2645-2664. New York, Pergamon, 184 p.
1976, The energy balance and deformation mechanisms of thrust sheets: Royal Society of London Philosophical Fruth, L. S., Jr., Orme, G. R., and Donath, F. A., 1966, Experimental compaction effects in carbonate sediments: Journal
Transactions, ser. A, v. 283, p. 289-312. of Sedimentary Petrology, v. 36, p. 747-754.
Engelder, J. T., 1974a, Cataclasis and the generation of fault gouge: Geological Society of America Bulletin, v. 85, Fry, N., 1979, Random point distributions and strain measurement in rocks: Tectonophysics, v. 60, p. 89-105.
p. 1515-1522. Gaither, A., 1953, A study of porosity and grain relationships in experimental sands: Journal of Sedimentary Petrology,
1974b, Microscopic wear grooves on slickensides: Indicators of paleoseismidty: Journal of Geophysical Research, v. 23, p. 180-195.
v. 79, p, 4387-4392. Gallagher, J. J., Jr., Friedman, M., Handin, J., and Sowers, G. M., 1974, Experimental studies relating to microfracture in
1979, Mechanisms for strain within the Upper Devonian clastic sequence of the Appalachian Plateau, western sandstone: Tectonophysics, v. 21, p, 203-248,
New York: American Journal of Science, v. 279, p. 527-542. Gamond, J. F., 1983, Displacement features associated with fault zones: A comparison between observed examples and
1982, Is there a genetic relationship between selected regional joints and contemporary stress within the lithosphere experimental models: Journal of Structural Geology, v. 5, p. 33-45.
of North America?: Tectonics, v. 1, p. 161-177. 1987, Bridge structures as sense of displacement criteria in brittle fault zones: Journal of Structural Geology, v. 9,
1985, Loading paths to joint propagation during a tectonic cycle: An example from the Appalachian Plateau, p. 609-620.
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

1356 R. H. GROSHONG, JR.

Garnett, J. A., 1974, A mechanism for the development of en-echelon gashes in kink zones: Tectonophysics, v. 23, Groshong, R. H., Jr., Pfiffner, O. A., and Pringle, L. R., 1984a, Strain partitioning in the Helvetic thrust belt of eastern
p. 129-138. Switzerland from the leading edge to the internal zone: Journal of Structural Geology, v. 6, p. 5-18.
Garrison, R. E., and Kennedy, W. J., 1977, Origin of solution seams and flaser structure in Upper Cretaceous chalks of Groshong, R. H., Jr., Teufel, L. W., and Gasteiger, C., 1984b, Precision and accuracy of the calcite strain-gage technique:
southern England: Sedimentary Geology, v. 19, p. 107-137. Geological Society of America Bulletin, v. 95, p. 357-363.
Gaviglio, P., 1986, Crack-seal mechanism in a limestone: A factor of deformation in strike-slip faulting: Tectonophysics, Grubenmann, U., and Niggli, P., 1924, Die Gesteinmetamorphose (Volume 1): Berlin, Borntraeger, 539 p.
v. 131, p. 247-255. Guiraud, M., and S6guret, M., 1987, Soft-sediment microfaulting related to compaction within the fluvio-deltaic infill of
Gay, N. C., and Jaeger, J. C., 1975a, Cataclastic deformation of geological materials in matrices of differing composition: the Soria strike-slip basin (northern Spain), in Jones, M. E., and Preston, R.M.F., eds., Deformation of sediments
I. Pebbles and conglomerates: Tectonophysics, v. 27, p. 303-322. and sedimentary rocks; Geological Society Special Publication No. 29: Oxford, England, Blackwell Scientific
1975b, Cataclastic deformation of geological materials in matrices of differing composition: II. Boudinage: Tec- Publications, p. 123-136.
tonophysics, v. 27, p. 323-331. Guthrie, J. M., Houseknecht, D. W., and Johns, W. D., 1986, Relationships among vitrinite reflectance, illite crystallinity,
Gay, N. C., and Ortlepp, W, D., 1979, Anatomy of a mining-induced fault zone: Geological Society of America Bulletin, and organic geochemistry in Carboniferous strata, Ouachita Mountains, Oklahoma and Arkansas: American
Part 1, v. 90, p. 47-58, Association of Petroleum Geologists Bulletin, v. 70, p. 26-33.
Gay, N. G , and Weiss, L. E., 1974, The relationship between principal stress directions and the geometry of kinks in Guzzetta, G., 1984, Kinematics of stylolite formation and physics of the pressure-solution process: Tectonophysics, v. 101,
foliated rocks: Tectonophysics, v. 21, p. 287-300. p. 383-394.
Geiser, P. A., 1974, Cleavage in some sedimentary rocks of the central Valley and Ridge province, Maryland: Geological Hadizadeh, J., and Rutter, E. H., 1983, The low temperature brittle-ductile transition in quartzite and the occurrence of
Society of America Bulletin, v. 85, p. 1399-1412. cataclastic flow in nature: Geologische Rundschau, v. 7, p. 493-509.
1975, Slaty cleavage and the dewatering hypothesis—An examination of some critical evidence: Geology, v. 3, Hancock, P. L., 1965, Axial-trace-fractures and deformed concretionary rods in south Pembrokeshire: Geological Magazine,
p. 717-720. v. 102, p. 143-163.
Geiser, P, A., and Engelder, T., 1983, The distribution of layer parallel shortening fabrics in the Appalachian foreland of 1972, The analysis of en-echelon veins: Geological Magazine, v. 109, p. 269-276.
New York and Pennsylvania: Evidence for two non-coaxial phases of the Alleghanian orogeny: Geological 1985, Brittle microtectonics: Principles and practice: Journal of Structural Geology, v. 7, p. 437-457.
Society of America Memoir 158, p. 161-175. Hancock, P. L., and Barka, A. A., 1987, Kinematic indicators on active normal faults in western Turkey: Journal of
Geiser, P. A., and Sansone, S., 1981, Joints, microfractures, and the formation of solution cleavage in limestone: Geology, Structural Geology, v. 9, p. 573-584.
v. 9, p. 280-285. Handin, J., 1966, Strength and ductility: Geological Society of America Memoir 97, p. 223-289.
Ghisetti, F., 1987, Mechanisms of thrust faulting in the Gran Sasso chain, central Apennines, Italy: Journal of Structural 1969, On the Coulomb-Mohr failure criterion: Journal of Geophysical Research, v. 74, p. 5343-5348.
Geology, v. 9, p. 955-967. Handin, J., and Fairbairn, H. W., 1955, Experimental deformation of Hasmark dolomite: Geological Society of America
Gilbert, G. K., 1928, Studies in basin-range structure: U.S. Geological Survey Professional Paper 153,92 p. Bulletin, v. 66, p. 1257-1273.
Gipson, M., 196S, Application of the electron microscope to the study of particle orientation and fissility in shale: Journal Handin, J. W., and Griggs, D., 1951, Deformation of Yule Marble: Part II—Predicted fabric changes: Geological Society
of Sedimentary Petrology, v. 35, p. 408-414. of America Bulletin, v. 62, p. 863-886.
Glover, J. E., 1969, Observations on stylolites in Western Australian rocks: Royal Society of Western Australia Journal, Handin, J., and Hager, R. V., 1957, Experimental deformation of sedimentary rocks under confining pressure: Tests at
v. 52, p. 12-17. room temperature on dry samples: American Association of Petroleum Geologists Bulletin, v. 41, p. 1-50.
Golding, H. G., and Conolly, J. R., 1962, Stylolites in volcanic rocks: Journal of Sedimentary Petrology, v. 32, 1958, Experimental deformation of sedimentary rocks under confining pressure: Tests at high temperature:
p. 534-538. American Association of Petroleum Geologists Bulletin, v. 42, p. 2892-2934.
Gorlov, N. V., 1971, The mechanism of the opening of extension fractures (in the instance of pegmatite veins, north- Handin, J., Hager, R. V., Friedman, M., and Feather, J. N., 1963, Experimental deformation of sedimentary rocks under
western White Sea region): Geotectonics, no. 3, p. 148-152. confining pressure: Pore pressure tests: American Association of Petroleum Geologists Bulletin, v. 47, p. 717-755.
Graham, R. H., 1978, Quantitative deformation studies in the Permian rocks of the Alpes-Maritimes: Goguel Symposium, Handin, J., Heard, H. G , and Magouirk, J. N., 1967, Effects of the intermediate principal stress on the failure of limestone,
Bureau de Recherches Géologiques et Minières, Mémoire 91, p. 219-238. dolomite, and glass at different temperatures and strain rates: Journal of Geophysical Research, v. 72, p. 611-640.
Gramberg, J., 1965, The axial cleavage fracture 1. Axial cleavage fracturing, a significant process in mining and geology: Hanmer, S., 1981, Segregation bands in plagioclase: Non-dilational quartz veins formed by strain-enhanced diffusion:
Engineering Geology, v. 1, p. 31-72. Tectonophysics, v. 79, p. T53-T61.
Gratier, J. P., and Guiguet, R., 1986, Experimental pressure solution-deposition on quartz grains: The crucial effect of the Hanna, S. S., and Fry, N., 1979, A comparison of methods of strain determination in rocks from Southwest Dyfed
nature of the fluid: Journal of Structural Geology, v. 8, p, 845-856. (Pembrokeshire) and adjacent areas: Journal of Structural Geology, v. 1, p. 155-162.
Graton, L. C., and Fraser, H. J., 1935, Systematic packing of spheres—with particular relation to porosity and permeabil- Hansen, E. C., and Borg, I. Y., 1962, The dynamic significance of deformation lamellae in quartz of a calcite-cemented
ity: Journal of Geology, v. 43, p. 785-909. sandstone: American Journal of Science, v. 260, p. 321-336.
Gray, D. R., 1977a, Differentiation associated with discrete crenulation cleavages: Lithos, v. 10, p. 89-101. Harms, J. C., 1965, Sandstone dikes in relation to Laramide faults and stress distribution in the southern Front Range,
1977b, Morphologic classification of crenulation cleavage: Journal of Geology, v. 85, p. 229-235. Colorado: Geological Society of America Bulletin, v. 76, p. 981-1002.
—1978, Cleavages in deformed psammitic rocks from southeastern Australia: Their nature and origin: Geological Hawkins, P. J., 1978, Relationship between diagenesis, porosity reduction, and oil emplacement in late Carboniferous
Society of America Bulletin, v. 89, p. 577-590. sandstone reservoirs, Bothamsall Oilfield, E Midlands: Journal of Geological Society of London, v. 135, p. 7-24.
1979, Microstructure of crenulation cleavages: An indicator of cleavage origin: American Journal of Science, Heald, M. T., 1955, Stylolites in sandstones: Journal of Geology, v. 63, p. 101-114.
v. 279, p. 97-128. 1956, Cementation of Simpson and St. Peter sandstones in parts of Oklahoma, Arkansas, and Missouri: Journal of
1981, Compound tectonic fabrics in singly folded rocks from southwest Virginia, U.S.A.: Tectonophysics, v. 78, Geology, v. 64, p. 16-30.
p.229-248. Heard, H. C., 1960, Transition from brittle to ductile flow in Solenhofen Limestone as a function of temperature, confining
Gray, D. R„ and Durney, D. W., 1979, Investigations on the mechanical significance of crenulation cleavage: Tectono- pressure, and interstitial fluid pressure: Geological Society of America Memoir 79, p. 193-226.
physics, v. 58, p. 35-79. 1963, The effect of large changes in strain rate in the experimental deformation of the Yule marble: Journal of
Gray, J., and Boucot, A. J., 1975, Color changes in pollen and spores: A review: Geological Society of America Bulletin, Geology, v. 71, p. 162-195.
v. 86, p. 1019-1033. Heard, H. C., and Carter, N. L., 1968, Experimentally induced "natural" intragranular flow in quartz and quartzite:
Green, H, W., II, 1984, "Pressure solution" creep: Some causes and mechanisms: Journal of Geophysical Research, v. 89, American Journal of Science, v. 266, p. 1-42.
p. 4313-4318. Heard, H. C., Wenk, H.-R., and Barber, D. J., 1978, Experimental deformation of dolomite single crystals to 800 °C
Gresley, W. S„ 1895, The indentation of the Bunter pebbles: Geological Magazine, v. 32, p. 239. [abs.]: EOS (American Geophysical Union Transactions), v. 59, p. 249.
Gretener, P. E., 1977, On the character of thrust faults with particular reference to the basal tongues: Bulletin of Canadian Heim, A., 1878, Untersuchungen über den Mechanismus der Gebirgsbildung im Anschluss an die geologische Mono-
Petroleum Geology, v. 25, p. 110-122. graphie der Tödi-Wingällen-Gruppe: Basal, Switzerland, Benno Schwabe, 3 volumes (I, II, Atlas).
Griffith, A. A., 1921, The phenomena of rupture and flow in solids: Royal Society of London Philosophical Transactions, Helmstaedt, H., andGreggs, R.G., 1980, Stylolitic cleavage and cleavage refraction in lower Paleozoic rocks of the Great
ser. A, v. 221, p. 163-198. Valley, Maryland: Tectonophysics, v. 66, p. 99-114.
1924, The theory of rupture: International Congress for Applied Mechanics, 1st, Delft, Proceedings, p. 55-63. Helwig, J., 1970, Slump folds and early structures, northeastern Newfoundland Appalachians: Journal of Geology, v. 78,
Griggs, D. T., 1935, The strain ellipsoid as a theory of rupture: American Journal of Science, 5th ser., v. 30, p. 121-137. p. 172-187.
— — 1 9 3 6 , Deformation of rocks under high confining pressure I. Experiments at room temperature: Journal of Heron, S. D., Judd, J. B., and Johnson, H. S., 1971, Clastic dikes associated with soil horizons in the North and South
Geology, v. 44, p. 541-577. Carolina Coastal Plain: Geological Society of America Bulletin, v. 82, p. 1801-1810.
1939, Creep of rocks: Journal of Geology, v. 47, p. 225-251. Hertz, H., 1896, Hertz's Miscellaneous Papers: London, England, Macmillan, p. 146-162.
1940, Experimental flow of rocks under conditions favoring recrystallization: Geological Society of America Hibbard, J., and Karig, D. E., 1987, Sheath-like folds and progressive fold deformation in Tertiary sedimentary rocks of
Bulletin, v. 51, p. 1001-1022. the Shimanto accretionary complex, Japan: Journal of Structural Geology, v. 9, p. 845-857.
1942, Strength and plasticity: Geological Society of America Special Paper 36, p. 107-130. Higgins, M. W., 1971, Cataclastic rocks: U.S. Geological Survey Professional Paper 687,97 p.
Griggs, D. T., and Bell, J. F., 1938, Experiments bearing on the orientation of quartz in deformed rocks: Geological Higgs, D. V., and Handin, J., 1959, Experimental deformation of dolomite single crystals: Geological Society of America
Society of America Bulletin, v. 49, p. 1723-1746. Bulletin, v. 70, p. 245-278.
Griggs, D. T., and Blacic, J. D., 1965, Quartz: Anomalous weakness of synthetic crystals: Science, v. 147, p. 292-295. Hills, E. S., 1953, Outlines of structural geology (3rd ed.): London, England, Methuen, 182 p.
Griggs, D. T., and Handin, J., 1960, Observation on fractures and a hypothesis of earthquakes: Geological Society of 1963, Elements of structural geology: New York, John Wiley & Sons, Inc., 483 p.
America Memoir 79, p. 347-364. Hiraga, H., and Shimamoto, T., 1987, Textures of sheared halite and their implications for the seismogenic slip of deep
Griggs, D. T., and Miller, W. B., 1951, Deformation of Yule marble: Part I—Compression and extension experiments on faults: Tectonophysics, v. 144, p. 69-86.
dry Yule marble at 10,000 atmospheres confining pressure, room temperature: Geological Society of America Hobbs, B. E„ 1968, Recrystallization of single crystals of quartz: Tectonophysics, v. 6, p. 353-401.
Bulletin, v, 62, p, 853-862. Hobbs, B. E., Means, W. D., and Williams, P. F., 1976, An outline of structural geology: New York, John Wiley & Sons,
Griggs, D. T., Turner, F., Borg, I., and Sosoka, J., 1951, Deformation of Yule marble: Part IV: Effects at 150 °C: Inc., 571 p.
Geological Society of America Bulletin, v. 62, p. 1385-1406. Hobbs, B. E., Ord, A., and Teyssier, C., 1986, Earthquakes in the ductile regime?: Pure and Applied Geophysics, v. 124,
1953, Deformation of Yule marble: Part V: Effects at 300 °C: Geological Society of America Bulletin, v. 64, p. 309-336.
p. 1327-1342. Hodgson, R. A., 1961a, Classification of structures on joint surfaces: American Journal of Science, v. 259, p. 493-502.
Griggs, D. T., Paterson, M. S., Heard, H. C., and Turner, F. J., 1960a, Annealing recrystallization in calcite crystals and 1961b, Regional study of jointing in Comb Ridge-Navajo Mountain area, Arizona and Utah: American Associa-
aggregates: Geological Society of America Memoir 79, p. 21-37. tion of Petroleum Geologists Bulletin, v. 45, p. 1-38.
Griggs, D. T., Turner, F. J„ and Heard, H. C., 1960b, Deformation of rocks at 500 to 800 °C: Geological Society of Hoek, E., and Bieniawski, Z. T., 1965, Brittle fracture propagation in rock under compression: International Journal of
America Memoir 79, p. 39-104. Fracture Mechanics, v. 1, p. 137-155.
Groshong, R. H., Jr., 1965, Systematic joints across a syncline in the Valley and Ridge province [abs.]: Pennsylvania Holeywell, R. C., and Tullis, T. E., 1975, Mineral reorientation and slaty cleavage in the Martinsburg Formation, Lehigh
Academy of Science Proceedings, v. 39, p. 23. Gap, Pennsylvania: Geological Society of America Bulletin, v. 86, p. 1296-1304.
1972, Strain calculated from twinning in calcite: Geological Society of America Bulletin, v. 83, p. 2025-2038. Honea, E., and Johnson, A. M., 1976, A theory of concentric, kink, and sinusoidal folding and of monoclonal flexuring of
1974, Experimental test of least-squares strain gage calculation using twinned calcite: Geological Society of compressible, elastic multilayers. IV. Development of sinusoidal and kink folds in multilayers confined by rigid
America Bulletin, v. 85, p. 1855-1864. boundaries: Tectonophysics, v. 30, p. 197-239.
1975a, "Slip" cleavage caused by pressure solution in a buckle fold: Geology, v. 3, p. 411-413. Horsfield, W. T., 1980, Contemporaneous movement along crossing conjugate normal faults: Journal of Structural
1975b, Strain, fractures, and pressure solution in natural single-layer folds: Geological Society of America Bulletin, Geology, v. 2, p. 305-310.
v. 86, p. 1363-1376. Hortenbach, R„ 1977, Pressure solution processes in carbonates and their importance: Zeitschrift fur Geologie Wissen-
1976, Strain and pressure solution in the Martinsburg Slate, Delaware Water Gap, New Jersey: American Journal schaften, v. 5, p. 617-621.
of Science, v. 276, p. 1131-1146. House, W. M., and Gray, D. R., 1982, Cataclasites along the Saltville thrust, U.S.A., and their implications for thrust-sheet
LOW-TEMPERATURE DEFORMATION MECHANISMS 1357

emplacement: Journal of Structural Geology, v, 4, p. 257-269, Lennox, P. G., 1987, Conjugate oblique planar fabrics in rock analogues—by kinking or not by kinking?: Geological
Houseknecht, D. W., 1984, Influence of grain size and temperature on intergranular pressure solution, quartz cementation, Society of Australia, International Conference on Deformation of Crustal Rocks, Abstracts, no. 19, p. 89-90,
and porosity in a quartzose sandstone: Journal of Sedimentary Petrology, v. 54, p. 348-361, Lindström, M., 1962, A structural study of the southern end of the French Jura: Geological Magazine, v. 99, p. 193-207.
1987, Assessing the relative importance of compaction processes and cementation to reduction of porosity in Lisle, R. J., 1977, Clastic grain shape and orientation in relation to cleavage from the Aberystwyth grits, Wales:
sandstones: American Association of Petroleum Geologists Bulletin, v. 71, p. 633-642. Tectonophysics, v. 39, p. 381-395.
Howell, B. F., Jr., 1986, History of ideas on the cause of earthquakes: EOS (American Geophysical Union Transactions), Lockwood, J. P., and Moore, J. G., 1979, Regional deformation of the Sierra Nevada, California, on conjugate microfault
v. 67, p. 1323-1326. sets: Journal of Geophysical Research, v. 84, p. 6041-6049.
Hubbert, M. K., 1951, Mechanical basis for certain familiar geologic structures: Geological Society of America Bulletin, Logan, B. W., and Semeniuk, V., 1976, Dynamic metamorphism; processes and products in Devonian carbonate rocks,
v. 62, p. 355-372. Canning Basin, Western Australia: Geological Society of Australia Special Publication No. 6,138 p.
Hugman, R.H.H., III, and Friedman, M., 1979, Effects of texture and composition on mechanical behavior of experimen- Logan, J. M., 1979, Brittle phenomena: Reviews of Geophysics and Space Physics, v. 17, p. 1121-1132.
tally deformed carbonate rocks: American Association of Petroleum Geologists Bulletin, v. 63, p. 1478-1489. Logan, J. M., Higgs, N. G., and Friedman, M., 1981, Laboratory studies on natural gouge from the U.S. Geological Survey
Iddings, J. P., 1886, The columnar structure in the igneous rocks on Orange Mountain, New Jersey: American Journal of Dry Lake Valley No. 1 well, San Andreas fault zone: American Geophysical Union Monograph 24, p. 121-134.
Science, 3rd ser., v. 31, p. 321-331. Logan, J. M., Iwasaki, T„ Friedman, M., and Kling, S. A., 1972, Experimental investigations of sliding friction in
Ingerson, E,, and Tuttle, O. F., 1945, Relations of lamellae and crystallography of quartz and fabric directions in some multilithologic specimens, in Pincus, H., ed., Geological factors in rapid excavation: Geological Society of America
deformed rocks: American Geophysical Union Transactions, v. 26, p. 95-105. Engineering Geology Case History No. 9, p. 55-67.
Ingram, R. L., 1953, Fissility of mudrocks: Geological Society of America Bulletin, v. 64, p. 869-878. Logan, J. M., and Rauenzahn, K. A., 1987, Frictional dependence of gouge mixtures of quartz and montmorillonite on
Jackson, R. E., and Dunn, D. E., 1974, Experimental sliding friction and cataclasis of foliated rocks: International Journal velocity, composition and fabric: Tectonophysics, v. 144, p. 87-108.
of Rock Mechanics and Mining Science, v. 11, p. 235-249. Lohest, M., 1909, De l'origine des veines et des géodes des terrains primaires de Belgique: Annales Société Géologique de
Jaeger, J. C., 1959, The frictional properties of joints in rocks: Pure and Applied Geophysics, v. 43, p. 148-158. Belgique, v. 36, B, p. 275-282.
Jaeger, J. C„ and Cook, N.G.W., 1979, Fundamentals of rock mechanics (3rd edition): London, England, Chapman and Lowry, W. D., 1956, Factors in loss of porosity by quartzose sandstones of Virginia: American Association of Petroleum
Hall, 593 p. Geologists Bulletin, v. 40, p. 489-500.
James, A.V.G., 1920, Factors producing columnar structure in lavas and its occurrence near Melbourne, Australia: Lucas, P. T., and Drexler, J. M., 1976, Altamont-Bluebell—A major naturally fractured stratigraphie trap, Uinta Basin,
Journal of Geology, v. 28, p. 458-469. Utah: American Association of Petroleum Geologists Memoir No. 24, p. 121-135.
Jamison, W. R., and Spang, J. H., 1976, Use of calcite twin lamellae to infer differential stress: Geological Society of Lucas, S. E., and Moore, J. C., 1986, Cataclastic deformation in accretionary wedges: Deep Sea Drilling Project Leg 66,
America Bulletin, v. 87, p. 868-872. southern Mexico, and on-land examples from Barbados and Kodiak Islands: Geological Society of America
Jamison, W. R., and Stearns, D. W., 1982, Tectonic deformation of Wingate sandstone, Colorado National Monument: Memoir 166, p. 89-103.
American Association of Petroleum Geologists Bulletin, v. 66, p. 2584-2608. Lundberg, N., and Moore, J. C., 1986, Macroscopic structural features in Deep Sea Drilling Project cores from forearc
Jaroszweski, E. T., 1969, New site of tectonic stylolites: Bulletin de I'Academie Polonaise des Sciences, Series des sciences regions: Geological Society of America Memoir 166, p. 13-44.
géologie et géographie, v. 17, p. 17-23. Maddock, R. H., 1983, Melt origin of fault-generated pseudotachylytes demonstrated by textures: Geology, v. 11,
Johnson, J. A.D., and Budd, S. R., 1975, The geology of the Zone B and Zone C Lower Cretaceous limestone reservoirs of p. 105-108.
Asab Field, Abu Dhabi: 9th Arab Petroleum Congress, Paper No. 109 (B-3), p. 1-24. Maltman, A. J., 1977, Some microstructures of experimentally deformed argillaceous sediments: Tectonophysics, v. 39,
Judson, S., and Barks, R. E., 1961, Microstriations on polished pebbles: American Journal of Science, v. 259, p, 371-381. p. 417-436.
Kahle, C. F., and Floyd, J. G , 1971, Stratigraphie and environmental significance of sedimentary structures in Cayugan 1987, Shear zones in argillaceous sediments—An experimental study, in Jones, M. E., and Preston, R.M.F., eds.,
(Silurian) tidal flat carbonates, northwestern Ohio: Geological Society of America Bulletin, v. 82, p. 2071-2098. Deformation of sediments and sedimentary rocks: Geological Society Special Publication No. 29: Oxford, Eng-
Kahn, J. S., 1956, The analysis and distribution of packing in sand-size sediments. 1. On the measurement of packing in land, Blackwell Scientific Publications, p. 77-87.
sandstones: Journal of Geology, v. 64, p. 385-395. 1988, The importance of shear zones in naturally deformed wet sediments; Tectonophysics, v. 145, p. 163-175.
Kanaori, Y., 1986, A SEM cathodoluminescence study of quartz in mildly deformed granite from the region of the Mandl, G., de Jong, L.N.J., and Maltha, A., 1977, Shear zones in granular material. An experimental study of their
Atotsugawa fault, central Japan: Tectonophysics, v. 131, p. 133-146, structure and mechanical genesis: Rock Mechanics, v. 9, p. 95-144.
Keller, W. D., Stone, C. G., and Hoersch, A. L., 1985, Textures of Paleozoic chert and novaculite in the Ouachita March, A., 1932, Mathematische Theorie der Regelung nach der Korngestalt bei affiner Deformation: Zeitschrift für
Mountains of Arkansas and Oklahoma and their geological significance: Geological Society of America Bulletin, Kristallographie, v. 81, p, 285-297.
v. 96, p. 1353-1363. Marshak, S., and Engelder, T., 1985, Development of cleavage in limestones of a fold-thrust belt in eastern New York:
Kerrich, R., 1978, An historical review and synthesis of research on pressure solution: Zentralblatt fur Geologie und Journal of Structural Geology, v. 7, p. 345-359.
Palaontologie, Teil I, p. 512-550. Marshak, S., Geiser, P. A., Alvarez, W., and Engelder, T., 1982, Mesoscopic fault array of the northern Umbrian
Kerrich, R., and Allison, I., 1978, Row mechanisms in rocks: Geoscience Canada, v. 5, p. 109-118. Apennine fold belt, Italy: Geometry of conjugate shear by pressure-solution slip: Geological Society of America
Kerrich, R., Beckinsalé, R. D., and Durham, J. J., 1977, The transition between deformation regimes dominated by Bulletin, v. 93, p. 1013-1022.
intercrystalline diffusion and intracrystalline creep evaluated by oxygen isotope thermometry: Tectonophysics, Marshall, J. D., 1982, Isotopic composition of displacive fibrous calcite veins: Reversals in pore-water composition trends
v. 38, p. 241-257. during burial diagenesis: Journal of Sedimentary Petrology, v. 52, p, 615-630.
Kirby, S. H., 1980, Tectonic stresses in the lithosphère: Constraints provided by the experimental deformation of rocks: Mawer, C. K,, 1985, Comment on "Fault-related rocks: Suggestions for terminology": Geology, v. 13, p. 378.
Journal of Geophysical Research, v. 85, p. 6353-6363, Maxwell, J. C., 1960, Experiments on compaction and cementation of sand: Geological Society of America Memoir 79,
Klassen-Neklyudova, M. V., 1964, Mechanical twinning of crystals: New York, Consultants Bureau, 213 p. p. 105-132.
Knill, J. L., 1960, A classification of cleavages, with special references to the Craignish district of the Scottish Highlands: 1962, Origin of slaty and fracture cleavage in the Delaware Water Gap area, New Jersey and Pennsylvania, in
International Geological Congress, 21st, Proceedings, pt. 18, p. 317-325. Engel, A.E.J., James, H. L., and Leonard, B. F., eds., Petrologic studies (Buddington volume): Boulder, Colorado,
Knipe, R. J., 1986, Microstructural evolution of vein arrays preserved in Deep Sea Drilling Project cores from the Japan Geological Society of America, p. 281-311.
Trench, Leg 57: Geological Society of America Memoir 166, p. 75-87. Maxwell, J. G , and Verrall, P., 1954, Low porosity may limit oil in deep sands: World Oil, v. 138, no. 5, p. 106-113, and
Knipe, R, J., and White, S. H., 1977, Microstructural variation of an axial plane cleavage around a fold—A H.V.E.M. no. 6, p. 102-104.
study: Tectonophysics, v. 39, p. 355-380. McClay, K. R., 1977, Pressure solution and Coble creep in rocks and minerals: A review: Geological Society of London
Knopf, E. B., 1949, Fabric changes in Yule marble after deformation in compression: Part II, experimental deformation of Journal, v. 134, p. 57-70.
Yule marble: American Journal of Science, v. 247, p. 537-569. McClintock, F. A., and Walsh, J. B., 1962, Friction on Griffith cracks under pressure: United States National Congress of
Knopf, E. B., and Ingerson, E., 1938, Structural petrology: Geological Society of America Memoir 6, 270 p. Applied Mechanics, 4th, Proceedings, p. 1015-1021.
Kohlstedt, D. L., and Weathers, M. S., 1980, Deformation-induced microstructures, palaeopiezometers, and differential McEwen, T. J., 1977, Pressure solution or indentation: Comment: Geology, v. 5, p. 249-251.
stresses in deeply eroded fault zones: Journal of Geophysical Research, v. 85, p. 6269-6285. 1981, Brittle deformation in pitted pebble conglomerates: Journal of Structural Geology, v. 3, p. 25-37.
Koopman, A., 1983, Detachment tectonics in the central Apennines, Italy: Geologica Ultraiectina, v. 30, p. 1-155. McKenzie, D., and Brune, J. B., 1972, Melting on fault planes during large earthquakes: Royal Astronomical Society
Kowallis, B. J., Wang, H. F., and Jang, B.-A., 1987, Healed microcrack orientations in granite from Illinois borehole Geophysical Journal, v. 29, p. 65-78.
UPH-3 and their relationship to the rock's stress history: Tectonophysics, v. 135, p. 297-306. Means, W. D., 1987, A newly recognized type of slickenside striation: Journal of Structural Geology, v. 9, p. 585-590.
Kranz, R. L., 1983, Microcracks in rocks: A review: Tectonophysics, v. 100, p. 449-480. Means, W. D., and Paterson, M. S., 1966, Experiments on preferred orientation of platy minerals: Contributions to
Kubler, B., 1968, Evaluation quantitative du métamorphisme par la cristallinité de l'illite. Etat des progrès réalisés ces Mineralogy and Petrology, v. 13, p. 108-133.
dernières années: Centre Recherche Pau Bulletin, v. 2, p. 385-397. Merino, E., Ortoleva, P., and Strickholm, P., 1983, Generation of evenly spaced pressure-solution seams during (late)
Labaume, P., 1987, Syn-diagenetic deformation of a turbidite succession related to submarine gravity nappe emplacement, diagenesis: A kinetic theory: Contributions to Mineralogy and Petrology, v. 82, p. 360-370.
Autapie Nappe, French Alps, in Jones, M. E., and Preston, R.M.F., eds., Deformation of sediments and sedimen- Michael, A. J., 1984, Determination of stress from slip data: Faults and folds: Journal of Geophysical Research, v. 89,
tary rocks: Geological Society Special Publication No. 29: Oxford, England, Blackwell Scientific Publications, p. 11517-11526.
p. 147-163. Milnes, A. G., 1979, Albert Heim's general theory of natural rock deformation (1878): Geology, v. 7, p. 99-103.
Lachenbruch, A. H., 1962, Mechanics of thermal contraction cracks in ice-wedge polygons in permafrost: Geological Mimran, Y., 1975, Fabric deformation induced in Cretaceous chalks by tectonic stresses: Tectonophysics, v. 26,
Society of America Special Paper 70,69 p. p. 309-316.
Lajtai, E. Z., 1971, A theoretical and experimental evaluation of the Griffith theory of brittle fracture: Tectonophysics, 1977, Chalk deformation and large-scale migration of calcium carbonate: Sedimentology, v. 24, p. 333-360.
v. II, p. 129-156. 1985, Technically controlled freshwater carbonate cementation in chalk: Society of Economic Paleontologists and
Land, L. S., and Dutton. S, P., 1978. Cementation of Pennsylvanian deltaic sandstone: Isotopic data: Journal of Mineralogists Special Publication 36, p, 371-379.
Sedimentary Petrology, v. 48, p. 1167-1176. M&ik, M., 1971, Observations concerning calcite veinlets in carbonate rocks: Journal of Sedimentary Petrology, v. 41,
Langheinrich, G., and Plessman, W., 1968, Zue Entstehungsweise von Schieferungs-Flâchen in Kalksteinen (Turon-Kalke p. 450-460.
eines Salzauftriebs-Sattels im Harz-Vorland); Geologische Mitteilungen, v. 8, p. 111-142. Mitra, G., 1984, Brittle to ductile transition due to large strains along the White Rock thrust. Wind River Mountains,
Lapworth, C., 1885, The Highland controversy in British history: Its causes, course and consequences: Nature, v. 32, Wyoming: Journal of Structural Geology, v. 6, p. 51-62.
p, 558-559. Mitra, S., 1976, A quantitative study of deformation mechanisms and finite strain in quartzite: Contributions to Mineral-
Laubscher, H. P., 1979, Elements of Jura kinematics and dynamics: Eclogae Geologicae Helvetiae, v. 72, p, 467-483. ogy and Petrology, v. 59, p. 203-226.
Laurent, Ph., Bernard, Ph., Vasseur, G., and Etchecopar, A., 1981, Stress tensor determination from the study of e twins in —1987, Regional variations in deformation mechanisms and structural styles in the central Appalachian orogenic
calcite: A linear programming method: Tectonophysics, v. 78, p. 651-660. belt: Geological Society of America Bulletin, v. 98, p. 569-590.
Lavecchia, G., 1985, II sovrascorrimento dei Monti Sibillini: Anaiisi cinematica e strutturale: Bolletino Societa Geologica Mitra, S-, and Beard, W. C., 1980, Theoretical models of porosity reduction by pressure solution for well-sorted
Italiana, v. 104, p. 161-194. sandstones: Journal of Sedimentary Petrology, v. 50, p. 1347-1360.
Lawn, B. R., and Wilshaw, T. R., 1975, Fracture of brittle solids: Cambridge, England, Cambridge University Press, Mitra, S., and Tullis, J., 1979, A comparison of intracrystalline deformation in naturally and experimentally deformed
204 p. quartzites: Tectonophysics, v. 53, p. T21-T27.
Lawrence, R, D., 1970, Stress analysis based on albite twinning of plagioclase feldspars: Geological Society of America Mogi, K., 1971, Effect of the triaxial stress system on the failure of dolomite and limestone: Tectonophysics, v. 11,
Bulletin, v. 81, p. 2507-2512. p. 111-127.
Lee, J. H., Peacor, D. R., Lewis, D. D., and Wintsch, R. P., 1986, Evidence for syntectonic crystallization for the Möhr, O., 1900, Welche Umstände bedingen die Elastizitätsgrenze und den Bruch eines Materials?: Zeitschrift Vereins
mudstone to slate transition at Lehigh Gap, Pennsylvania, U.S. A.: Journal of Structural Geology, v. 8, p. 767-780. Deutsches Ingenieure, v. 44, p. 1524-30; 1572-77.
Leith, C. K„ 1905, Rock cleavage: U.S. Geological Survey Bulletin 239, 216 p. Moody, J. B., and Hundley-Goff, E. M., 1980, Microscopic characteristics of orthoquartzite from sliding friction experi-
Lemmleyn, G. G., and Kliya, M. O., 1960, Distinctive features of the healing of a crack in a crystal under conditions of ments. II. Gouge: Tectonophysics, v. 62, p. 301-319.
declining temperature: International Geology Review, v. 2, p. 125-128. Moody, J. D., and Hill, M. J., 1956, Wrench-fault tectonics: Geological Society of America Bulletin, v. 67, p. 1207-1246.
1358 R. H. GROSHONG, JR.

Moon, C. F., 1972, The microstructure of clay sediments: Earth-Science Reviews, v. 8, p. 303-321. can Journal of Science, v. 283, p. 936-966.
Moore, J. C., and Allwardt, A., 1980, Progressive deformation of a Tertiary trench slope, Kodiak Islands, Alaska: Journal 1986a, Ability of the Fry method to characterize pressure-solution deformation: Tectonophysics, v. 122,
of Geophysical Research, v. 85, p. 4741-4756. p. 187-193.
Moore, J. C., and Byrne, T., 1987, Thickening of fault zones: A mechanism of mélange formation in accreting sediments: 1986b, Ability of the Fry method to characterize pressure-solution deformation—Reply: Tectonophysics, v. 131,
Geology, v. 15, p. 1040-1043. p. 201-203.
Moore, J. G , and Geigle, J. E., 1974, Slaty cleavage: Incipient occurrences in the deep sea: Science, v. 183, p. 509-510. 1987, Ability of the Fry method to characterize pressure-solution deformation—Reply. Tectonophysics, v. 138,
Moore, J. G , Roeske, S., Lundberg, N., Schoonmaker, J., Cowan, D. S., Gonzales, E., and Lucas, S. E., 1986, Scaly p. 326.
fabrics from Deep Sea Drilling Project cores from forearcs: Geological Society of America Memoir 166, p. 55-73. Ord, A., and Christie, J. M., 1984, Flow stresses from microstructures in mylonitic quartzites of the Moine thrust zone,
Morganstern, N. R., and Tchalenko, J. S., 1967a, Microscopic structures in kaolin subjected to direct shear Geotechnique, Assynt area, Scotland: Journal of Structural Geology, v. 6, p. 639-654.
v. 17, p. 309-328. Park, W. G , and Schott, E. H., 1968, Stylolites: Their nature and origin: Journal of Sedimentary Petrology, v. 38,
1967b, Microstructural observations on shear zones from slips in natural clays: Geotechnical Conference, Oslo, p. 175-191.
Proceedings, v. 1, p. 147-152. Parker, J. M., Ill, 1942, Regional systematic jointing in slightly deformed sedimentary rocks: Geological Society of
Morrow, C. A., Shi, L. Q., and Byerlee, J. D., 1984, Permeability of fault gouge under confining pressure and shear stress: America Bulletin, v. 53, p. 381-408.
Journal of Geophysical Research, v. 89, p. 3193-3200. 1969, Jointing in south-central New York: Discussion: Geological Society of America Bulletin, v. 80, p. 919-922.
Mosher, S., 1987, Pressure-solution deformation of the Purgatory Conglomerate, Rhode Island (U.S.A.): Quantification of Paterson, M. S., 1958, Experimental deformation and faulting in Wombeyan marble: Geological Society of America
volume change, real strains and sedimentary shape factor: Journal of Structural Geology, v. 9, p. 221-232. Bulletin, v. 69, p. 465-476.
Mossop, G. D., 1972, Origin of the peripheral rim, Redwater reef, Alberta: Bulletin of Canadian Petroleum Geology, 1973, Nonhydrostatic thermodynamics and its geologic applications: Reviews of Geophysics, v. 11, p. 355-389.
v. 20, p. 238-280. 1978, Experimental rock deformation: Berlin, Springer-Verlag, 254 p.
Muehlberger, W. R., 1961, Conjugate joint sets of small dihedral angle: Journal of Geology, v. 69, p. 211-219. Paterson, M. S., and Turner, F. J., 1970, Experimental deformation of constrained crystals of calcite in extension, in
Mügge, O., and Heide, F., 1931, Einfache Schiebung am Anorthit: Neues Jahrbuch für Mineralogie, Geologie und Paulitsch, P., ed., Experimental and natural rock deformation: New York, Springer-Verlag, p. 109-141.
Paläontologie Abhandlungen, A., v. 64, p. 161-170. Pavoni, N., 1980, Comparison of focal mechanisms of earthquakes and faulting in the Helvetic zone of the central Valais,
Mullenax, A. C,, Gray, D. R., 1984, Interaction of bed-parallel stylolites and extension veins in boudinage: Journal of Swiss Alps: Ectogae Geologicae Helvetiae, v. 73, p. 551-5S8.
Structural Geology, v. 6, p. 63-72. Peck, D. L., and Minakami, T., 1968, The formation of columnar joints in the upper part of Kilauean lava lakes, Hawaii:
Murreil, S.A.F., 1963, A criterion for brittle fracture of rocks and concrete under triaxial stress and effect of pore pressure Geological Society of America Bulletin, v. 79, p. 1151-1166.
on the criterion, in Fairhurst, G , ed., Rock mechanics: Rock Mechanics Symposium, 5th, University of Minnesota, Peng, S., and Johnson, A. M., 1972, Crack growth and faulting in cylindrical specimens of Chelmsford granite: Interna-
Proceedings: Oxford, England, Pergamon, p. 563-577. tional Journal of Rock Mechanics and Mining Science, v. 9, p. 37-86.
Nàdai, A., 1931, Plasticity: New York, McGraw-Hill, 349 p. Peterson, G. L., 1966, Structural interpretation of sandstone dikes, northwest Sacramento Valley, California: Geological
1950. Theory of flow and fracture of solids: New York, McGraw-Hill Book Company, 571 p. Society of America Bulletin, v. 77, p. 833-842.
Narahara, D. K., and Wiltschko, D. V., 1986, Deformation in the hinge region of a chevron fold, Valley and Ridge 1968, Flow structures in sandstone dikes: Sedimentary Geology, v. 2, p. 177-190.
province, central Pennsylvania: Journal of Structural Geology, v. 8, p. 157-168. Petit, J. P., 1987, Criteria for the sense of movement on fault surfaces in brittle rocks: Journal of Structural Geology, v. 9,
Narr, W., and Burruss, R. C., 1984, Origin of reservoir fractures in Little Knife Field, North Dakota: American p. 597-608.
Association of Petroleum Geologists Bulletin, v. 68, p. 1087-1100. Petit, J. P., and Laville, E., 1987, Morphology and microstructures of hydroplastic slickensides in sandstone, in Jones,
Naumann, C. F., 1858, Lehrbuch der Geonosie (2nd edition): Leipzig, Engelmann, 3 volumes, 2,053 p. M. E., and Preston, R.M.F., eds., Deformation of sediments and sedimentary rocks: Geological Society Special
Neal, J. T., Langer, A. M., and Kerr, P. F., 1968, Giant desiccation polygons of Great Basin playas: Geological Society of Publication No. 29: Oxford, England, Blackwell Scientific Publications, p. 107-121.
America Bulletin, v. 79, p. 69-90. Petit, J. P., Proust, F., and Tapponnier, P., 1983, Criteres de sens de mouvement sur les miroirs de faille en roches non
Nelson, K. D., 1982, A suggestion for the origin of mesoscopic fabric in accretionary mélange, based on features observed calcaires: Bulletin Société Géologique de France, v. 25, p. 589-608.
in the Chrystalls Beach Complex, South Island, New Zealand: Geological Society of America Bulletin, v. 93, Pfiffner, O. A., and Burkhard, M., 1987, Determination of paleo-stress axes orientations from fault, twin and earthquake
p. 625-634. data: Annales Tectonicae, v. 1, p. 48-57.
Nelson, R. A., 1979, Natural fracture systems: Description and classification: American Association of Petroleum Geolo- Phillips, J.-G, 1982, Character and origin of cataclasite developed along the low-angle Whipple detachment fault,
gists Bulletin, v. 63, p. 2214-2232. Whipple Mountains, California, in Frost, E. G., and Martin, D. L., eds., Mesozoic-Cenozoic tectonic evolution of
1981, Significance of fracture sets associated with stylolite zones: American Association of Petroleum Geologists the Colorado River region, California, Arizona, and Nevada: San Diego, California, Cordilleran Publishers,
Bulletin, v. 65, p. 2417-2425. p. 109-116.
1983, Localization of aggregate stylolites by rock properties: American Association of Petroleum Geologists Phillips, W. J., 1972, Hydraulic fracturing and mineralization: Geological Society of London Journal, v. 128, p. 337-359.
Bulletin, v. 67, p. 313-319. Philpotts, A. R., 1964, Origin of pseudotachylites: American Journal of Science, v. 262, p. 1008-1035.
1985, Geologic analysis of naturally fractured reservoirs: Houston, Texas, Gulf Publishing Company, 320 p. Pickering, K. T., 1987, Wet-sediment deformation in the Upper Ordovician Point Leamington Formation: An active
Nelson, R. A., and Stearns, D. W., 1977, Interformational control of regional fracture orientations: Rocky Mountain thrust-imbricate system during sedimentation, Notre Dame Bay, north-central Newfoundland, in Jones, M. E., and
Association of Geologists Symposium, p. 95-101. Preston, R.M.F., eds., Deformation of sediments and sedimentary rocks: Geological Society Special Publication
Nemat-Nasser, S., and Horii, H., 1982, Compression-induced nonplanar crack extension with application to splitting, No. 29: Oxford, England, Blackwell Scientific Publications, p. 213-239.
exfoliation and rockburst: Journal of Geophysical Research, v. 87, p. 6805-6821. Pierson, B. J., and Shinn, E. A., 1985, Cement distribution and carbonate mineral stabilization in Pleistocene limestones of
Neumann, E.-R., 1969, Experimental recrystallization of dolomite and comparison of preferred orientations of calcite and Hogsty Reef, Bahamas, in Schneidermann, N., and Harris, P. M., eds., Carbonate cements: Society of Economic
dolomite in deformed rocks: Journal of Geology, v. 77, p. 426-438. Paleontologists and Mineralogists Special Publication 36, p. 153-168.
Nevin, C. M„ 1949, Principles of structural geology (4th edition): New York, John Wiley & Sons, Inc., 410 p. Pittman, E. D., 1981, Effect of fault-related granulation on porosity and permeability of quartz sandstones, Simpson
Newson, J. F., 1903, Clastic dikes: Geological Society of America Bulletin, v. 14, p. 227-268. Group (Ordovician), Oklahoma: American Association of Petroleum Geologists Bulletin, v. 65, p. 2381-2387.
Nicholson, R., and Pollard, D. D., 1985, Dilation and linkage of echelon cracks: Journal of Structural Geology, v. 7, Platt, J. P., and Leggett, J. K., 1986, Stratal extension in thrust footwalls, Makran accretionary prism: Implications for
p. 583-590. thrust tectonics: American Association of Petroleum Geologists Bulletin, v. 70, p. 191-203.
Nickelsen, R. P., 1966, Fossil distortion and penetrative rock deformation in the Appalachian Plateau, Pennsylvania: Platt, J. P., and Vissers, R.L.M., 1980, Extensional structures in anisotropic rocks: Journal of Structural Geology, v. 2,
Journal of Geology, v. 74, p. 924-931. p. 397-410.
1972, Attributes of rock cleavage in some mudstones and limestones of the Valley and Ridge province, Pennsylva- Plessmann, W„ 1965, Gesteinslösung, ein Hauptfaktor beim Schieferungsprozess: Geologische Mitteilungen, v. 4 (1963),
nia: Pennsylvania Academy of Science, v. 46, p. 107-112. p. 69-82.
1979, Sequence of structural stages of the Alleghany orogeny at the Bear Valley strip mine, Shamokin, Pennsylva- 1972, Horizontai-stylolithen im französisch-schweizerischen Tafel—und Faltenjura und ihre Einpassung in den
nia: American Journal of Science, v. 279, p. 225-271. regionalen Rahmen: Geologische Rundschau, v. 61, p. 332-347.
1986, Cleavage duplexes in the Marcellus Shale of the Appalachian foreland: Journal of Structural Geology, v. 8, Hessmann, W., and Spaeth, G., 1971, Sedimentgänge und tektonisches Schichtfliessen (Biegungsfliessen) im Rechtsrhein-
p. 361-371. ischen Schiefergebirge: Geologische Mitteilungen, v. 11, p. 137-164.
Nickelsen, R. P., and Hough, V.N.D., 1967, Jointing in the Appalachian Plateau of Pennsylvania: Geological Society of Poirier, J.-P., 1985, Creep of crystals; high-temperature deformation processes in metals, ceramics and minerals: New
America Bulletin, v. 78, p. 609-630. York, Cambridge University Press, 260 p.
Nicolas, A., and Poirier, J. P., 1976, Crystalline plasticity and solid state flow in metamorphic rocks: New York, John Pollard, D. D., Segall, P., and Delaney, P. T., 1982, Formation and interpretation of dilatent echelon cracks: Geological
Wiley & Sons, 444 p. Society of America Bulletin, v. 93, p. 1291-1303.
Nissen, H. U., 1964a, Calcite fabric analysis of deformed oolites from the South Mountain fold, Maryland: American Powell, C. McA., 1972, Tectonic dewatering strain in the Michigamme slate, Michigan: Geological Society of America
Journal of Science, v. 262, p. 892-903. Bulletin, v. 83, p. 2149-2158.
1964b, Dynamic and kinematic analysis of deformed crinoid stems in a quartz graywacke: Journal of Geology, 1979, A morphological classification of rock cleavage: Tectonophysics, v. 58, p. 21-34.
v. 72, p. 346-360. Price, N. J., 1966, Fault and joint development in brittle and semi-brittle rock: London, England, Pergamon Press, 176 p.
Norris, D. K., and Barron, K., 1969, Structural analysis of features on natural and artificial faults: Geological Survey of Price, N. J., and Hancock, P. L., 1972, Development of fracture cleavage and kindred structures: International Geological
Canada Paper 68-52, p. 136-167. Congress, 24th, Montreal, Section 3, p. 584-592.
Odom, I. E., 1967, G a y fabric and its relation to structural properties in mid-continent Pennsylvanian sediments: Journal Price, R. A., 1967, The tectonic significance of mesoscopic subfabrics in the southern Rocky Mountains of Alberta and
of Sedimentary Petrology, v. 37, p. 610-623. British Columbia: Canadian Journal of Earth Sciences, v. 4, p. 39-70.
Oertel, G., 1962, Stress, strain and fracture in clay models of geologic deformation: Geotimes, v. 6, p. 26-31. Prior, D. J., Knipe, R. J., Bates, M. P., Grant, N. T., Law, R. D., Lloyd, G. E., Welbon, A., Agar, S. M., Brodie, K. H.,
1965, The mechanism of faulting in clay experiments: Tectonophysics, v. 2, p. 343-393. Maddock, R. H., Rutter, E. H., White, S. H., Bell, T. H., Ferguson, C. C., and Wheeler, J., 1987, Orientation of
1970, Deformation of a slaty, lapillar tuff in the Lake District, England: Geological Society of America Bulletin, specimens: Essential data for all fields of geology: Geology, v. 15, p. 829-831.
v. 81, p. 1173-1188. Proctor, R. J., Payne, C. M., and Kalin, D. C., 1970, Crossing the Sierra Madre fault zone in the Glendora tunnel, San
1983, The relationship of strain and preferred orientation of phyllosilicate grains in rocks—A review: Tectono- Gabriel Mountains, California: Engineering Geology, v. 4, p. 5-63.
physics, v. 100, p. 413-447. Ramberg, H., 1955, Natural and experimental boudinage and pinch-and-swell structures: Journal of Geology, v. 63,
Oertel, G., and Curtis, C. D., 1972, Clay-ironstone concretion preserving fabrics due to progressive compaction: Geologi- p. 512-526.
cal Society of America Bulletin, v. 83, p. 2597-2606. Ramsay, J. G., 1967, Folding and fracturing of rocks: New York, McGraw-Hill, Inc., 568 p.
Oertel, G., and Phakey, P. P., 1972, The texture of a slate from Nantlle, Caernarvon, North Wales: Texture, v. I, p. 1-8. 1980a, Shear zone geometry: A review: Journal of Structural Geology, v. 2, p. 83-99.
Oldershaw, A. E., and Scoffin, T. P., 1967, The source of ferroan and non-ferroan calcite cements in the Halkin and 1980b, The crack-seal mechanism of rock deformation: Nature, v. 284, p. 135-139.
Wenlock limestones: Geological Journal, v. 5, p. 309-320. Ramsay, J. G., and Graham, R. H., 1970, Strain variation in shear belts: Canadian Journal of Earth Sciences, v. 7,
Olgaard, D. L., and Brace, W. F., 1983, The microstructure of gouge from a mining-induced seismic shear zone: p. 786-813.
International Journal of Rock Mechanics and Mining Science, v. 20, p. 11-19. Ramsay, J. G., and Huber, M. I., 1983, The techniques of modern structural geology, Volume 1: Strain analysis: London,
Olsson, W. A., 1974, Grain size dependence of yield stress in marble: Journal of Geophysical Research, v. 79, England, Academic Press, 307 p.
p. 4859-4862. 1987, The techniques of modern structural geology, Volume 2: Folds and fractures: London, England, Academic
Olsson, W. A., and Peng, S. S., 1976, Microcrack nucleation in marble: International Journal of Rock Mechanics and Press, p. 309-700.
Mining Science, v. 13, p. 53-59. Ranalli, G., 1984, Grain size distribution and flow stress in tectonites: Journal of Structural Geology, v. 6, p. 443-447.
Onasch, C. M., 1983a, Dynamic analysis of rough cleavage in the Martinsburg Formation, Maryland: Journal of Raymond, L A., 1984, Classification of mélanges: Geological Society of America Special Paper 198, p. 7-20.
Structural Geology, v. 5, p. 73-81. 1985, Comment on "Fault-related rocks: Suggestions for terminology": Geology, v. 13, p. 218.
1983b, Origin and significance of microstructures in sandstones of the Martinsburg Formation, Maryland: Ameri- Reches, Z., 1978, Analysis of faulting in three-dimensional strain field: Tectonophysics, v. 47, p. 109-129.
LOW-TEMPERATURE DEFORMATION MECHANISMS 1359

1983, Faulting in rocks in three-dimensional strain fields. II. Theoretical analysis: Tectonophysics, v. 95, Shainin, V. E., 1950, Conjugate sets of en-enchelon tension fractures in the Athens Limestone at Riverton, Virginia:
p. 133-156. Geological Society of America Bulletin, v. 61, p. 509-517.
Reches, Z., and Dieterich, J. H., 1983, Faulting of rocks in three-dimensional strain fields. I. Failure of rocks in polyaxial, Shand, S. J., 1916, The pseudotachylyte of Parijs (Orange Free State): Geological Society of London Quarterly Journal,
servo-control experiments: Tectonophysics, v. 95, p. 111-132. v. 72, p. 198-221.
Reches, Z., and Johnson, A. M., 1976, A theory of concentric, kink and sinusoidal folding and of monoclinal flexuring of Sharpe, D., 1847, On slaty cleavage: Geological Society of London Quarterly Journal, v. 3, p. 74-105.
compressible, elastic multilayers: Tectonophysics, v. 35, p. 295-334. Shinn, E. A., and Robbin, D. M., 1983, Mechanical and chemical compaction in fine-grained shallow-water limestones:
Reid, H. F., 1910, Mechanics of the earthquake, in Lawson, A. G , ed., The California earthquake of 1906, Volume II: Journal of Sedimentary Petrology, v. 53, p. 595-618.
Washington, D.C., Carnegie Institution of Washington, 192 p. Shuaib, S. M., 1973, Subsurface petrographic study of joints in variegated siltstone-sandstone and Khairbad limestone,
Reik, G. A., and Currie, J. B., 1974, A study of relations between rock fabric and joints in sandstone: Canadian Journal of Pakistan: American Association of Petroleum Geologists, v. 57, p.. 1775-1778.
Earth Sciences, v. 11, p. 1253-1268. Sibley, D. F., and Blatt, H., 1976, Intergranular pressure solution and cementation of the Tuscarora orthoquartzite:
Reimold, W. U., Oskierski, W., and Huth, J., 1987, The pseudotachylite from Champagnac in the Rochecouart Meteorite Journal of Sedimentary Petrology, v. 46, p. 881-896.
Crater, France: Journal of Geophysical Research, v. 92, p. E737-E748. Sibson, R. H., 1977, Fault rocks and fault mechanisms: Journal of the Geological Society of London, v. 133, p. 191-213.
Rejebian, V. A., Harris, A. G„ and Huebner, J. S., 1987, Conodont color and textural alteration: An index to regional 1980, Transient discontinuities in ductile shear zones: Journal of Structural Geology, v. 2, p. 165-171.
metamorphism, contact metamorphism, and hydrothermal alteration: Geological Society of America Bulletin, 1986a, Brecciation processes in fault zones: Inferences from earthquake rupturing: Pure and Applied Geophysics,
v. 99, p. 471-479. v. 124, p. 159-175.
Reks, 1. J., and Gray, D. R., 1982, Pencil structure and strain in weakly deformed mudstone and siltstone: Journal of 1986b, Earthquakes and rock deformation in crustai fault zones: Annual Review of Earth and Planetary Sciences,
Structural Geology, v. 4, p. 161-176. v. 14, p. 149-175.
Renton, J. J., Heald, M. T., and Cecil, C. B., 1969, Experimental investigation of pressure solution of quartz: Journal of 1987, Earthquake rupturing as a mineralizing agent in hydrothermal systems: Geology, v. 15, p. 701-704.
Sedimentary Petrology, v. 39, p. 1107-1117. Siddans, A.W.B., 1976, Deformed rocks and their textures: Royal Society of London Philosophical Transactions, ser. A,
Richter, D., 1963, Verkürzung von Fossilien und Entstehung von Flaser—und Knollenkalken durch Lösungsvorgänge in v. 283, p. 43-54.
geschieferten kalkigen Gesteinen: Geologische Mitteilungen, v. 4, p. 235-248. Sieh, K. E., 1984, Lateral offsets and revised dates of large prehistoric earthquakes at Pallett Creek, southern California:
Rickard, M. J , 1961, A note on cleavages in crenulated rocks: Geological Magazine, v. 98, p. 324-332. Journal of Geophysical Research, v. 89, p. 7641-7670.
Rickard, M. J., and Rixon, L. K., 1983, Stress configurations in conjugate quartz-vein arrays: Journal of Structural Simpson, C., and Schmid, S. M., 1983, An evaluation of criteria to deduce the sense of movement in sheared rocks:
Geology, v. 5, p. 573-578. Geological Society of America Bulletin, v. 94, p. 1281-1288.
Riedel, W., 1929, Zur Mechanik geologischer Brucherscheinungen: Centraiblatt für Mineralogie, Geologie und Paläontol- Simpson, J., 1985, Stylolite-controlled layering in an homogeneous limestone: Pseudo-bedding produced by burial
ogie, Abteilung B, Geologie und Paläontologie, p. 354-368. diagenesis: Sedimentology, v. 32, p. 495-505.
Rigby, J. K., 1953, Some transverse stylolites; Journal of Sedimentary Petrology, v. 23, p. 265-271. Sippel, R. F., 1968, Sandstone petrology, evidence from luminescence petrography: Journal of Sedimentary Petrology,
Rispoli, R., 1981, Stress fields about strike-slip faults inferred from stylolites and tension gashes: Tectonophysics, v. 75, v. 38, p. 530-554.
p. T29-T36. Skempton, A. W., 1966, Some observations on tectonic shear zones: Congress of the International Society for Rock
Rittenhouse, G., 1971a, Mechanical compaction of sands containing different percentages of ductile grains: A theoretical Mechanics, 1st, Proceedings, v. 1, p. 329-335.
approach: American Association of Petroleum Geologists Bulletin, v. 55, p. 92-96. Smith, D. A., 1980, Sealing and nonsealing faults in Louisiana Gulf Coast salt basin: American Association of Petroleum
197 lb, Pore-space reduction by solution and cementation: American Association of Petroleum Geologists Bulletin, Geologists Bulletin, v. 64, p. 145-172.
v. 55, p. 80-91. Smith, D. L., and Evans, B., 1984, Diffusional crack healing in quartz: Journal of Geophysical Research, v. 89,
1973, Pore-space reductions in sandstones—controlling factors and some engineering implications: Offshore p. 4125-4136.
Technology Conference, 5th Annual, Houston, Texas (preprint), p. 1683-1688. Smith, K. G., 1952, Structure plan of clastic dikes: American Geophysical Union Transactions, v. 33, p. 889-892.
Roberts, J. C., 1961, Feather-fracture, and the mechanics of rock-jointing: American Journal of Science, v. 259, Sorby, H. C., 1853, On the origin of slaty cleavage: Edinburgh New Philosophical Journal, v. 55, p. 137-149.
p. 481-492. 1860, On the origin of "cone-in-cone": British Association for the Advancement of Science, report of 29th
1965, Quartz microfracturing in the north crop of the South Wales Coalfield: Geological Magazine, v. 102, meeting, 1859, Transactions of Sections, Geology, p. 124.
p. 59-72. 1863, On the direct correlation of mechanical and chemical forces: Royal Society of London Proceedings, v. 12,
Robertson, E. C., 1955, Experimental study of the strength of rocks: Geological Society of America Bulletin, v. 66, p. 538-600.
p.1275-1314. 1880, On the structure and origin of non-calcareous stratified rocks: Geological Society of London Quarterly
Robin, P.-Y. F., 1978, Pressure solution at grain-to-grain contacts: Geochimica et Cosmochimica Acta, v. 42, Journal (Proceedings), v. 36, p. 46-92.
p.1383-1389. 1908, On the application of quantitative methods to the structure and history of rocks: Geological Society of
Rod, E., 1966, A discussion of the paper: "Fault plane features: An alternative explanation" by R. E. Riecker: Journal of London Quarterly Journal, v. 64, p. 171-233.
Sedimentary Petrology, v. 36, p. 1163-1165. Spang, J. H., 1972, Numerical method for dynamic analysis of calcite twin lamellae: Geological Society of America
Roering, C., 1968, The geometrical significance of natural en-echelon crack-arrays: Tectonophysics, v. 5, p. 107-123. Bulletin, v. 83, p. 467-472,
Roy, A. B., 1978, Evolution of slaty cleavage in relation to diagenesis and metamorphism: a study from the Hunsruck- Spang, J. H., and Brown, S. P., 1981, Dynamic analysis of a small imbricate thrust and related structures, Front Ranges,
schiefer: Geological Society of America Bulletin, v. 89, p. 1775-1785. southern Canadian Rocky Mountains, in McClay, K. R., and Price, N. J., eds,, Thrust and nappe tectonics: Boston,
Rummel, F., 1987, Fracture mechanics approach to hydraulic fracturing stress measurements, in Atkinson, B. K., ed., Massachusetts, Blackwell Scientific Publications, p. 143-149.
Fracture mechanics of rock: London, England, Academic Press, p. 217-239. Spang, J. H., and Friedman, M., 1978, Analysis of twinning in naturally and experimentally deformed dolomite: EOS
Rutter, E. H., 1983, Pressure solution in nature, theory and experiment: Geological Society of London Journal, v. 140, (American Geophysical Union Transactions), v. 59, p. 1186.
p. 725-740. Spang, J. H., and Groshong, R. H., Jr., 1981, Deformation mechanisms and strain history of a minor fold from the
1986, On the nomenclature of mode of failure transitions in rocks: Tectonophysics, v. 122, p. 381-387. Appalachian Valley and Ridge province: Tectonophysics, v. 72, p. 323-342.
Rutter, E. H., Mäddock, R. H„ Hall, S. H., and White, S. H., 1986, Comparative microstructures of natural and Spang, J. H., and Van Der Lee, J., 1975, Numerical dynamic analysis of quartz deformation lamellae and calcite and
experimentally produced clay-bearing fault gouges: Pure and Applied Geophysics, v. 124, p. 3-30. dolomite twin lamellae: Geological Society of America Bulletin, v. 86, p, 1266-1272.
Ryan, M. P., and Sammis, C. G., 1978, Cyclic fracture mechanisms in cooling basalt: Geological Society of America Spang, J. H., Oldershaw, A. E., and Groshong, R. H., Jr., 1974, The nature of thin twin lamellae in calcite [abs.]: EOS
Bulletin, v. 89, p. 1295-1308. (American Geophysical Union Transactions), v. 55, p. 419.
Rye, D. M., and Bradbury, H. J.. 1988, Fluid flow in the crust: An example from a Pyrenean thrust ramp: American Spang, J. H., Oldershaw, A. E., and Stout, M. Z., 1979, Development of cleavage in the Banff Formation at Pigeon
Journal of Science, v. 288, p. 197-235. Mountain, Front Ranges, Canadian Rocky Mountains: Canadian Journal of Earth Sciences, v. 16, p. 1108-1115.
Sandberg, P., 1985, Aragonite cements and their occurrence in ancient limestones, in Schneidermann, N., and Harris, Spears, D. A., 1976, The fissility of some Carboniferous shales: Sedimentology, v. 23, p. 721-725.
P. M., eds., Carbonate cements: Society of Economic Paleontologists and Mineralogists Special Publication 36, Spiers, C. J., 1979, Fabric development in calcite polycrystals deformed at 400 °C: Bulletin of Mineralogie, v. 102,
p. 33-57. p. 282-289.
Sander, B., 1911, Über Zusammenhänge zwischen Teilbewegung und Gefüge in Gesteinen: Tschermaks Mineralogische Spray, J. G., 1987, Artificial generation of pseudotachylyte using friction welding apparatus: Simulation of melting on a
Petrographische Mitteilungen, v. 30, p. 281-314. fault plane: Journal of Structural Geology, v. 9, p. 49-60.
1930, Gefügekunde der Gesteine: Vienna, Springer, 352 p. Sprunt, E. S., and Nur, A., 1976, The reduction of porosity by pressure solution: Experimental verification: Geology, v. 4,
Schlanger, S. O., 1964, Petrology of the limestones of Guam: U.S. Geological Survey Professional Paper 403-D, 52 p. p. 463-466.
Schmid, E , and Boas, W., 1950, Plasticity of crystals: London, England, F. A. Hughes, 353 p. 1977, Destruction of porosity through pressure solution: Geophysics, v. 42, p. 726-741.
Schmid, S. M., 1975, The Glarus overthrust: Field evidence and mechanical model: Eclogae Geoiogicae Helvetiae, v. 68, 1979, Microcracking and healing in granites: New evidence from cathodoluminescence: Science, v. 205,
p. 247-280. p. 495-497.
1982, Microfabric studies as indicators of deformation mechanisms and flow taws operative in mountains build- Spry, A., 1969, Metamorphic textures: Oxford, England, Pergamon Press, 350 p.
ing, in Hsu, J., ed., Mountain building processes; London, England, Academic Press, p. 95-110. Stanton, R. L., 1972, Ore petrology: New York, McGraw-Hill Book Company, 713 p.
Schmid, S. M., and Paterson, M. S., 1977, Strain analysis in an experimentally deformed oolitic limestone, in Saxena, Stearns, D. W., 1969, Fracture as a mechanism of flow in naturally deformed layered rocks, in Baer, A. J., and Norris,
S. K., and Bhattachargi, S„ eds., Energetics of geological processes: New York, Springer-Verlag, p. 67-93. D. K., eds., Conference on Research in Tectonics, Proceedings: Canada Geological Survey Paper 68-52, p. 79-90.
Schmid, S. M., Boland, J. N., and Paterson, M. S., 1977, Superplastic flow in fine grained limestone: Tectonophysics, 1972, Structural interpretation of fractures associated with the Bonita fault: New Mexico Geological Society Field
v. 43, p. 257-292. Conference Guidebook No. 23, p. 161-164.
Schmid, S. M., Casey, M., and Starkey, J., 1981, The microfabric of calcite tectonites from the Helvetic nappes (Swiss Stel, H., 1981, Crystal growth in cataclasites: ¡Diagnostic microstructures and implications: Tectonophysics, v. 78,
Alps), in McCtay, K. R„ and Price, N. J., eds., Thrust and nappe tectonics: Boston, Massachusetts, Blackwell p. 585-600.
Scientific Publications, p. 151-158. 1986, The effect of cyclic operation of brittle and ductile deformation on the metamorphic assemblage in
Schmid, S. M., Paterson, M. S., and Boland, J. N., 1980, High temperature flow and dynamic recrystallization in Carrara cataclasites and mylonites: Pure and Applied Geophysics, v. 124, p. 289-307.
marble: Tectonophysics, v. 65, p. 245-280. Stockdale, P. B., 1922, Stylolites: Their nature and origin: Indiana University Studies, v. 9, p. 1-97.
Scholz, C. H., 1968a, Correction to paper by C. H. Scholz "Experimental study of the fracturing process in brittle rock": 1926, The stratigraphic significance of solution in rocks: Journal of Geology, v. 34, p. 399-414.
Journal of Geophysical Research, v. 73, p. 4794. Stocker, R. L., and Ashby, M. F., 1973, On the rheology of the upper mantle: Reviews of Geophysics and Space Physics,
1968b, Experimental study of the fracturing process in brittle rock: Journal of Geophysical Research, v. 73, v. 11, p. 391-426.
p. 1447-1454. Summers, R., and Byerlee, J., 1977, A note on the effect of fault gouge composition on the stability of frictional sliding:
Schwander, H. W., Bürgin, A., and Stern, W. B„ 1981, Some geochemical data on stylolites and their host rocks: Eclogae International Journal of Rock Mechanics and Mining Sciences, v. 14, p. 155-160.
Geoiogicae Helvetiae, v. 74, p. 217-224. Syme Gash, P. J., 1971, A study of surface features relating to brittle and semi-brittle fracture: Tectonophysics, v. 12,
Schweitzer, J., and Simpson, C., 1986, Cleavage development in dolomite of the Elbrook Formation, southwest Virginia: p. 349-391.
Geological Society of America Bulletin, v. 97, p. 778-786. Tada, R., and Siever, R., 1986, Experimental knife-edge pressure solution of halite: Geochimica et Cosmochimica Acta,
Secor, D. T., 1965, Role of fluid pressure in jointing: American Journal of Science, v. 263, p. 633-646. v. 50, p. 29-36.
Segal!, P., and Pollard, D. D„ 1980, Mechanics of discontinuous faults: Journal of Geophysical Research, v. 85, Talbot, C. J., 1970, The minimum strain ellipsoid using deformed quartz veins: Tectonophysics, v. 9, p. 47-76.
p. 4337-4350. Tapp, B., and Cook, J., 1988, Pressure solution zone propagation in naturally deformed carbonate rocks: Geology, v. 16,
1983a, Joint formation in granitic rock of the Sierra Nevada: Geological Society of America Bulletin, v. 94, p. 182-185.
p. 563-575. Tapp, B., and Wickham, J., 1987, Relationships of rock cleavage fabrics to incremental and accumulated strain in the
1983b, Nucleation and growth of strike slip faults in granite: Journal of Geophysical Research, v. 88, p. 555-568. Conococheague Formation, U.S.A.: Journal of Structural Geology, v. 9, p. 457-472.
Seifert, K. W.. 1965, Deformation bands in albite: American Mineralogist, v. 50, p. 1469-1472. Taylor, G. I., 1938, Plastic strain in metals: Journal of Institute of Metals, Great Britain, v. 62, p. 307-324.
Downloaded from gsabulletin.gsapubs.org on February 27, 2013

1360 R. H. GROSHONG, JR.

Taylor, J. M., 19S0, Pore-space reduction in sandstones: American Association of Petroleum Geologists Bulletin, v. 34, Waters, A. C., and Campbell, C. D., 1935, Mylonites from the San Andreas fault zone: American Journal of Science,
p. 701-716. v. 29, p. 473-503.
Tchalenko, J. S., 1968, The evolution of kink-bands and the development of compression textures in sheared clays: Watts, M. J., and Williams, G. D., 1980, Fault rocks as indicators of progressive shear deformation in the Guingamp
Tectonophysics, v. 6, p. 159-174. region, Brittany: Journal of Structural Geology, v. 1, p. 323-332.
1970, Similarities between shear zones of different magnitude: Geological Society of America Bulletin, v. 81, Wawersik, W. R., and Brace, W. F., 1971, Post-failure behavior of a granite and diabase: Rock Mechanics, v. 3, p. 61-85.
p. 1625-1640. Weaver, C. E., 1984, Shale-slate metamorphism in southern Appalachians: Amsterdam, the Netherlands, Elsevier, 239 p.
Teichmiiller, M., and Teichmiiller, R., 1981, The significance of coalification studies to geology—A review: Bulletin Weber, K. J., and Bakker, M., 1981, Fracture and vuggy porosity: Society of Petroleum Engineers, 56th Annual Fall
Centres Recherches Exploration-Production Elf-Aquitaine, v. 5, p. 491-534. Technical Conference and Exhibition, San Antonio, Texas, 11 p.
Teufel, L. W., 1980, Strain analysis of experimental superposed deformation using calcite twin lamellae: Tectonophysics, Weber, K. J., Mandl, G., Pilaar, W. F., Lehner, F., and Precious, R. G., 1978, The role of faults in hydrocarbon
v. 65, p. 291-309. migration and trapping in Nigerian growth fault structures: 10th Annual Society of Petroleum Engineers of Ameri-
1981, Pore volume changes during frictional sliding of simulated faults: American Geophysical Union Monograph can Institute of Mechanical Engineers Offshore Technology Conference Preprint No. OTC-3356, p, 2643-2653.
24, p. 135-145. Wegmann, C. E., 1932, Note sur le Boudinage: Bulletin Société Géologique de France, series 5, II, p. 477-489.
Teufel. L. W., and Logan, J. M., 1978, Effect of displacement rate on the real area of contact and temperatures generated Wegmann, C. E., and Schaer, J. P., 1957, Lunules tectoniques et traces de mouvements dans les plis du Jura: Eclogae
during frictional sliding of Tennessee Sandstone: Pure and Applied Geophysics, v. 116, p. 840-865, Geologicae Helvetiae, v. 50, p. 492-496.
Thomson, A., 1959, Pressure solution and porosity, in Ireland, H. A., ed., Silica in sediments—A symposium with Wellman, H. W., 1962, A graphical method for analysing fossil distortion caused by tectonic deformation: Geological
discussions: Society of Economic Paleontologists and Mineralogists Special Publication No. 7, p. 92-109. Magazine, v. 49, p. 348-353.
1973, Soft-sediment faults in the Tesnus Formation and their relationship to paleoslope: Journal of Sedimentary Wenk, H. R., 1978, Are pseudotachylites products of fracture or fusion: Geology, v. 6, p. 507-511.
Petrology, v. 43, p. 525-528. Wenk, H. R., Venkitasubramanyan, C. S., and Baker, D. W., 1973, Preferred orientation in experimentally deformed
Tissot, B. P., and Welte, D. H., 1978, Petroleum formation and occurrence: New York, Springer-Verlag, 538 p. limestone: Contributions to Mineralogy and Petrology, v. 38, p. 81-114.
Tj'ia, H. D., 1964, Slickensides and fault movements: Geological Society of America Bulletin, v. 75, p. 683-686. Weyl, P. K., 1959, Pressure solution and the force of recrystallization—A phenomenological theory: Journal of Geophysi-
— — 1 9 7 2 , Fault movement, reoriented stress field and subsidiary structures: Pacific Geology, v. 5, p. 49-70. cal Research, v. 64, p. 2001-2025.
Tobin, D. G., and Donath, F. A., 1971, Microscopic criteria for defining deformational modes in rock: Geological Weymouth, J. H., and Williamson, W, O., 1953, The effects of extrusion and some other processes on the microstructure
Society of America Bulletin, v. 82, p. 1463-1476. of clay: American Journal of Science, v. 251, p. 89-108.
Trurnit, P., 1968, Pressure solution phenomena in detrital rocks: Sedimentary Geology, v. 2, p. 8 9 - I I 4 . Whisonant, R. C., 1970, Influence of texture upon the response of detrital quartz to deformation of sandstones: Journal of
Truswell, J. F., 1972, Sandstone sheets and related intrusions from Coffee Bay, Transkei, South Africa: Journal of Sedimentary Petrology, v. 40, p. 1018-1025.
Sedimentary Petrology, v. 42, p. 578-583. White, J. C., and White, S. H., 1983, Semi-brittle deformation within the Alpine fault zone, New Zealand: Journal of
Tullis, J. A., 1970, Preferred orientation in rocks produced by Dauphin^ twinning: Science, v. 168, p. 1342-1344. Structural Geology, v. 5, p. 579-589.
Tullis, J. A., and Tullis, T. E., 1972, Preferred orientation of quartz produced by mechanical Dauphinfc twinning: White, J. W., 1973, An invariant description of failure for an isotropic medium: Journal of Geophysical Research, v. 78,
American Geophysical Union Geophysical Monograph Series, v. 16, p. 67-82. p. 2438-2441.
Tullis, J. A., and Yund, R. A., 1985, Dynamic recrystallization of feldspar: A mechanism for ductile shear zone White, S., 1973a, Deformation lamellae in naturally deformed quartz: Nature Physical Science, v. 245, p, 26-28.
formation: Geology, v. 13, p. 238-241. 1973b, Syntectonic recrystallization and texture development in quartz: Nature, v. 244, p. 276-278.
1987, Transition from cataclastic flow to dislocation creep of feldspar: Mechanisms and microstructures: Geology, 1973c, The dislocation structures responsible for the optical effects in some naturally deformed quartzites: Journal
v. 15, p. 606-609. of Material Science, v. 8, p. 490-499.
Tullis, J. A., Christie, J. M., and Griggs, D. T., 1973, Microstructures and preferred orientations of experimentally 1976, The effects of strain on the microstructures, fabrics, and deformation mechanisms in quartzites: Royal
deformed quartzites: Geological Society of America Bulletin, v. 84, p. 297-314. Society of London Philosophical Transactions, ser. A, v. 283, p. 69-86.
Tullis, J. A., Snoke, A. W., and Todd, V. R., 1982, Penrose conference report: Significance and petrogenesis of mylonitic 1982, Fault rocks of the Moine thrust zone: A guide to their nomenclature-. Texture and Miciostructuies, v. 4,
rocks: Geology, v. 10, p. 227-230. p. 211-221.
Tullis, T. E., 1976, Experiments on the origin of slaty cleavage and schistosity: Geological Society of America Bulletin, White, S. H., 1979, Difficulties associated with paleostress estimates: Bullétin Mineralogique, v. 102, p. 210-215.
v. 87, p. 745-753. White, S. H., and Knipe, R. J., 1978, Microstructure and cleavage development in selected slates: Contributions to
1980, The use of mechanical twinning in minerals as a measure of shear stress magnitudes: Journal of Geophysical Mineralogy and Petrology, v. 66, p. 165-174.
Research, v. 85, p. 6263-6268. Wickham, J. S., 1973, An estimate of strain increments in a naturally deformed carbonate rock: American Journal of
Turner, F. J., 1953, Nature and dynamic interpretation of deformation lamellae in calcite of three marbles: American Science, v. 273, p. 23-47.
Journal of Science, v, 251, p. 276-298. Wickham, J. S., and Anthony, M., 1977, Strain paths and folding of carbonate rocks near Blue Ridge, central Appala-
Turner, F. J., and Heard, H. C., 1965, Deformation of calcite crystals at different strain rates: University of California, Los chians: Geological Society of America Bulletin, v. 88, p. 920-924.
Angeles, Geological Science Publications, v. 46, p. 103-126. Williams, P. F., 1972, Development of metamorphic layering and cleavage in low grade metamorphic rocks at Bermagui,
Turner, F. J., and Weiss, L. E., 1963, Structural analysis of metamorphic tectonites: New York, McGraw-Hill Book Co., Australia: American Journal of Science, v, 272, p. 1-47.
545 p. Williams, P. F., Collins, A. R., and Wiltshire, R. G., 1969, Cleavage and penecontemporaneous deformation structures in
1965, Deformation kinks in brucite and gypsum: National Academy of Science Proceedings, v. 54, p. 359-364. sedimentary rocks: Journal of Geology, v. 77, p. 415-425.
Turner, F. J., Griggs, D. T., and Heard, H. C., 1954a, Experimental deformation of calcite crystals: Geological Society of Willis, B., 1923, Geologic structures: New York, McGraw-Hill Book Company, Inc., 295 p.
America Bulletin, v. 65, p. 883-934. Wilson, L. R., 1971, Palynological techniques in deep-basin stratigraphy: Shale Shaker, v. 21, p. 124-139.
Turner, F. J„ Griggs, D. T., Heard, H. C., and Weiss, L. E., 1954b, Plastic deformation of dolomite rock at 380 °C: Wilson, T. V., and Sibley, D. F., 1978, Pressure solution and porosity reduction in shallow buried quartz arenite:
American Journal of Science, v. 252, p. 477-488. American Association of Petroleum Geologists Bulletin, v. 62, p. 2329-2334.
Tuttle, O. F., 1949, Structural petrology of planes of liquid inclusions: Journal of Geology, v. 57, p. 331-356. Wiltschko, D. V., and Sutton, S. J., 1982, Deformation by overburden of a coarse quartzite conglomerate: Journal of
Twiss, R. J., 1974, Structure and significance of planar deformation features in synthetic quartz: Geology, v, 2, Geology, v. 90, p. 725-733.
p. 329-332. Winslow, M. A., 1983, Clastic dike swarms and the structural evolution of the foreland fold and thrust belt of the southern
1976, Some planar deformation features, slip systems, and submicroscopic structures in synthetic quartz: Journal Andes: Geological Society of America Bulletin, v. 94, p. 1073-1080.
of Geology, v. 84, p. 701-724. Wintsch, R. P., 1978, A chemical approach to the preferred orientation of mica: Geological Society of America Bulletin,
—1977, Theory and applicability of a recrystallized grain size palaeopiezometer: Pure and Applied Geophysics, v. 89, p. 1715-1718.
v. 115, p. 227-244. Wise, D. U., 1964, Microjointing in basement, middle Rocky Mountains of Montana and Wyoming: Geological Society
Tyler, J. H., 1975, Fracture and rotation of brittle clasts in a ductile matrix: Journal of Geology, v. 83, p. 501-510. of America Bulletin, v. 75, p. 287-306,
Underbill, J. R., and Woodcock, N. H., 1987, Faulting mechanisms in high-porosity sandstones; New Red Sandstone, Wise, D. U., Dunn, D. E„ Engelder, J. T., Geiser, P. A., Hatcher, R. D., Kish, S. A., Odom, A. L., and Schamel, S„ 1984,
Arran, Scotland, in Jones, M. E., and Preston, R.M.F., eds., Deformation of sediments and sedimentary rocks, Fault-related rocks: Suggestions for terminology: Geology, v. 12, p. 391-394.
Geological Society Special Publication No. 29: Oxford, England, Blackwell Scientific Publications, p. 91-105. 1985a, Reply to Mawer on "Fault-related rocks: Suggestions for terminology": Geology, v. 13, p. 379.
Urai, J. L., 1985, Water-enhanced dynamic recrystallization and solution transfer in experimentally deformed carnalite: 1985b, Reply to Raymond on "Fault-related rocks: Suggestions for terminology": Geology, v. 13, p. 218-219.
Tectonophysics, v. 120, p. 285-317. Wojtal, S., and Mitra, G., 1986, Strain hardening and strain softening in fault zones from foreland thrusts: Geological
Van Hise, C. R., 1898, Metamorphism of rocks and rock flowage: Geological Society of America Bulletin, v. 9, Society of America Bulletin, v. 97, p. 674-687.
p. 269-328. Wong, P. D., and Oldershaw, A., 1981, Burial cementation in the Devonian, Kaybob Reef Complex, Alberta, Canada:
Van Loon, A. J., Brodzikowski, K„ and Gotowala, R., 1984, Structural analysis of kink bands in unconsolidated sands: Journal of Sedimentary Petrology, v. 51, p. 507-520.
Tectonophysics, v. 104, p. 351-374. Wood, D. S., and Holm, P. E., 1980, Quantitative analysis of strain heterogeneity as a function of temperature and strain
van Werveke, L., 1883, Eigenthiimliche Zwilligsbildung an Feldspat und Diallag: Neues Jahrbuch, v. 2, p. 97-101. rate: Tectonophysics, v. 66, p. 1-14.
Vance, J. A., 1961, Polysynthetic twinning in plagioclase: American Mineralogist, v. 46, p, 1097-1119. Woodcock, N. H., 1976, Structural style in slump sheets: Ludlow Series, Powys, Wales: Geological Society of London
Venkitasubramanyan, C. S., 1971, Qualitative analysis of three-dimensional strain using twins in calcite and dolomite: Journal, v. 132, p. 399-415,
Tectonophysics, v. 11, p. 217-231. Woodland, B. G., 1982, Gradational development of domainal slaty cleavage, its origin and relation to chlorite porphyro-
Vintanage, P. W., 1954, Sandstone dikes in the South Platte area, Colorado: Journal of Geology, v. 62, p. 493-500. blasts in the Martinsburg Formation, eastern Pennsylvania: Tectonophysics, v. 82, p. 89-124.
Voll, G., 1960, New work on petrofabrics: Liverpool and Manchester Geological Journal, v. 2, p. 503-567. Woodworth, J. B., 1896, On the fracture system of joints, with remarks on certain great fractures; Boston Society of
von Kirm&n, T., 1911, Festigheitversuche unter allseittgem Druck: Zeitschrift des Vereines Deutsche Ingenieure, v. 55, Natural History Proceedings, v. 27, p. 163-184.
p. 1749-1757. Wright, T. O., and Piatt, L. B., 1982, Pressure dissolution and cleavage in the Martinsburg Shale: American Journal of
Wallace, R. E., 1951, Geometry of shearing stress and relation to faulting: Journal of Geology, v. 59, p. 118-130. Science, v. 282. p. 122-135.
Wallace, R. E„ and Morris, H. T., 1986, Characteristics of faults and shear zones in deep mines: Pure and Applied Yagishita, K., and Morris, R. C., 1979, Microfabrics of a recumbent fold in cross-bedded sandstones: Geological Maga-
Geophysics, v. 124, p. 107-126. zine, v. 116, p. 105-116.
Wang, H. F., and Simmons, G., 1978, Microcracks in crystalline rock from 5.3-km depth in the Michigan Basin: Journal
of Geophysical Research, v. 83, p. 5849-5856.
Wanless, H. R., 1979, Limestone response to stress: Pressure solution and dolomitization: Journal of Sedimentary Petrol-
ogy, v. 49, p. 437-462. M A N U S C R I P T R E C E I V E D BY THE SOCIETY NOVEMBER 9 , 1 9 8 7
1982, Limestone response to stress: Pressure solution and dolomitization—Reply: Journal of Sedimentary Petrol- REVISED M A N U S C R I P T R E C E I V E D A P R I L 20,1988

ogy, v. 52, p. 328-332. MANUSCRIPT ACCEPTED APRIL 2 1 , 1 9 8 8

Printed in U.S.A.

Vous aimerez peut-être aussi