You are on page 1of 16

Eur. J. Mineral.

2015, 27, 471–486


Published online 13 March 2015

Petrophysical properties of a granite-protomylonite-ultramylonite sequence:


insight from the Monte Grighini shear zone, central Sardinia, Italy
STEFANO COLUMBU1,*, GABRIELE CRUCIANI1, DARIO FANCELLO1, MARCELLO FRANCESCHELLI1
and GIOVANNI MUSUMECI2

1
Dipartimento di Scienze Chimiche e Geologiche, Università di Cagliari, Via Trentino 51, 09127 Cagliari, Italy
*Corresponding author, e-mail: columbus@unica.it
2
Dipartimento di Scienze della Terra, Università di Pisa, Via S. Maria 53, 56126 Pisa, Italy

Abstract: The Monte Grighini intrusive complex (central-west Sardinia) has been affected by a 1.5 km wide late-Variscan shear
zone; microstructures and deformation fabrics indicate that high strain domains resulted in sharp transitions among protomylonite to
ultramylonite. A representative geological cross section of the shear zone has been sampled, as well as samples from outside the
mylonitic belt for use as slightly deformed protolith reference samples. Physical and mechanical analyses were performed on each
group in order to evaluate a possible correlation between petrophysical properties (density, porosity and mechanical strength) and the
degree of mylonitic deformation. Ultramylonites are generally characterized by a higher solid-phase density (2.82–2.85 g/cm3) than
less deformed protomylonites and mylonites (2.73–2.79 g/cm3), testifying to the formation of high-density minerals with an increase
in deformation. Conversely, bulk density tends to decrease as deformation increases (from 2.52–2.57 g/cm3 to 2.47–2.48 g/cm3) due
to the total porosity increase occurring during mylonitization. The mean values of total porosity range from 6.33 to 9.27 % in
protomylonites and mylonites and from 12.01 to 12.85 % in ultramylonites. The relatively high value of porosity in more deformed
ultramylonites is the result of the sum of closed micropores and open pores produced by different processes. Mechanical strength
measurements show a noticeable anisotropy, with higher strength values measured applying the load perpendicular to foliation
(Z axis) and lower values obtained parallel to it (X,Y axes). In the Monte Grighini mylonite the anisotropy is due to the distribution of
pores and to the preferred orientation of new minerals, as highlighted by the correlation between scalar physical properties (solid, real
and bulk densities, porosity) and vector properties (resistance to punching or uniaxial compressive strengths). The results show that
the degree of deformation of mylonitic rocks can be characterized by petrophysical properties, in order to shed light on the tectonic
processes experienced at depth.
Key-words: petrophysical properties; mylonitic granitoids; porosity; density; mechanical strength; anisotropy.

1. Introduction controls the rheology of shear zones. Furthermore, strain


softening processes associated with ductile flow (White
Geological and geophysical evidence shows that deforma- et al., 1980) allow the transition from mylonite to ultra-
tion in the Earth’s crust is heterogeneous, with large strain mylonite leading the shear zones to become rock layers
and displacements localized in faults and shear zones, with an intrinsic lower strength than the undeformed rock
which are also efficient pathways for magma and/or fluid outside the shear zone. The mylonite to ultramylonite
transfer from the lower to the middle-upper crustal levels transition has been the subject of several papers aimed at
(Hollister & Crawford, 1986; Scholz, 1988). understanding the microstructural evolution, behavior and
Strain localization depends on several intrinsic (grain rheology of these rocks. Many aspects have been investi-
size, mineralogy, phase transformations) and extrinsic gated, from geochemical and minerochemical features
(pressure, temperature, strain rate, fluid pressure) factors (Hippertt, 1998; Wibberley & McCaig, 2000) to the struc-
(Scholz, 1988; Shimamoto, 1989; Tullis & Yund, 1992; tural analysis (Hippertt & Hongn, 1998; Ree et al., 2005).
Wintsch et al., 1995) that determine which deformation Petrophysical properties have also been determined with a
mechanisms accommodate the fault or shear zone activity focus on seismic wave velocities (Siegesmund et al., 1991;
(Handy, 1989). According to the shear zone models by Cirrincione et al., 2010) and their relation to rock textures.
Scholz (1988) and Chester (1995), in the lower-middle In this paper a different approach, based on the determina-
crust below the brittle–ductile transition, the ductile flow tion of physical and mechanical properties, is used. For this
of quartz and feldspar is the dominant mechanism that purpose we analyzed the mylonitic granitoids of the Monte

eschweizerbart_xxx

0935-1221/15/0027-2447 $ 7.20
DOI: 10.1127/ejm/2015/0027-2447 # 2015 E. Schweizerbart’sche Verlagsbuchhandlung, D-70176 Stuttgart
472 S. Columbu et al.

Grighini shear zone in the Variscan belt of Sardinia (Elter External Zone, the External Nappe Zone, the Internal
et al., 1990; Musumeci, 1992). Nappe Zone and the Inner Zone. The Inner Zone is the
The Monte Grighini shear zone is the exhumed deep deepest part of the Variscan in Sardinia; it is composed of
portion of a kilometer-wide, late Variscan, transtensive several types of migmatites (Cruciani et al., 2001, 2008) and
shear zone exploited by synkinematic intrusions that can metabasite with eclogite and granulite facies relics
be regarded as a proxy of the mid-crustal portion of active (Franceschelli et al., 2002; Cruciani et al., 2012). The
shear zones in the continental crust. In particular, the occur- Monte Grighini Complex is a NW-SE elongated dome-
rence of a wide belt of ultramylonites in which most of the shaped metamorphic complex of Variscan age (Musumeci
strain and displacement was partitioned and the complete et al., 2015), exposed in central Sardinia (Fig. 1a), in the
exposure of the transition from protomylonitic to mylonitic External Nappe Zone (Carmignani et al., 1994).
to ultramylonitic fabrics, makes this structure a suitable The Monte Grighini Complex consists of a pile of
example to investigate the variation of the physical proper- Variscan metamorphic units (from bottom to top: the Monte
ties across a ductile shear zone resulting from strain locali- Grighini Unit, the Castello Medusa Unit and the Gerrei Unit),
zation processes. Furthermore, the igneous paragenesis derived from the middle Ordovician to Devonian sedimen-
(quartz, K-feldspar, plagioclase with a minor amount of tary succession. In particular the Monte Grighini Unit con-
biotite and muscovite) allows a better interpretation of the sists of Late Ordovician metavolcanic-volcanoclastic rocks
petrophysical data and makes these rocks representative of (Cruciani et al., 2013) overlain by a sequence of metapelites
the average composition of the continental crust (i.e., quart- with metasiltites and marble lenses at the top. These meta-
zofeldspathic crust). Several hand specimens of protomylo- morphic rocks were intruded by Late Carboniferous intrusive
nites to ultramylonites were collected in the field. For each rocks organized in two NW-SE striking belts that consist of
sample, solid, real and bulk densities, total, closed and open I-type metaluminous monzogranite with minor tonalite and
porosity to helium and to water, water absorption coeffi- small diorite bodies and S-type peraluminous muscovite-
cient, and the mechanical strength measured with uniaxial bearing leucogranite. The monzogranite is composed of
compressive and point load tests were determined. Bulk, quartz, K-feldspar, plagioclase and biotite, while the mineral
real and solid densities provide different information on the assemblage of the leucogranite consists of quartz, K-feldspar,
petrophysical features of the rock samples. Solid density is a plagioclase, muscovite, biotite, tourmaline  garnet. Rb/Sr
function of rock mineralogy, and should show the modal and Ar/Ar radiometric dating give ages of about 305–295 Ma
compositional variations linked to the mylonitization pro- for both metaluminous and peraluminous intrusive rocks
cess. Real and bulk densities are influenced by closed and (Carmignani et al., 1987; Del Moro et al., 1991). The units
total porosity, respectively, that are related to the textural experienced a polyphase tectonic and metamorphic evolu-
and microstructural aspects of the rock (e.g., foliation devel- tion. Early deformation related to syn-collisional nappe stack-
opment). The mechanical strength depends on the scalar ing developed southwest verging folds and axial planar
physical properties (i.e., total porosity, bulk and solid den- cleavage and/or schistosity, which represents the main folia-
sities) and textural, microstructural and mineralogical fea- tion in the field. These structures developed under different
tures (e.g., foliation orientation, grain size, mineral metamorphic conditions in the three units, and range from
hardness). The mechanical properties, measured in three lower greenschist facies in the Gerrei Unit to upper greens-
mutually perpendicular directions in a shear zone reference chist facies and amphibolite facies in the Castello Medusa
frame, can make a valuable contribution to study the aniso- Unit and Monte Grighini Unit, respectively. For the Monte
tropy ratios between the various levels of shear zones and Grighini Unit, geothermobarometric data indicate peak meta-
indirectly understand the mechanical strains occurring at morphic conditions of 7.5 kbar and 500 C followed by a
depth in the deformation process. nearly isothermal decompression up to 4 kbar (Spano et al.,
This paper, in addition to a basic mineralogical, petro- 2012). After nappe stacking and the metamorphic peak, late
graphic and physical characterization of the mylonitic and deformation in the Monte Grighini complex consisted of
ultramylonitic rocks, highlights the variations of the phy- NW-trending upright folds with development of large-scale
sical and mechanical properties at various stages of mylo- antiforms and synforms, followed by development of a wide
nite formation. The aim of this work is to provide a new strike-slip shear zone with synkinematic emplacement of
insight on the petrophysical properties of the exhumed intrusive rocks and final extensional structures (see below).
mylonitic rocks in order to better understand the deforma-
tional-metamorphic processes that they underwent at
depth.
2.2. Late Variscan Monte Grighini shear zone

The Monte Grighini shear zone (Fig. 1a and b) is an


2. Geological setting approximately 12 km long and 1.5 km wide ductile shear
zone that affected the Monte Grighini Unit and the intru-
2.1. Overview of the Monte Grighini Complex sive complex; the progressive transition from magmatic to
solid-state mylonitic fabrics indicates that the intrusions
The Variscan chain of Sardinia consists of four tectono- were synkinematic (Elter et al., 1990; Musumeci, 1992).
metamorphic zones that, from south to north, are: the Within the shear zone (Fig. 1a and b) the distribution of
eschweizerbart_xxx
Petrophysical properties of granite–ultramylonite sequence 473

Fig. 1. (a) Geological sketch map of the Monte Grighini Complex, central Sardinia (modified from Musumeci et al., 2015); (b) deformation
map of the shear zone showing the distribution of protomylonites, mylonites and ultramylonites (modified from Musumeci, 1992). The inset
shows a tectonic sketch map of the Variscan basement of Sardinia (modified from Carmignani et al., 2001); PAL: Posada-Asinara Line.
(online version in colour)

deformation fabrics is consistent with an increase of defor- shear bands that dip steeply toward SW. The shear bands
mation from a protomylonite zone to the east through a overprint the previous magmatic and metamorphic folia-
wide mylonite zone to a narrow ultramylonite zone to the tions and are the main fabric observed in the field. SC-type
west. The ultramylonite zone is a narrow belt (50–100 m in shear bands (Passchier & Trouw, 2005) consists of shear
width) that consists of ultramylonitic and phyllonitic rocks. plane (C-planes) striking N 150 –160 E while the foliation
Westward it is bounded by southwestward dipping detach- (S planes) strikes N120 E in the protomylonite zone and
ment faults and a wide cataclasite zone that marks the progressively rotate to become parallel to the C-planes.
westernmost side of the shear zone and the tectonic contact The complete parallelism between shear planes and
with the overlying, low grade Gerrei Unit (Fig. 1a). S-planes is reached in the ultramylonite zone that consists
Retrograde deformation, from lower amphibolite to of homogeneously foliated, fine-grained ultramylonites
lower greenschist facies, is almost entirely partitioned in and phyllonites, with N150 –160 striking foliation dip-
the intrusive rocks, which are the dominant lithology in the ping steeply (65 –75 ) toward WSW. The stretching linea-
shear zone. Medium-grade metamorphic rocks of the tions (marked by feldspar porphyroclasts, elongated quartz
Monte Grighini unit show a weak mylonitic overprint and mica trails) and slickenside striae trend NW-SE and
defined by thin (centimeter to decimeter thick) shear gently plunge (5 –20 ) toward NW in both the mylonitic
bands cross-cutting earlier foliations. In the intrusive com- and ultramylonitic foliation. According to Musumeci
plex poorly deformed igneous fabrics (WNW-striking, (1992), a prevailing simple shear deformation charac-
steeply dipping magmatic foliations) are preserved in the terizes the mylonite and ultramylonite zone with the high
protomylonite zone along the northeastern side of the shear strain values (0.8 , K , 1.1) reached in the ultramylonite
zone (Fig. 1b); the shear zone deformation increases zone, which is also marked by the highest values of shear
toward the west and southwest with the development of strain (g ¼ 20) and displacement (Musumeci, 1991). All
mylonitic fabrics that consist of NW-striking SC-type kinematic indicators in the mylonitic and ultramylonitic
eschweizerbart_xxx
474 S. Columbu et al.

rocks—such as SC-type shear bands, asymmetric mica fish

Table 1. Methodologies, physical properties and specimen characteristics of analyzed rock samples. The average volume of specimens is calculated for three groups: protolith samples (G70,

Ultra-mylonite

5.96  2.93
2.74  1.01
2.10  0.04
2.74  1.01
2.74  1.01
2.10  0.04
and sigma-type porphyroclasts—are consistent with a dex-
tral sense of shear. In the cataclasite zone, southwest gently
to moderately dipping foliation and brittle shear planes
with down-dip to slightly oblique slickenlines point out a

Volume average  standard deviation


top-to-the SSW sense of shear. Thus the Monte Grighini

Protomyl./Mylonite
shear zone is an exhumed dextral transtensive shear zone
with a horizontal displacement of at least 7 km (Musumeci,

4.06  0.27
16.31  5.45
4.06  0.27
2.12  0.03
4.06  0.27
2.12  0.03
1991). Shear zone activity triggered the emplacement of
synkinematic intrusions and drove the amphibolite facies
exhumation of the Monte Grighini Unit from middle to
upper crustal levels. Brittle deformation recorded the final
coupling with the Gerrei Unit at a very shallow crustal
level. At the scale of the Variscan orogeny, the Monte

2.11  0.04
3.80  1.13
3.80  1.13
2.11  0.04
3.80  1.13
Protoliths
Grighini shear zone belongs to the network of Late

n.d.
Carboniferous strike-slip to extensional shear zones that
accommodate large scale crustal extension and triggered
exhumation of deep metamorphic rocks and emplacement
of large volume of intrusive rocks (Jegouzo, 1980; Burg

(6 for each lithotype)


et al., 1981; Echtler & Malavieille, 1990; Rey et al., 1992).

Total number
3. Materials and methods
Petrographic and microstructural analyses were car-

60
13
78
78
13
78
ried out on polished thin sections by optical micro- G92, G97), protomylonites and mylonites (A1, A2, A3, A7, A8, A9, A10), and ultramylonites (A4, A5, A6).

Specimen characteristics
scopy and scanning electron microscopy with an FEI

Cubic and undisturbed


Cubic and undisturbed
Cubic and undisturbed

Cubic and undisturbed


Quanta 200 SEM at the Centro Grandi Strumenti of
University of Cagliari. A Panalytical X’Pert diffract-

Type or shape
ometer was used to identify the minerals of very fine-
grained ultramylonite-phyllonite samples, where

Powdered
Powdered
phases were unrecognizable by optical nor by back
scattered electron (BSE) observation.
Physical and mechanical properties were determined, in
the laboratory of the University of Cagliari, on 78 pseudo- Bulk volume; dry, wet and hydrostatic mass
cubic rock specimens (6 for each rock layer or group;
Tables 1, S1; supplementary material available online on
the GSW website of the journal), with an average volume
between 3.80 and 4.06 cm3, variable edge length (with a, b,
Physical properties determined

Uniaxial compressive strength

c ¼ 15  4 mm), and on 13 powdered rock specimens, with


an average volume ranging between 2.10 and 2.12 cm3.
The variable size of the specimens is due to the attempt to
Resistance to puncture
Solid and real volume

exclude weakness surfaces that may affect the reliability of


the results.
The samples were dried at 105  5 C and the dry solid
mass (mD) was determined. The volume of solid phases
Dry mass

(VS) of powdered rock specimens (on 10 g and with particle


size less than 0.063 mm) and the real volume (VR, where
VR ¼ VS þ VC, and VC is the volume of pores closed to
helium) of the rock specimens were determined by a helium
Ultrapycnometer 1000 (Quantachrome Instruments). Then,
Water absorption, analytical

Digital compression testing


Test type and methodology

and hydrostatic balance

the wet solid mass (mW) of the samples was determined


after water absorption by immersion for ten days. Through a
Helium pycnometry

hydrostatic analytical balance, the bulk volume VB (with VB


Point load tester

¼ VS þ VO þ VC, where VO ¼ (VBVR) is the volume of


open pores to helium) is calculated as:

VB ¼ ½ðmW  mHY Þ=rw T X  100




eschweizerbart_xxx
Petrophysical properties of granite–ultramylonite sequence 475

where mHY is the hydrostatic mass of the wet specimen and compressive strength. The strength measurements were
rwTX is the water density at a temperature TX. Total por- acquired by using a Wizard Basic microprocessor digital
osity (T), open porosity to water and helium (O H2O; O readout unit. The compressive strength (RC), determined
He, respectively), closed porosity to water and helium (C with the load applied to the X-Y-Z sample axes, was
H2O; C He), bulk density (rB), real density (rR), solid calculated as:
density (rS) were computed as:
RC ¼ P=A

T ¼ ½ðVB  VS Þ=VB  100  where P is the breaking load and A is the area of the
specimen section perpendicular to the direction of the load.
O H2 O ¼ f½ðmW  mD Þ=rw TX =VB g 100 

O He ¼ ½ðVB  VR Þ=VB  100 


4. Petrographic and deformation features of the
sampled section
C H2 O ¼ T  O H2 O In the northwest portion of the shear zone (Fig. 1b), a
decameter thick belt of alternating protomylonitic, mylo-
C He ¼ T  O He nitic to ultramylonitic layers derived from a biotite-bearing
monzogranite crops out along the north side of the Riu
rS ¼ mD =VS ; rR ¼ mD =VR ; rB ¼ mD =VB s’Iscibi valley (Fig. 1). Sampling for this study was con-
ducted on a meter-wide, NW-SE oriented section, called
‘‘A series’’, that exposes the various deformation fabrics
The weight imbibition coefficient (ICW) and the saturation and strain domains of the shear zone. This section (Fig. 2)
index (SI) were computed as: consists of a decimeter-thick ultramylonite layer (samples
A4, A5, A6) that is enclosed between two protomylonitic
ICW ¼ ½ðmW  mD Þ=mD  100 
layers (samples A1, A2, A3 and samples A7, A8) and a
mylonite (samples A9, A10). The boundaries between the
different protomylonite, mylonite and ultramylonite
SI ¼ ðO H2 O=O HeÞ domains are sharp and parallel to the pervasive mylonitic
¼ f½ðmW  mD Þ=rw T X =VO g 100 

foliation that homogeneously affects the entire outcrop


with an average N160 –N170 strike and a 60 –65 west-
ward dip. Samples G92, G97, G70 (see Fig. 1 for sample
The punching strength index was determined with a Point Load location) contain undeformed to poorly deformed fabrics
Tester in compliance with the International Society for Rock for comparison with the mylonite section. These protolith
Mechanics (ISRM, 1972, 1985) on six cubic oriented rock reference samples underwent variable degrees of very low
specimens for each sample (Table 1). The load is exerted via to low-strain.
the application of a concentrated load with two opposing con-
ical punches. The load was applied along three main directions:
the X axis, parallel to foliation and to mineralogical lineation; 4.1. Protolith reference samples
the Y axis, perpendicular to mineralogical lineation and parallel
to foliation; the Z axis, perpendicular to foliation. Sample G92 (Fig. 1a and 3a) is an almost undeformed mon-
The distance (D) between the two points of contact zogranite, which can be considered the igneous protolith. It is a
punches on a specimen are between 11 and 20 mm with an coarse-grained, equigranular monzogranite with no noticeable
average of 16.1  0.91 mm. The geometric requirements of mylonitic imprint except for a few sporadic oriented mica
the ISRM recommendations (1985) relating to the size of the domains. Subeuhedral and undeformed K-feldspar pheno-
specimens (i.e., 0.3 , D/W , 1 and L . 0.5W, where W and crysts commonly show myrmekite microstructures and some-
2L are the width perpendicular to the direction of the load and times perthite veins. Plagioclase, ranging in composition
the length of the specimen, respectively) were observed. The between albite and oligoclase, is typically zoned (core An5,
resistance to puncturing (IS) was calculated as P/De2 where P rim An12). Anhedral quartz grains are undeformed. Biotite and
is the breaking load and De is the ‘‘equivalent diameter of the muscovite occur as individual phenocrysts or are clustered into
carrot’’ (ISRM, 1985), with De ¼ 4A/p and A ¼ W  D. The slightly oriented mica-rich trails.
index value is referred to a standard cylindrical specimen Sample G97 was collected from the northern part of the
with diameter D ¼ 50 mm for which IS has been corrected intrusion. It is a coarse-grained, well foliated, dark colored
with a shape coefficient (F) and calculated as: monzogranite in which feldspar phenocrysts are sur-
rounded by oriented biotite trails. Igneous fabric is still
ISð50Þ ¼ IS F ¼ IS ðDe =50Þ0:45
 
evident, but it is strongly affected by the first appearance of
mylonitic microstructures. Feldspars are commonly sur-
A testing machine (digital model 50-C54, Controls) with rounded by poorly oriented trails of phyllosilicates and
load capacity of 3000 kN, max working pressure of 700 bar by a fine-grained matrix, which consists mainly of recrys-
and power of 750 W was used to determine the uniaxial tallized quartz and minor K-feldspar grains.
eschweizerbart_xxx
476 S. Columbu et al.

Fig. 2. (a) Samples mylonites of the Monte Grighini shear zone; (b) detail of the protomylonite-ultramylonite boundary located between A3
and A4.

Sample G70 is a muscovite-bearing leucogranite col- porphyroclasts, sometimes with perthite exsolution, exhi-
lected in the mylonite zone close to the boundary with bit domino-type, shear band-type and sometimes mosaic-
the protomylonite zone. Mylonitic textures are well devel- type microstructures. Plagioclase porphyroclasts, ranging
oped with pervasive foliations and noticeable S-C struc- in composition between albite and oligoclase (An10-20), are
tures recognizable at both meso and microscopic scale. The commonly altered and fractured. Quartz occurs mostly as
leucogranite mineral assemblage is similar to that of the fine-grained polycrystalline lenses or ribbons marking the
monzogranite except for the absence of biotite. Muscovite foliation and, more rarely, as porphyroclasts with undulose
is found in two different generations; a magmatic musco- extinction. Phyllosilicates (mostly biotite with average
vite with XMg ¼ 0.55 and Si ¼ 6.02 a.p.f.u. consisting of XMg ¼ 0.20 and muscovite with XMg ¼ 0.38 and Si ¼
sub-millimetric porphyroclasts commonly fish-shaped 6.191 a.p.f.u.) occur as fine-grained crystals, clustering in
(Fig. 3b) and a mylonitic muscovite in well oriented micro- sub-millimetric trails parallel to foliation. The boundary
crystalline trails. Mica-fish tend to help define the S planes, with the ultramylonite domain is marked by reduced grain
whereas the microcrystalline trails follow the C planes. size of muscovite and quartz and development of well-
Although strongly fractured, K-feldspar and plagioclase oriented mica trails.
porphyroclasts retain their original shape, whereas quartz The second protomylonite domain (samples A7, A8;
is almost completely dynamically recrystallized to form Fig. 2), about 25 cm thick, is located between the
the microcrystalline matrix. ultramylonite and the mylonite layers. The microstruc-
tural features are very similar to those of the protomy-
lonitic domain previously described. Both plagioclase
4.2. Protomylonite domain and K-feldspar have experienced brittle deformation,
with formation of domino-type microstructures.
The first domain of protomylonite (samples A1, A2, A3; Phyllosilicates and quartz underwent grain-size reduc-
Fig. 2), about 50 cm thick, consists of foliated dark-colored tion with the formation of millimeter-thick elongated
rocks, characterized by coarse-grained (2–5 mm) K-feld- ribbons of recrystallized fine-grained material.
spar and quartz porphyroclasts embedded in a fine-grained
(20–200 mm) quartz and phyllosilicate-rich matrix
(Fig. 3c). A panoramic overview of these textures in sam- 4.3. Mylonite domain
ple A1, obtained combining different BSE images, is
shown in Fig. 4. This domain, about 20–30 cm thick, shows a texture
Samples A1, A2, and A3 were collected in order to marked by porphyroclasts hosted in a fine-grained matrix.
follow the gradual increase in deformation that charac- Two samples (A9, A10; Fig. 2) with variable grain size,
terizes this domain. Following the porphyroclasts vs. phyllosilicate content and with porphyroclasts/matrix ratio
matrix criteria for fault rock classification (Sibson, 1977; approximately between 40/60 and 30/70, were collected
Wise et al., 1984), samples are protomylonites (recrystal- from this domain. Deformation varies greatly from one
lized matrix 30–35 % vol. of rock). K-feldspar sample to the other but also between different
eschweizerbart_xxx
Petrophysical properties of granite–ultramylonite sequence 477

Fig. 3. Photomicrographs showing the microstructure of: (a) monzogranite protolith sample G92; (b) leucogranite protolith sample G70; (c)
protomylonite sample A1; (d) mylonite sample A10; (e), ultramylonite sample A5, and (f) phyllonite sample A4. Cross-polarized light.
Mineral abbreviations according to Fettes & Desmons (2007).

Fig. 4. Panoramic BSE image of protomylonite sample A1. K-feldspar occurs as porphyroclasts with domino-type structures and perthitic
exsolution, and quartz is found in recrystallized ribbons that define the foliation.
eschweizerbart_xxx
478 S. Columbu et al.

microdomains within the same sample. Following the por- biotite—are not recognizable by either optical microscope
phyroclasts/matrix criteria for fault rocks classification or SEM analysis). K-feldspar relics are missing in the phyl-
(Sibson, 1977; Wise et al., 1984) samples from this domain lonite. This fabric results from the combination of mechan-
can be classified as mylonites stricto sensu. ical fracturing and dissolution of feldspar porphyroclasts
These mylonites are characterized by discontinuous and the complete recrystallization of quartz and mica,
bands of strongly fractured K-feldspar porphyroclasts, which forms very fine-grained crystals in ribbons and trails.
forming domino-type and shear band type microstructures, Porphyroclast abundance (,10 %) indicates that samples
alternating with dynamically recrystallized quartz-rich from this domain are ultramylonites (Sibson, 1977; Wise
layers. The porphyroclast grain size (1–3 mm) is smaller et al., 1984).
than in the protomylonite. Dynamically recrystallized
quartz becomes more abundant, forming millimeter-thick
polycrystalline ribbons or lenses that wrap around coarser
feldspar porphyroclasts (Fig. 3d). Igneous biotite and mus-
5. Analytical results
covite tend to disappear completely and phyllosilicates
The following properties were determined: bulk, real and
recrystallize as intergrowths of fine-grained micas forming
solid density, open porosity to helium and water, closed
thin trails parallel to quartz ribbons. The boundary with the
porosity, water imbibition coefficient (expressed as
ultramylonite is sharp, whereas toward the protomylonite a
weight), and water open porous saturation index (Tables
gradual decrease of strain results in a progressive grain size
S1a, b). Point load and uniaxial compressive strengths are
increase and in the appearance of larger and more abundant
given in Tables S2a and b.
porphyroclasts.
Solid density depends only on the mineralogy and thus
expresses a weight average value of density between the
4.4. Ultramylonite domain various crystalline or amorphous phases present in the
rock. The real and bulk densities are influenced, not only
The ultramylonite domain is about 15 cm thick and is by the density of the solid, but also by the porosity (from
characterized by a strong, pervasive foliation marked by a closed porosity and the total porosity, respectively). The
fine alternation of millimeter-thick grey quartzofeldspathic kinetics of water absorption over time depends on the
and brown phyllosilicate-rich layers. Based on the amount porosity (porous shape, size) and on the porous tortuosity.
of phyllosilicates, this domain has been further subdivided The porous tortuosity depends on the geometry and degree
into ultramylonite (samples A5, A6) and phyllonite (sample of connectivity of the porous network, and it exerts a strong
A4). The ultramylonite consists of a fine-grained (10–30 influence on the rock permeability.
mm) recrystallized matrix in which alternating quartz-rich
and mica-rich sub-millimetric layers surround rare rounded 5.1. Density and porosity
K-feldspar porphyroclasts and quartz rich aggregates
(Fig. 3e). The phyllonite consists of a centimeter-thick Average solid density, calculated for six specimens from
layer characterized by a homogeneous, very fine-grained each sample, shows an increasing trend with shear strain
matrix (Fig. 3f) formed by phyllosilicates (mostly biotite) (Tables S1a, b; Fig. 5). Leucogranite G70 has a solid mean
with a minor amount of quartz, K-feldspar and plagioclase density of 2.68 g/cm3, while the poorly deformed monzo-
(the mineral assemblage of this sample was determined by granites (G92 and G97) show a solid density of 2.73 and
powder XRD analysis because mineral phases—except 2.72 g/cm3, respectively. In the protomylonite-mylonite

Fig. 5. Bulk density (rB), real density (rR) and solid density (rS) values comparing the leuco-monzogranite protolith samples (G70, G92,
G97) to the mylonite section shown in Fig. 2, where A1, A2, A3, A7 and A8 are protomylonites, A9 and A10 are mylonites and A4, A5, A6
are ultramylonites.
eschweizerbart_xxx
Petrophysical properties of granite–ultramylonite sequence 479

domains, the values vary between 2.73 g/cm3 (sample A7) and solid density. In fact, when the solid density increases
and 2.79 g/cm3 (sample A10), whereas in the ultramylonite (and simultaneously bulk density decreases), a total porosity
domain they vary between 2.82 g/cm3 (sample A6) and increase is expected. Total porosity ranges (in mean values)
2.85 g/cm3 (sample A5). from 3.41 to 4.41 % in the protolith granitoids, from 6.33 %
The real density (from 2.66 g/cm3 in protomylonites to to 9.27 % in protomylonites and mylonites and from 12.01
2.72 g/cm3 in ultramylonites) shows a similar trend, to 12.85 % in ultramylonites (Tables S1a, b). Open porosity
although less marked, to that observed for the solid density tends to increase with shear strain. Closed porosity also
(Fig. 5), due to the increase of closed porosity with the increases in higher strain domains. Helium open porosity
degree of deformation (see below). The bulk density shows varies from 2.19 to 3.91 % in granitoid protoliths, from 3.89
an opposite effect compared to that observed for real and % to 6.14 % in protomylonites and mylonites, and from 7.90
solid density (Fig. 5), with higher mean values (2.61–2.62 to 8.91 % in ultramylonites. Helium closed porosity varies
g/cm3) in slightly deformed monzogranite, intermediate from 0.51 to 1.54 % in granitoids, from 2.44 to 3.75 % in
mean values in protomylonite to mylonite samples protomylonites and mylonites, and from 3.82 to 4.94 % in
(2.52–2.57 g/cm3) and lower mean values in ultramylonite ultramylonites.
(2.47–2.48 g/cm3). Bulk density depends on solid density
and total porosity. Since solid density increases with the
degree of shear strain, a bulk density decrease must be due 5.2. Physical-mechanical properties
to a noticeable porosity increase.
Porosity data (Fig. 6) match with density data, showing a The physical-mechanical properties of the samples were
positive correlation with the increase of shear strain, in investigated using Point Load and compressive stress tests
agreement with the opposite behavior between bulk density (results in Tables S2a, b). Figure 7 shows the mean punching

Fig. 6. Open porosity to water (O H2O) and to helium (O He), total porosity (T) and closed porosity to helium (C He) of the same sample
sequence as shown in Fig. 2.

Fig. 7. Punching resistance index [IS (50)] values determined along three main load directions (Y, X, Z) of the same sample sequence as shown
in Fig. 2.
eschweizerbart_xxx
480 S. Columbu et al.

index [IS(50)] for each sample in three preferential directions: examined quartzofeldspathic rocks, this physical varia-
X, Y, Z (in a shear zone reference frame, see Section 3). The tion showing a positive correlation with degree of
maximum strength is found along the Z axis because other deformation is driven by K-feldspar (with density of
directions are affected by foliation planes and/or other dis- 2.56 g/cm3) breakdown leading to the formation of
continuities. In most of the samples, the lowest values are higher-density minerals (i.e., muscovite and quartz with
obtained by applying the load along the Y axis direction, 2.88 and 2.65 g/cm3, respectively). This is well recog-
whereas intermediate values are generally found by load nizable in the ultramylonite rocks where the strong
application along the X axis. Absolute strength values increase in muscovite content occurred at the expense
decrease in response to an increasing degree of mylonitiza- of K-feldspar according to well-known reaction soft-
tion (with exception of some formerly fractured or strongly ening mechanisms observed in mylonitic granitoids
altered specimens of sample A2 that showed higher porosity (Hemley & Jones, 1964; O’Hara, 1988; Hippert,
than other specimens). Taking into account that the point load 1998; Wibberley, 1999; Ree et al., 2005; Wintsch &
test is strongly affected by pre-existing micro-fractures that Yeh, 2013 and references therein). This reaction
influence absolute values, it is nevertheless interesting to implies a volume loss similar to that reported by
observe the relative values between different directions and Wintsch & Yeh (2011, 2013) who estimated the
their relations with physical features. The uniaxial compres- volume change (DV ¼ 16 %) for the reaction:
sive stress test, similar to the point load test, shows that the 3KAlSi3O8 þ 2Hþ ¼ KAl2(Si3Al)O10(OH)2 þ 6SiO2
maximum strength (Tables S2a, b) is generally measured þ 2Kþ. The relevance of solid density is that it
along the Z direction, perpendicular to foliation. Between depends only on mineral assemblage so, in absence of
the other two directions, the higher strength is detected alteration processes, as in this case, it does not change
along the X axis (i.e., along the stretching lineation). during exhumation.
The increase of the total porosity with shear strain is due
to the reduction in the number of feldspar porphyroclasts
that lead to the development of a more porous, homoge-
6. Discussion neous, fine-grained matrix. In the ultramylonite domain,
the porosity mainly consists of (i) fine intra-matrix pores
The variation in the intensity of deformation in the Monte (mainly closed) developed during the recrystallization of
Grighini shear zone resulted in different fabrics (protomy- the new microcrystalline phases (e.g., quartz, muscovite)
lonitic to ultramylonitic) and consequently in different phy- and (ii) pores (mainly open) along the foliation and as
sical features that have been examined in the mylonitic inter- to intracrystalline microcracks (microfractured
samples and in the protolith reference samples. These phy- clasts). The increase of porosity might in small part also
sical properties were measured at atmospheric conditions, be related to epigenetic alteration processes and/or physi-
so absolute values may be affected by exhumation process. cal-mechanical relaxation, the latter due to exhumation at a
On the other hand, the relative values of each parameter and shallow crustal level. The Monte Grighini rocks do not
their variation through the protolith-ultramylonite sequence show significant evidence of alteration, and analytical
can be considered meaningful because the whole sequence results (i.e. the decrease of the saturation index with
underwent the same conditions from depth to the surface. shear strain, Fig. 8; Tables S1a, b) suggest negligible
Here we discuss the reliability of each parameter and why effects of exhumation. If exhumation had significantly
these results are considered significant. We will show that affected these rocks, an increase of the saturation index
the correlation among different parameters gives results in the mylonites would be expected due to the increase of
consistent with deformation-metamorphic processes open porosity to water produced by dilatation of pores in
already described in literature. Our results from the Monte response to the decrease of lithostatic load. The observed
Grighini shear zone highlight the relationships between decrease of the saturation index suggests that open porosity
petrophysical properties and strain intensity in the transition is probably influenced by mylonitic processes that lead to
from protomylonite to ultramylonite case-study. the development of smaller pore radius and higher tortu-
Increasing deformation strongly reduced the grain size and osity, both hindering water absorption.
decreased the porphyroclast/matrix ratio from protomylonite The real and bulk densities are parameters linked to the
to ultramylonite at both large (through the shear zone) and variable incidence of solid density and porosity. The var-
small (across the studied section) scales. The decrease of the iation in real density is mainly influenced by the presence
porphyroclast/matrix ratio is mostly related to brittle fractur- of closed porosity (with correlation coefficient R2 ¼ 0.92
ing of the feldspar porphyroclasts and to the production of for the deformed samples A1–A10, Fig. 9a) and to a lesser
fine-grained quartz and white mica from feldspar breakdown extent by solid density. Furthermore, the real density is
reactions. Heterogeneous distribution of deformation led to useful to distinguish among the different sample popula-
alternation of protomylonites, mylonites and ultramylonites tions when plotted vs. total porosity (Fig. 9b). In this
with sharp boundaries between domains. diagram the mylonites (samples A9, A10) show physical
A marked increase of solid density (from 2.68 to behavior intermediate between those of the protomylonites
2.85 g/cm3), and subordinately of real density, were (samples A1, A2, A3, A7, A8) and ultramylonites (samples
observed from protomylonite to ultramylonite. In the A4, A5, A6).
eschweizerbart_xxx
Petrophysical properties of granite–ultramylonite sequence 481

A decrease of absolute values of mechanical strength


was observed especially when the load was applied parallel
to the foliation. We do not assume that absolute strength
values, measured at surface conditions, are representative
of the conditions at depth in the crust because the latter are
affected by other factors such as lithostatic confining pres-
sure, temperature, fluid pressure. Nevertheless, the com-
parison between strength indexes in different directions
and the relationship between these indexes and the other
scalar properties are reliable. The significant agreement
among the various determined parameters (see below)
supports our assumptions. An increase in the mylonitic
deformation commonly correlates with a decrease in
mechanical resistance (Fig. 7). This strength decrease
acquires greater relevance considering the fact that, in
undeformed igneous rocks, a grain size decrease involves
a compressive strength increase following a well-known
relation (from Brace, 1961 to Přikryl, 2001, and references
therein). In mylonitic rocks, the opposite behavior is
Fig. 8. Open porosity to helium (O He) vs open porosity to water observed where strength decreases with decreasing grain
(O H2O) of leucogranite (G70), monzogranite (G92, G97), proto- size. This is an expected behavior considering that mylo-
mylonite (A1, A2, A3, A7, A8), ultramylonite (A4, A5, A6) and nitization produces a pervasive homogeneous foliation and
mylonite samples (A9, A10). The line indicates water saturation a total porosity increase. The increase in total porosity has
index at 100 %.
an important effect on reducing mechanical strength as
stated by Tuğrul & Zarif (1999), who found a negative
Bulk density decreases in response to an increasing linear correlation between these parameters. Moreover, the
degree of deformation (Tables S1a, b; Fig. 10a, b) due to growth of new phyllosilicates (i.e., fine-grained musco-
the opposite effects of total porosity and solid density vite) in the presence of a circulating fluid phase, promotes
increases, where the total porosity has the stronger influ- ductile deformation (Wibberley, 1999), since phyllosili-
ence. Open porosity plays an important role in the devel- cates with a strong shape preferred orientation, defining
opment of mylonitic and ultramylonitic fabrics, enhancing the foliation, are characterized by low resistance. The
fluid circulation throughout the structure itself; in fact, correlation diagrams between total porosity and point
several authors (Hippertt, 1998; Ishii et al., 2007; load strengths [IS(50)] measured along the X, Y, Z axes
Bhattacharyya & Mitra, 2011, 2014; Fukuda & Okudaria, (Fig. 11a–c) show significant differences depending on the
2013) point out the role of the fluid phase in fault rheology direction of the applied load. This behavior is probably
because it can reduce the frictional coefficient and enable related to the higher breaking strength that minerals oppose
mica-forming softening reactions. if stressed along their long direction (X axis).

Fig. 9. (a) Closed porosity to water (C H2O) vs. real density (rR) diagram, of the A1–A10 mylonitic sequence and protolith samples (G70,
G92, G97). For each group, regression lines and correlation coefficients (R2) are reported. (b) Total porosity (T) vs. real density (rR)
diagram for the same samples.

eschweizerbart_xxx
482 S. Columbu et al.

Fig. 10. (a) Total porosity (T) vs. bulk density (rB) diagram and (b) solid density (rS) vs. bulk density (rB) of leucogranite (G70),
monzogranite (G92, G97 samples), protomylonite (A1, A2, A3, A7, A8), ultramylonite (A4, A5, A6) and mylonite samples (A9, A10). Both
diagrams show a negative correlation.

Fig. 11. Punching resistance index [IS (50)] determined along perpendicular axes (a) Y, (b) X, (c) Z and (d) XYZ vs. total porosity (T) from
protolith and mylonite zone samples, where XY is the foliation plane and X is the stretching direction.

The negative correlation between porosity and maxi- foliation planes (e.g., Santa Rosa mylonite, Kern &
mum strength along the X and Y axes (with a high correla- Wenk, 1990). There is no appreciable correlation (R2 ¼
tion coefficient: R2 ¼ 0.73 and 0.62, respectively; 0.20) along the Z axis, due to both a greater data dispersion
Fig. 11d) is consistent with the suggestion that pores are and a greater punching resistance of the ultramylonite
preferentially located between the phyllosilicate-rich group (influenced more probably by solid density rather
eschweizerbart_xxx
Petrophysical properties of granite–ultramylonite sequence 483

Fig. 12. Punching resistance index [IS (50)] determined along (a) Y, (b) X, (c) Z and (d) XYZ vs. bulk density (rB) from protolith and mylonite
zone samples, where XY is the foliation plane and X is the stretching direction.

than by total porosity). The punching index [IS(50)] is rocks. The strong control exerted by mechanical aniso-
positively correlated with bulk density (Fig. 12a–d); the tropy also fits very well with the anisotropy measurements
sample populations are clearly recognizable, with positive obtained by several authors using seismic wave velocity
correlation coefficients within each group slightly higher from Jones & Nur (1982) to Meltzer & Christensen (2001).
than those between the IS(50) and total porosity. Jones & Nur (1982) state that only little anisotropy can be
Uniaxial compressive strength values have been related detected in mylonitic rocks, unless a large amount of
to the main physical properties, obtaining similar results to phyllosilicate is present, as in our example. Seismic aniso-
those described for the point load test (not shown). Better tropy calculations performed using electron backscatter
correlation is shown together by protomylonites and mylo- diffraction (EBSD) data (Dempsey et al., 2011; Lloyd
nites, whereas ultramylonitic samples, especially along the et al., 2011) reveal that the main effect is produced by
Y axis, show inconsistent behavior. The breaking surfaces crystal preferred orientation of micas, and to a lesser extent
have revealed that when applying the load in the direction quartz, along the foliation planes. Lloyd et al. (2011) state
of the Y and X axes, the rupture always occurs parallel to that even a mica content as low as 10 % (in a quartzofelds-
foliation, dividing the samples into layers of constant pathic matrix) has a significant effect on anisotropy,
thickness. If the load is applied along the Z axis, the rupture whereas mica content higher than 20 % controls aniso-
occurs along random directions, except for those speci- tropy. The petrophysical properties also highlight the con-
mens that have pre-existing micro-fractures. tribution of porosity in controlling the velocity and
These results highlight the strong control exerted by direction of seismic waves. Oriented microcracks contri-
mechanical anisotropy (i.e., pervasive foliation) in mylo- bute largely to the directional dependence of elastic wave
nite and ultramylonite rocks. In fact, mylonitic foliations velocities, in addition to preferred orientation of minerals
are preferred planes of weakness with a high concentration (Kern & Wenk, 1990). This is particularly effective under
of pores, favoring fluid circulation along channeled path- conditions of low effective pressure in the rock volume
ways and strain partitioning with respect to the surrounding (high pore pressure) where most of the intercrystalline

eschweizerbart_xxx
484 S. Columbu et al.

cracks are open to fluid circulation. In the reported exam-  decrease of bulk density resulting from an increase of
ple a preferred orientation of microcracks can be deter- open and closed porosity hosted mainly among foliation
mined by the negative correlation between porosity and planes and subordinately in the fine-grained matrix;
maximum strength along the X and Y axes. This means  general decrease of uniaxial compressive and point-load
that, at depth, seismic anisotropy of ductile shear zones is strengths that show a negative correlation with the total
mostly controlled by preferred orientation of minerals porosity when the load is applied parallel to foliation (Y
(mostly micas) and interconnected intergranular micro- and X directions), where mechanical deformation and
cracks. Microcracks are critical to the localization of both breaking is favored; along the Z direction, the strengths
fluid flow and deformation in the development of shear are higher than along the Y and X axes, with a slightly
zones in particular under retrograde metamorphic condi- negative correlation between total porosity and degree
tions. The decrease of bulk density, as consequence of the of mylonitization.
increase of total porosity, is directly related to lower seis-
mic velocity of P-waves (Gardner et al., 1974). This means Mylonitization results in a drastic change of physical and
that a decrease of P-wave velocity could be expected to mechanical properties with, in particular, a significant
correspond to the transition from protomylonite to ultra- variation in the mechanical strength that allows strain and
mylonite, thus marking the position of high strained rocks displacement to be partitioned into the weak shear zones.
at depth in crustal-scale shear zones. Finally, the example studied as a proxy of active shear
In the Monte Grighini shear zone, the development of a zone at depth, also provides evidence that high strain
pervasive mylonitic foliation is marked by intense grain domains (ultramylonite zones) in a large-scale shear zone
size reduction and growth of new phyllosilicates with are marked by a drastic change in physical parameters with
strong shape preferred orientation. The phyllosilicates, respect to the surrounding less deformed rocks, allowing
organized in millimeter-thick continuous layers, form the them to be identified in seismic profiles.
dominant fabric in the ultramylonites. This feature might The petrophysical parameters determined in this study
suggest that the softening reactions are the most effective are in agreement with known deformational and meta-
mechanism in mylonite development. However, other morphic processes commonly occurring in quartzofelds-
mechanisms could affect the development of shear zones; pathic shear zones: bulk density increase is consistent with
for example Gilotti (1992) highlights the role of grain size softening reactions, for which an involvement of fluid
reduction in controlling the rheology even in absence (or phases is needed; porosity increase explains how fluid
with a minor contribution) of softening reactions. In the phases are driven through the shear zone; anisotropy of
Monte Grighini shear zone the decrease of grain size and mechanical strength is consistent with the oriented recrys-
the recrystallization of oriented micas-rich layers seems to tallization of phyllosilicates.
indicate that both mechanisms contributed to the develop-
ment of mylonitic and ultramylonitic fabrics and strongly
controls the physical and mechanical properties of rocks.
Considering that the composition of the examined intru-
Acknowledgements: The Authors wish to thanks two
sive rocks is representative of the upper continental crust, it
anonymous Reviewers for helpful criticism and sugges-
is clear that in quartzofeldspathic rocks the content of
tions. We are also grateful to Associate Editor Jane Gilotti
phyllosilicates and grains size are potentially the most
for editorial handling and improvements to the manuscript.
important factors in controlling both mechanical aniso-
This work benefited from Contributo di Ateneo per la
tropy and physical properties such as porosity and conse-
Ricerca (CAR) Funds, Università di Cagliari.
quently fluid circulation.

7. Conclusions References
The results from the Monte Grighini shear zone show a
correlation between petrophysical properties, fabric devel- Bhattacharyya, K. & Mitra, G. (2011): Strain softening along the
MCT zone from the Sikkim Himalaya: relative roles of Quartz
opment and degree of deformation in a protomylonite-
and Micas. J. Struct. Geol., 33, 1105–1121.
mylonite-ultramylonite sequence affecting quartzofelds-
—, — (2014): Spatial variations in deformation mechanisms along
pathic rocks representative of the continental crust. The
the Main Central thrust zone: implications for the evolution of
transition from a protomylonite to ultramylonite fabric the MCT in the Darjeeling-Sikkim Himalaya. J. Asian Earth
shows: Sci., 96, 132–147.
Brace, W.F. (1961): Dependence of fracture strength of rocks on
 marked increase of solid density, and subordinately of grain size. In: Proc. 4th Symp. Rock Mech., Univ. Park, Penn.,
real density. In quartzofeldspathic rocks, this physical PA, p 99–103.
variation is driven by K-feldspar breakdown leading to Burg, J.-P., Iglesias, M., Laurent, Ph., Matte, Ph., Ribeiro, A. (1981):
the formation of higher-density minerals (i.e., musco- Variscan intracontinental deformation: the Coimbra-Cordoba
vite and quartz), and to a volume loss; shear zone. Tectonophysics, 78, 161–177.
eschweizerbart_xxx
Petrophysical properties of granite–ultramylonite sequence 485

Carmignani, L., Cherchi, G.P., Del Moro, A., Franceschelli, M., Franceschelli, M., Carcangiu, G., Caredda, A.M., Cruciani, G.,
Ghezzo, C., Musumeci, G., Pertusati, P.C. (1987): The mylonitic Memmi, I., Zucca, M. (2002): Transformation of cumulate
granitoids and tectonic units of the Mount Grighini Complex mafic rocks to granulite and re-equilibration in amphibolite
(W-Sardinia): a preliminary note. IGCP project n 5: and greenschist facies in NE Sardinia, Italy. Lithos, 63, 1–18.
‘‘Correlation of Prevariscan and Variscan events of the Alpine- Fukuda, J.-I. & Okudaira, T. (2013): Grain-size-sensitive creep of
Mediterranean mountain belt’’. Newsletter, 7, 25–26. plagioclase accompanied by solution-precipitation and mass
Carmignani, L., Carosi, R., Di Pisa, A., Gattiglio, M., Musumeci, G., transfer under mid-crustal conditions. J. Struct. Geol., 51, 61–73.
Oggiano, G., Pertusati, P.C. (1994): The Hercynian chain in Gardner, G.H.F., Gardner, L.W., Gregory, A.R. (1974): Formation
Sardinia (Italy). Geodin. Acta, 7(1), 31–47. velocity and density—the diagnostic basics for stratigraphic
Carmignani, L., Oggiano, G., Barca, S., Conti, P., Salvadori, I., traps. Geophysics, 39, 770–780.
Eltrudis, A., Funedda, A., Pasci, S. (2001): Geologia della Gilotti, J.A. (1992): The rheologically critical matrix in arkosic mylo-
Sardegna. Note illustrative della Carta Geologica della nites along the Särv thrust, Swedish Caledonides. in ‘‘Structural
Sardegna a scala 1:200.000. Mem. Descr. Carta Geol. It., 60, geology of fold and thrust velts’’, S. Mitra & G. Fisher, eds. Johns
283 p. Hopkins University Press, Baltimore, 145–160.
Chester, F.M (1995): A rheologic model for wet crust applied to Handy, M.R. (1989): Deformation regimes and the rheological evo-
strike-slip faults. J. Geophys. Res., 100, 13033–13044. lution of fault zones in the lithosphere: the effects of pressure,
Cirrincione, R., Fazio, E., Heilbronner, R., Kern, H., Mengel, K., temperature, grain size and time. Tectonophysics, 163, 119–152.
Ortolano, G., Pezzino, A., Punturo, R. (2010): Microstructure Hemley, J.J. & Jones, W.R. (1964): Chemical aspects of hydrother-
and elastic anisotropy of naturally deformed leucogneiss from a mal alteration with emphasis on hydrogen metasomatism. Econ.
shear zone in Montalto (southern Calabria, Italy). Geol. Soc. Geol., 59, 538–569.
London, special publ., 332, 49–68. Hippertt, J.F. (1998): Breakdown of feldspar, volume gain and lateral
Cruciani, G., Franceschelli, M., Caredda, A.M., Carcangiu, G. mass transfer during mylonitization of granitoid in a low meta-
(2001): Anatexis in the Hercynian basement of NE Sardinia, morphic grade shear zone. J. Struct. Geol., 20(2/3), 175–193.
Italy: a case study of the migmatite of Porto Ottiolu. Mineral. Hippertt, J.F. & Hongn, F.D. (1998): Deformation mechanisms in
Petrol., 71, 195–223. the mylonite/ultramylonite transition. J. Struct. Geol., 20(2),
Cruciani, G., Franceschelli, M., Jung, S., Puxeddu, M., Utzeri, D. 1435–1448.
(2008): Amphibole-bearing migmatites from the Variscan Belt Hollister, L.S. & Crawford, M.L. (1986): Melt-enhanced deforma-
of NE Sardinia, Italy: partial melting of mid-Ordovician igneous tion: a major tectonic process. Geology, 14, 558–561.
sources. Lithos, 105, 208–224. International Society for Rock Mechanics (1972): Suggest method
Cruciani, G., Franceschelli, M., Groppo, C., Spano, M.E. (2012): for determining the point load strength index. ISRM (Lisbon,
Metamorphic evolution of non-equilibrated granulitized eclo- Portugal), Committee on Field Tests, Document n.1, p 8–12.
gite from Punta de li Tulchi (Variscan Sardinia) determined —— (1985): Suggest method for determining the point load
through texturally controlled thermodynamic modelling. J. strength. ISRM Commission for Testing Methods, Working
Metamorphic Geol., 30, 667–685. Group on Revision of the Point Load Test Methods. Int. J.
Cruciani, G., Franceschelli, M., Musumeci, G., Spano, M.E., Rock Mech. Min. Sci. Geomech. Abstr., 22, 51–60.
Tiepolo, M. (2013): U–Pb zircon dating and nature of metavol- Ishii, K., Kanagawa, K., Shigematsu, N., Okudaira, T. (2007): High
canics and metarkoses from the Monte Grighini Unit: new ductility of K-feldspar and development of granitic banded
insights on Late Ordovician magmatism in the Variscan belt in ultramylonite in the Ryoke metamorphic belt, SW Japan. J.
Sardinia, Italy. Int. J. Earth Sci., 102, 2077–2096. Struct. Geol., 29, 1083–1098.
Del Moro, A., Laurenzi, M., Musumeci, G., Pardini, G. (1991): Rb/ Jegouzo, P. (1980): The South Armorican shear zone. J. Struct.
Sr and Ar/Ar chronology of Hercynian Mt. Grighini intrusive Geol., 2, 39–47.
and metamorphic rocks, (central-western Sardinia. Plinius, 4, Jones, T.D. & Nur, A. (1982): Seismic velocity and anisotropy in
121–122. mylonites and the reflectivity of deep crustal fault zones.
Dempsey, E.D., Prior, D.J., Mariani, E., Toy, V.G., Tatham, D.J. Geology, 10, 260–263.
(2011): Mica-controlled anisotropy within mid-to-upper crustal Kern, H. & Wenk, H.R. (1990): Fabric-related velocity anisotropy
mylonites: an EBSD study of mica fabrics in the Alpine Fault and shear wave splitting in rocks from the Santa Rosa Mylonite
Zone, New Zealand. in ‘‘Deformation mechanisms, rheology Zone, California. J. Geophys. Res., 95(B7), 11213–11223.
and tectonics: microstructures, mechanics and anisotropy’’, Lloyd, G.E., Halliday, J.M., Butler, R.W.H., Casey, M., Kendall, J.-
D.J. Prior, E.H. Rutter, D.J. Tatham, eds. Geol. Soc., London, M., Wookey, J., Mainprice, D. (2011): From crystal to crustal:
Spec. Publ., 360, 33–47. petrofabric-derived seismic modeling of regional tectonics. in
Echtler, H. & Malavieille, J. (1990): Extensional tectonics, basement ‘‘Deformation mechanisms, rheology and tectonics: microstruc-
uplift and Stephano-Permian collapse basin in a late Variscan tures, mechanics and anisotropy’’, D.J. Prior, E.H. Rutter, D.J.
metamorphic core complex (Montagne Noire, Southern Massif Tatham, eds. Geol. Soc., London, Spec. Publ., 360, 49–78.
Central). Tectonophysics, 177, 125–138. Meltzer, A. & Christensen, N. (2001): Nanga Parbat crustal aniso-
Elter, F.M., Musumeci, G., Pertusati, P.C (1990): Late Hercynian tropy: implications for interpretation of crustal velocity structure
shear zones in Sardinia. Tectonophysics, 176(3/4), 387–404. and shear-wave splitting. Geophys. Res. Lett., 28, 2129–2132.
Fettes, D. & Desmons, J. (Eds.) (2007): Metamorphic Rocks: a Musumeci, G. (1991): Displacement calculation in a ductile shear
classification and glossary of terms. Recommendations of the zone: Monte Grighini shear zone (Central-Western Sardinia).
international union of geological science Subcommission on the Boll. Soc. Geol. It., 110, 771–777.
systematic of metamorphic rocks. Cambridge University Press, —— (1992): Ductile wrench tectonics and exhumation of Hercynian
Cambridge, UK, 244 p. metamorphic basement in Sardinia; Monte Grighini Complex. In:
eschweizerbart_xxx
486 S. Columbu et al.

Symposium ‘‘Palaeozoic orogenies in Europe: Tectonics, mag- selected granitic rocks from Turkey. Eng. Geol., 51,
matism and evolution.’’ Geodin. Acta, 5, 119–133. 303–317.
Musumeci, G., Spano, M.E., Cherchi, G.P., Franceschelli, M., Tullis, J. & Yund, R.A. (1992): The brittle-ductile transition in
Pertusati, P.C., Cruciani, G. (2015): Geological map of the feldspar aggregates: an experimental study. in ‘‘Fault mechanics
Monte Grighini Variscan Complex (Sardinia, Italy). J. Maps, and transport properties of rocks’’, B. Evans & T.-F. Wong, eds.
11(2), 287–298. Academic, San Diego, CA, 89–117.
O’Hara, K. (1988): Fluid flow and volume loss during mylonitisa- White, S.H., Burrows, S.E., Carreras, J., Shaw, N.D., Hvmes, F.J.
tion: an origin for phyllonite in an overthrust setting, North (1980): On mylonites in ductile shear zones. J. Struct. Geol., 2,
Carolina, U.S.A. Tectonophysics, 156, 21–36. 175–187.
Passchier, C.W. & Trouw, R.A.J. (2005): Microtectonics. Springer- Wibberley, C. (1999): Are feldspar-to-mica reactions necessarily
Verlag, Berlin, 366 p. reaction-softening processes in fault zones? J. Struct. Geol.,
Přikryl, R. (2001): Some microstructural aspects of strength varia- 21, 1219–1227.
tion in rocks. Int. J. Rock Mech. Min. Sci., 38, 671–682. Wibberley, C.A.J. & McCaig, A.M. (2000): Quantifying orthoclase
Ree, J.-H., Kim, H.S., Han, R., Jung, H. (2005): Grain-size reduction and albite muscovitisation sequences in fault zones. Chem.
of feldspars by fracturing and neocrystallization in a low-grade Geol., 165, 181–196.
granitic mylonite and its rheological effect. Tectonophysics, Wintsch, R.P., Yeh, M.-W (2011): Lower greenschist facies oscilla-
407, 227–237. tions across the ‘brittle-ductile’ transition induced by alternating
Rey, P., Burg, J.P., Caron, J.M. (1992): Middle and Late reaction softening and hardening. in ‘‘The interrelationship
Carboniferous extension in the Variscan belt: structural and between deformation and metamorphism’’, D.G.A.M. Aerden
petrological evidences from the Vosges massif (eastern & S.E. Johnson, eds. Abstracts Volume, University of Granada,
France). Geodin. Acta, 5, 17–36. Granada, 32–33.
Scholz, C.H. (1988): The brittle-plastic transition and the depth of —, — (2013): Oscillating brittle and viscous behavior through the
seismic faulting. Geol. Rund., 77, 319–328. earthquake cycle in the Red River Shear Zone: monitoring flips
Shimamoto, T. (1989): The origin of S-C mylonites and a new fault- between reaction and textural softening and hardening.
zone model. J. Struct. Geol., 11, 51–64. Tectonophysics, 587, 46–62.
Sibson, R.H. (1977): Fault rocks and fault mechanisms. J. Geol. Soc. Wintsch, R.P., Christoffersen, R., Kronenberg, A.K. (1995): Fluid-
London, 33, 191–213. rock reaction weakening of fault zones. J. Geophys. Res., 100,
Siegesmund, S., Kern, H., Vollbrecht, A. (1991): The effect of 13021–13032.
oriented microcracks on seismic velocities in an ultramylonite. Wise, D.U., Dunn, D.E., Engelder, J.T., Geiser, P.A., Hatcher,
Tectonophysics, 186, 241–251. R.D., Kish, S.A., Odom, A.L., Schamel, S. (1984): Fault-
Spano, M., Cruciani, G., Franceschelli, M., Massonne, H.-J., related rocks: suggestion for terminology. Geology, 12,
Musumeci, G. (2012): Variscan metamorphic evolution of the 391–394.
Monte Grighini Unit in central Sardinia. Géol. France, 1,
206–207. Received 4 July 2014
Tuğrul, A. & Zarif, I.H. (1999): Correlation of mineralogical Modified version received 25 February 2015
and textural characteristics with engineering properties of Accepted 27 February 2015

eschweizerbart_xxx

You might also like