You are on page 1of 13

Fluid conduits in carbonate-hosted seismogenic normal faults of

central Italy
F. Agosta
1
and D. L. Kirschner
Department of Earth and Atmospheric Sciences, Saint Louis University, Saint Louis, Missouri, USA
Received 5 June 2002; revised 13 November 2002; accepted 4 February 2003; published 29 April 2003.
[1] We studied the structures and stable isotope geochemistry of carbonate fault rocks in
four normal faults of central Italy. The faults juxtapose Meso-Cenozoic carbonates of
the footwalls against continental basins of the hanging walls. Footwall rocks exposed
along fault scarps have been exhumed from depths of 1 km. The fault rocks are
systematically arranged in each fault and can be separated into five distinct domains.
Farthest from the main fault contact are undeformed host rock and fractured host rock
(domain 1). Progressively closer to the fault contact, in the core of the fault, are gouge
(domain 2), cataclasite (domain 3), and cement-dominated horizons with planar slip
surfaces (domain 4). Thin horizons of brecciated hanging wall sediments (domain 5) are
adjacent, and locally accreted, to the footwall. Structural and stable isotope data are
consistent with compartmentalization of fluid in the faults during exhumation. The data
are most consistent with these fluids being predominantly evolved meteoric water rather
than fluids from the mantle, crustal magmas, and/or devolatilizing carbonate rocks.
Meteoric water infiltrated domains 4 and 5 of the faults, either from hanging wall
sediments or directly at the land surface. The gouge and cataclasites of domains 2 and 3
were impermeable barriers to movement of meteoric water into domain 1 and
undamaged host rocks of the footwall. The results of this study do not support models
of earthquake nucleation and rupture that envision large volumes of deep-seated fluids
passing upward through shallow portions of these seismogenic faults. INDEX TERMS:
8010 Structural Geology: Fractures and faults; 8045 Structural Geology: Role of fluids; 1040 Geochemistry:
Isotopic composition/chemistry; 1045 Geochemistry: Low-temperature geochemistry; 7299 Seismology:
General or miscellaneous; KEYWORDS: stable isotope, normal fault, fluid, architecture, earthquake, Italy
Citation: Agosta, F., and D. L. Kirschner, Fluid conduits in carbonate-hosted seismogenic normal faults of central Italy,
J. Geophys. Res., 108(B4), 2221, doi:10.1029/2002JB002013, 2003.
1. Introduction
[2] Many faults in the upper crust have a core of highly
deformed rocks surrounded by lesser deformed to unde-
formed rocks [Sibson, 1977; Chester and Logan, 1986;
Byerlee, 1993; Scholz and Anders, 1994; Caine et al.,
1996]. The fault core is a narrow, highly deformed zone
that forms around a fault surface(s) and is the site of
comminution, dissolution/precipitation, mineral reactions,
and other fault-related processes. A damage zone surrounds
the fault core and contains numerous fractures, slip surfaces,
and recognizable fragments of host rock. The core and
damage zones are surrounded by relatively undamaged host
rock.
[3] Development of this zonation, associated mesoscale
to microscale structures, and the mechanics of seismic and
aseismic deformation might influence the movement of
fluids in the shallow crust [McCaig, 1988; Knipe et al.,
1991; Antonellini and Aydin, 1994, 1995; Olivier, 1986;
Evans et al., 1997; Jones et al., 1998; Hanemberg et al.,
1999; Cello et al., 2001]. Several models have been
proposed to relate subsurface fluid flow to the structure of
faults [Scholz, 1990; Bruhn et al., 1994; Mozley and Good-
win, 1995; Matthai et al., 1998; Ferrill et al., 1999; Rawling
et al., 2001]. In low-porosity rocks, such as the carbonates
of this study, these models assume that fault cores have a
lower long-term permeability than the surrounding damaged
zones. This is broadly consistent with the results of numer-
ical 3-D fluid flow models demonstrating that fractures are
the primary feature controlling fluid flow in low-porosity
rocks [Caine et al., 1999].
[4] Fluid flow in faults depends not only on the structural
arrangement of fault rocks but also on the dynamic inter-
actions imposed by deformation on the fault zone-fluid flow
system. The role of tectonic loading on fluid flow in faults
was examined by Scholz et al. [1973] and later modeled by
Sibson et al. [1975], Sibson [1990, 1992], and Byerlee
[1993]. They proposed cyclic, heterogeneous fluid flow in
faults during the seismic cycle due to variations in devia-
toric stress. Individual faults can be both conduits and
barriers for fluid movement during one seismic cycle. For
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. B4, 2221, doi:10.1029/2002JB002013, 2003
1
Now at Department of Geological and Environmental Sciences,
Stanford University, Stanford, California, USA.
Copyright 2003 by the American Geophysical Union.
0148-0227/03/2002JB002013$09.00
ETG 18 - 1
a normal fault, numerous studies envision an increase in
permeability during the interseismic period that increases to
a maximum value just after failure [e.g., Brown and Bruhn,
1996]. Healing and sealing of fractures following migration
of fluid into a normal fault [Cello, 2000] may result in an
increase in pore fluid pressure if the healing/sealing rate is
high enough [cf. Marone, 1998]. When tectonic loading and
pore fluid pressure are sufficiently high, earthquake faulting
may occur, resulting in the discharge of fluids from the fault
into the surrounding host rock [Muir-Wood and King,
1993]. Clearly, fluid flow varies spatially and temporally
in faults.
[5] In this framework, we investigate the structures and
stable isotope geochemistry of four normal faults in central
Italy to understand fluid-fault interactions. The faults, which
are hosted in carbonate rocks, are well exposed along the
margins of large intramontane basins and were chosen for
study because they extend 10 km into the crust [Ghisetti and
Vezzani, 1999], are the loci of numerous earthquakes
[Boschi et al., 1997], and exhibited the largest isotopic
variations between fault rocks and host rocks in an earlier
structural/geochemical study of the same region [Ghisetti et
al., 2001]. This study documented contrasting paleofluid
circulations between contraction- and extension-related
structures of the orogen. Semiclosed to closed fluid system
conditions occurred during the earlier shortening of the
thrust system, while semiopen to open system conditions
prevailed during later uplift, exhumation, and normal fault-
ing. We now focus on the development of these fluid
conduits along basin-bounding normal faults to better
address the fluid migration through faults that are structur-
ally zoned.
2. Regional Setting
[6] Oligocene to Pliocene continental collision and A-
type subduction of the Adriatic-Apulian plate below the
European margin resulted in the development of an accre-
tionary wedge in which Meso-Cenozoic sediments were
deformed [Royden et al., 1987; Patacca et al., 1990;
Doglioni, 1991]. The result of this deformation is the
Apennine fold-and-thrust belt, a narrow, NW-SE elongated
belt of tectonic units arranged in arcs of different size and
curvature [cf. Locardi, 1988]. Concurrent shortening in
the foreland and downfaulting and crustal thinning in the
hinterland characterize much of the deformation in the
Apennines. Models taking into account coeval contraction
in the foreland and extension in the hinterland include
rollback of the subducting Adriatic slab [Malinverno and
Ryan, 1986; Jolivet et al., 1998], north-south Africa-Eurasia
convergence [Mantovani et al., 1996], and upwelling of
asthenospheric magma [Locardi, 1988].
[7] The contraction and extension has resulted in the
development of four distinct structural domains in the
central Apennines, from west to east (Figure 1a). Farthest
west is the Tyrrhenian Sea back-arc region, with a 10-km-
thick crust in the central portion of the basin and water
depths of 23 km. To the east is the thinned and down-
faulted peri-Tyrrhenian inner belt, which is composed of a
25- to 30-km-thick crust dissected by NW-SE trending
normal faults and associated basins [Bosi et al., 1995].
Farther east is the strongly shortened peri-Adriatic outer
belt, which is composed of a 35-km-thick crust that was
actively shortening until late Pliocene-early Pleistocene
[Ghisetti and Vezzani, 1999]. The peri-Adriatic outer belt
is currently undergoing extension and collapse by active
normal faulting. The Adriatic-Apulia foredeep, which is the
easternmost structural domain, is actively shortening.
[8] The four normal faults investigated in this study are
located in the peri-Adriatic outer belt of central Apennines
(Figure 1b). The tectonostratigraphic units of this region
are composed of Mesozoic-Tertiary shallow water and
transitional platform-to-basin carbonates of the Adriatic-
Apulian margin [Vezzani and Ghisetti, 1998]. This region
has been uplifted at rates of 1.42.5 mm/yr since 3.5 and
1.6 Ma, respectively [Ghisetti and Vezzani, 1999]. It has
been dissected by late Pliocene to Pleistocene high-angle
normal faults that have either reactivated preexisting thrust
faults [DAgostino et al., 1998] and/or formed along the
back limb of large folds [Ghisetti et al., 1994]. Many
earthquakes in central Italy are due to slip along these
normal faults. These faults are oriented approximately
NW-SE and lie in a 30- to 50-km-wide, 800-km-long
seismogenic zone that extends through the central and
southern Apennines with an overall direction of extension
toward the NNE [Piccardi et al., 1999, and references
therein].
3. Geology of the Normal Faults
[9] The four normal faults investigated in this study are
the (1) Venere-Gioia dei Marsi, (2) Campo Imperatore, (3)
Monte Capo di Serre, and (4) Roccacasale faults. These
faults are some of the larger-displacement normal faults in
the region with throws of 12 km. The footwalls of these
emergent faults expose fault rocks exhumed from 1 km
depth or less. The faults occur in a seismically active region
and two of the faults are known to have been seismogenic.
We assume the other two faults have also been seismically
active during part or all of their history because of their
similarity and proximity to historically documented seismo-
genic faults.
[10] We describe in detail the structures and fault rocks of
the four fault zones (Figures 2 and 3). The structures and
fault rocks are similar among the faults and are distributed
systematically in the faults. In the discussion, we provide a
generalized model for the fault zones that includes five
structural domains.
3.1. Venere-Gioia dei Marsi Fault
[11] The Venere-Gioia dei Marsi fault is part of a NW
trending, SW dipping fault zone along the eastern margin of
the Fucino Plain. Mesozoic and younger carbonate rocks
compose the footwall of the fault. The faults hanging wall
is composed of late Pliocene to late Pleistocene fluvio-
lacustrine sediments of the Fucino basin [Galadini and
Messina, 1994]. The basin is a half-graben structure that
formed by movement of the Venere-Gioia dei Marsi and
other normal faults along the eastern margin of the basin
[Mostardini and Merlini, 1988].
[12] The Venere-Gioia dei Marsi fault ruptured during the
M
s
= 7.0 Avezzano earthquake of 1915, resulting in more
than 30,000 casualties and structural damage in a 500 km
2
area [Boschi et al., 1997]. The 1915 earthquake epicenter
ETG 18 - 2 AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS
was near the town of Gioia dei Marsi [Galadini et al.,
1995], 1 km north of our study section. Focal mechanism
solutions place the T axis in the E-W [Gasparini et al.,
1985] to NNE-SSE directions [Basili and Valensise, 1991],
consistent with oblique extension, while trench investiga-
tions documented pure dip-slip motion of this fault during
the 1915 rupture [Micchetti et al., 1996]. Normal, right-
lateral slip of this fault has been documented by local
geomorphological analyses [Piccardi et al., 1999]. This
fault and others in the Fucino area have been responsible
for numerous earthquakes in the past, including those
between 885 and 1349 A.D., 550 and 885 A.D., 6000 and
3600 B.C., and 10,800 and 5600 B.C. [Micchetti et al.,
1996; Galadini and Galli, 1999]. The fault has accommo-
dated an average displacement rate of 0.41.0 mm/yr, with
down-dip motion toward 229 [Micchetti et al., 1996;
Piccardi et al., 1999].
[13] This fault is made of three NW striking segments.
We investigated the central segment, south of Venere, where
several quarries crosscut the faults contact and footwall.
Here the fault strikes between 110 and 140, and dips
between 55 and 65 SW (Figure 2a). The fault surface is
polished and striated with striations and slickenfibers con-
sistent with normal dip-slip motion and associated minor
left-lateral movement. A noticable scarp demarcates the
fault trace. Inversely graded, cemented slope scree drapes
over, and is partly offset by, the fault in some places. The
faults footwall is composed of dark white to gray, centi-
Figure 1. (a) Schematic cross section across central Italy [after Ghisetti et al., 2001]. Four distinct
structural domains are present: (1) Tyrrhenian Sea back-arc region, (2) thinned and downfaulted peri-
Tyrrhenian (p-T) inner belt, (3) strongly shortened peri-Adriatic (p-A) outer belt, and (4) Adriatic-Apulia
foredeep. Study area is in the peri-Adriatic outer portion of the belt, in the Abruzzo region of central Italy.
This region has been uplifted and dissected by late Pliocene to Pleistocene high-angle normal faults. (b)
Simplified tectonic map of central Apennines [after Ghisetti et al., 1994] and the four faults studied in
detail: A, Venere-Gioia dei Marsi; B, Campo Imperatore; C, Serra Capo di Monte; and D, Roccacasale
faults. Faults juxtapose deformed carbonates of footwalls against continental sediments of hanging walls.
AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS ETG 18 - 3
F
i
g
u
r
e
2
.
G
e
o
l
o
g
i
c
a
l
m
a
p
s
a
n
d
o
r
i
e
n
t
a
t
i
o
n
d
a
t
a
o
f
m
i
n
o
r
s
t
r
u
c
t
u
r
e
s
o
f
(
a
)
V
e
n
e
r
e
-
G
i
o
i
a
d
e
i
M
a
r
s
i
a
n
d
C
a
m
p
o
I
m
p
e
r
a
t
o
r
e
f
a
u
l
t
s
a
n
d
(
b
)
M
o
n
t
e
C
a
p
o
d
i
S
e
r
r
e
a
n
d
R
o
c
c
a
c
a
s
a
l
e
f
a
u
l
t
z
o
n
e
s
.
S
t
r
u
c
t
u
r
a
l
d
a
t
a
a
r
e
p
l
o
t
t
e
d
i
n
l
o
w
e
r
h
e
m
i
s
p
h
e
r
e
,
e
q
u
a
l
-
a
r
e
a
s
t
e
r
e
o
n
e
t
p
r
o
j
e
c
t
i
o
n
s
.
R
o
s
e
d
i
a
g
r
a
m
s
d
e
p
i
c
t
t
r
e
n
d
s
o
f
m
i
n
o
r
s
t
r
u
c
t
u
r
e
s
c
o
n
s
i
s
t
e
n
t
w
i
t
h
t
h
e
r
i
g
h
t
-
h
a
n
d
r
u
l
e
f
o
r
s
t
r
i
k
e
a
n
d
d
i
p
.
N
i
s
n
u
m
b
e
r
o
f
d
a
t
a
d
i
s
p
l
a
y
e
d
;
M
i
s
t
h
e
p
e
r
c
e
n
t
o
f
d
a
t
a
i
n
t
h
e
l
o
n
g
e
s
t
p
e
t
a
l
o
f
t
h
e
r
o
s
e
d
i
a
g
r
a
m
s
.
ETG 18 - 4 AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS
Figure 3. Representative outcrops of faults and fault rocks. (a) Campo Imperatore normal fault zone
viewed from SW. Fault separates core and damage zones of footwall from Quaternary sediments (Q.A.)
of hanging wall. (b) Venere-Gioia dei Marsi normal fault viewed from SW. Fault separates core and
damage zones of footwall from Quaternary sediments (Q.A.) of hanging wall Subsidiary normal faults in
footwall damage zone are oriented orthogonal to the main fault. (c) Hand sample from Campo Imperatore
fault. Fault surface is planar and coated with carbonate cements and thin bands of fine-grained cataclasite.
Within cement there are microscopic, rounded survivor clasts, while those in cataclasite are commonly
angular. Cemented horizons are commonly sheared. (d) Hand sample from Roccacasale fault. Fault rock
adjacent to fault surface contains matrix and clast-supported cataclasites with centimeter-thick horizons of
red and pink cements. Youngest slip surface is between striated alluvium (Q.A.) and relatively
undamaged alluvium (Q.A. at top of sample). (e) Photomicrograph of Campo Imperatore hand sample.
Survivor clasts and thin calcite veins are in cement-rich carbonate horizon. Sizes of survivor clasts
increase away from fault surface (f.s.). (f) Photomicrograph of Roccacasale hand sample. Fault surface
(f.s.) juxtaposes sheared alluvium (Q.A.) against matrix-supported cataclasite of footwall.
AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS ETG 18 - 5
meter- to meter-layered, lower Cretaceous limestones [Vez-
zani and Ghisetti, 1998].
[14] Two high-angle, E-W oriented left-lateral faults off-
set the fault in the western portion of the study area. These
faults are coated with pink, centimeter-thick cements; stria-
tions on internal slip surfaces are consistent with both
strike- and dip-slip motions. Low-angle stylolites that strike
parallel to the main fault and dip 1015 NE are also
present in this area. They are associated with millimeter-
thick veins of white carbonate, which display a radiating
pattern about a central point.
[15] The fault zone is 70100 m thick. In the footwall,
the fault core contains several distinct fault rocks. A 5- to
20-cm-thick cement-supported cataclasite with centimeter-
to millimeter-diameter survivor clasts is adjacent to the fault
surface. Next to this horizon is a meter-thick, clast- and
matrix-supported cataclasite with centimeter- to millimeter-
diameter survivor clasts. Adjacent to the cataclasite is a
meter-thick gouge horizon with clasts that are arranged in
centimeter- to decimeter-thick layers of uniform clast size.
The clasts are more rounded closer to the fault. Three sets of
fractures and one set of minor faults are present in the fault
core (Figure 2a).
[16] Numerous fractures, slip surfaces, and some veins
crosscut the host rock in the damage zone. This zone, which
is several decameters thick, is separated from the relatively
undeformed host rock by a subsidiary, dip-slip fault that
forms a distinctive morphologic scarp subparallel to the main
fault (Figure 3b). There are two primary sets of fractures, one
set of minor faults, and two sets of veins in the damage zone
(Figure 2a). The centimeter-thick veins are filled with white
carbonate. On the basis of a few crosscutting relations, the
SE striking veins formed before the east striking veins.
3.2. Campo Imperatore Fault
[17] The Campo Imperatore fault delimits the northern
margin of the Campo Imperatore Plain, which is an intra-
montane Quaternary basin. This basin is 17 km long, up to
4 km wide, and filled with middle to middle upper Pleis-
tocene gravels and glacial sediments, and upper Pleistocene
fluvial, lacustrine, and alluvial fan sediments [Giraudi,
1994]. The fault has been active since the middle Pleisto-
cene. It offsets the middle Pleistocene sediments by several
tens of meters, the middle upper Pleistocene sediments by
10 m, and the upper Pleistocene alluvial fans and terraces
by 1 m [Giraudi, 1994].
[18] The study area is on the western margin of the
Campo Imperatore Plain. The fault cuts through an early
Jurassic dolomitic host rock of the Vado di Corno tectonic
unit [Vezzani and Ghisetti, 1998], and juxtaposes these
decimeter-thick, cement-supported carbonates against pink,
very cohesive, fluvial and lacustrine Quaternary sediments
of the hanging wall. Slope debris locally drapes the fault
surface. The fault contains several segments that strike
between 105 and 130, and dip 5060 SW (Figure 2a).
The fault surface is polished, striated, and coated with new
carbonate minerals. The striations are consistent with obli-
que, right-lateral slip along the fault segments striking 105
110, and pure dip-slip along the 120130 segments.
[19] In the study area, the deformed footwall of the fault is
up to 130 m thick. The fault core with an average thickness
of 1520 m contains decimeter- to meter-thick clast-sup-
ported cataclasites, and decameter-thick fault gouge. The
white, clast-supported cataclasites close to the fault consist
of centimeter-diameter, angular to subrounded survivor
clasts, which are occasionally embedded in a fine-grained
matrix and white carbonate cements. In the footwall, farther
from the fault, the core zone is composed primarily of gouge,
which contains decimeter- and centimeter-diameter, angular
fragments arranged in thin layers of uniform clast size. The
gouge is thicker in areas where the strike of the fault changes
significantly or where subsidiary faults merge into the main
fault. There are three sets of high-angle fractures, three sets
of minor faults, and three sets of veins in the faults core
(Figure 2a). On the basis of crosscutting relations, the NW
striking, NE dipping reverse faults were the first structures to
form in the core zone. The east striking, south dipping veins
are filled with white and pink carbonates and display two
phases of growth. The SE and NE striking veins are filled
with white carbonate minerals; the NE striking veins exhibit
evidence for several episodes of growth. The damage zone,
which is very distinctive in the footwall, contains three sets
of fractures (Figure 2a). Minor faults subparallel to the main
fault separate the damage zone from the relatively undam-
aged host rock.
[20] Major element analyses of samples collected along
the fault are consistent with the host rocks and cataclasites
fine-grained matrix being a mixture of dolomite and low-
Mg calcite (Table 1). Cements collected on the fault surface
are composed primarily of low-Mg calcite.
3.3. Monte Capo di Serre Fault
[21] The Monte Capo di Serre fault borders the western
flank of the Le Riparate Mountain, a few kilometers south-
east of the Campo Imperatore Plain. The fault is a metric-
scale undulated surface, both parallel and perpendicular to
the strike. It juxtaposes Cretaceous limestones of the foot-
wall against Jurassic dolomites of the hanging wall [Vezzani
and Ghisetti, 1998]. Slope debris with angular to sub-
rounded clasts embedded in yellow to pink carbonate
cements covers the fault surface, and is locally crosscut
by numerous minor fractures. Centimeter-thick portions of
Table 1. Relative Abundance of Principal Elements in Carbonate
Host Rocks and Cemented Cataclasites of Domain 4 in Three
Faults
a
Rock Type
Ca,
wt %
Mg,
wt %
Fe,
wt %
Sr,
ppm
Si,
wt %
LOI,
wt %
Campo Imperatore Fault
Host rock 23.9 10.8 0.03 96 0.05 65.1
Yellow cement 32.8 4.23 0.16 36 0.83 61.5
White cement 25.8 9.66 0.01 56 0.06 64.4
Yellow cement 33.2 4.72 0.05 87 0.15 61.8
Monte Capo di Serre
Host rock 39.2 0.37 0.02 162 0.07 60.3
Pink cement 39.4 0.20 0.04 77 0.17 60.0
White cement 39.5 0.11 b.d. 26 0.03 60.3
Pink cement 39.1 0.17 0.03 70 0.13 60.4
Roccacasale
Host rock 39.2 0.35 0.06 178 0.29 59.9
Red cement 34.3 3.22 0.10 45 0.13 62.1
Yellow cement 28.7 7.40 0.23 52 0.44 64.0
Pink cement 26.6 8.67 0.10 55 0.35 64.8
a
Data obtained by XRF analysis.
ETG 18 - 6 AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS
slope debris and talus have been locally sheared and
accreted onto the faults footwall.
[22] In the study area, the Monte Capo di Serre fault
strikes between 095 and 150 and dips 5070 W (Figure
2b). A fault scarp, up to 9 m high, characterizes the fault
trace. The fault surface is polished and striated with kine-
matic indicators consistent with normal oblique faulting.
Uncommon 5- to 10-cm-long ring cracks (i.e., cracks with
an arcuate shape [see Scholz, 1990, Figure 2.12]) formed on
the fault surface in the western portion of the study area.
Cemented cataclasites adjacent to the fault surface contain
millimeter- and centimeter-diameter, angular to subrounded
survivor clasts embedded in carbonate cement. Locally, up
to three different millimeter- and centimeter-thick red hori-
zons are present within these rocks, each of which contain
very small survivor clasts embedded in carbonate cements.
These red horizons are characterized by internal wavy
surfaces on the footwall side and polished, striated, planar
surfaces on the hanging wall side.
[23] The fault zone is not well developed along the Capo
di Serre fault; in the footwall, its width ranges from a few
meters up to 35 m and consists of a discontinuous fault
core flanked by a narrow damage zone. The fault core
contains decimeter-thick, clast-supported cataclasites, and
discontinuous meter-thick fault gouge. The cataclasites are
made of centimeter-diameter, subangular survivor clasts
with rare matrix and white and light pink cements. The
fault gouge is present only along the eastern portion of the
fault in the study area. It consists of centimeter- and
decimeter-diameter, angular to subrounded fragmented
clasts, commonly embedded in a fine-grained matrix.
Numerous millimeter- to centimeter-thick white veins are
present locally along the fault contact. They occur where the
fault zone is poorly developed and parallel the NW striking
and gently SW dipping bedding of the limestone host rock.
There are two sets of fractures and one set of minor faults in
the faults core (Figure 2b).
[24] In the damaged host rock of the footwall, the original
sedimentary fabric is well preserved. There are two sets of
fractures in the damage zones (Figure 2b). The fractures are
parallel and orthogonal to the fault surface, similar to those
in the fault core. The few minor faults crosscutting the
damage zone are parallel and oblique to the main fault, and
are associated with poorly cemented breccia.
[25] The Monte Capo di Serre fault zone host rocks and
cements are primarily low-Mg calcite (Table 1). East of the
study area, centimeter-thick flame structures cut across the
entire fault core. These structures contain centimeter- and
decimeter-thick, angular clasts embedded in pink carbonate
cements. They may be evidence of high-pressure, coseismic
fluid flow through the fault.
3.4. Roccacassale Fault
[26] The Roccacasale fault is part of a hundreds of meters
wide fault zone crosscutting the western flank of Morrone
Mountain and bordering the eastern part of the Sulmona
basin [Vezzani and Ghisetti, 1998]. This basin is a NW
elongated trough that developed after the late Pliocene and
is the easternmost intramontane basin of central Apennines.
It is filled with alluvial and lacustrine Pleistocene sediments,
upper Pleistocene-Holocene fluvial sediments, and very
recent colluvium and alluvial fans [Miccadei et al., 1998].
According to historical records [Boschi et al., 1997], there
were no significant earthquakes in this area until 1706.
Since then, there have been earthquakes of intensity V
(Mercalli scale) in 1789 and 1933 A.D.
[27] Our investigation focused on a prominent fault scarp
above the Roccacasale village. The fault separates upper
Cretaceous limestones and Jurassic dolostones of the foot-
wall from Quaternary fluvial, lacustrine, and alluvial depos-
its of the hanging wall [Vezzani and Ghisetti, 1998]. The
study portion of the Roccacasale fault comprises two main
segments, which strike 105110 and 120130, respectively
(Figure 2b). The fault surface is polished and striated, and
forms a distinctive fault scarp. The surface formed on a
centimeter-thick, yellow to pink, cemented horizon that is
occasionally associated with centimeter-thick fault breccia.
These cements consist of millimeter-diameter, angular to
subrounded survivor clasts embedded in light colored
carbonate cements. Two thin horizons containing milli-
meter-diameter clasts embedded in red carbonate cements,
plus a few millimeter-thick veins filled with white carbo-
nates, are present within this rock.
[28] The fault is covered locally by cohesive, pinkish
slope scree that drapes the fault surface. Kinematic indica-
tors along the fault surface are consistent with oblique slip
(normal with minor right-lateral movement) on the 105110
segment, and pure dip slip on the 120130 segment. East of
the study area, the fault is buried beneath Quaternary
continental deposits of the hanging wall. In the footwall,
the fault zone consists of a narrow, discontinuous fault core
and a very wide damage zone. The fault core contains
decimeter-thick, clast- and matrix-supported cataclasites,
and discontinuous meter-thick gouge. The cataclasites con-
tain centimeter-diameter, subangular clasts in a white to light
colored matrix. The fault gouge is made of centimeter- and
decimeter-diameter, angular to subrounded fragmented
clasts embedded in fine-grained matrix. There is one primary
set of fractures and three sets of minor faults in the core of
the fault (Figure 2b). The damage zone is hundreds of meters
thick and has two sets of fractures and two sets of minor
faults (Figure 2b). There are many fractures near these minor
faults, resulting in a very incohesive rock in outcrop.
[29] The cemented horizons in the fault core are com-
posed of dolomite survivor clasts embedded in a low-Mg
calcite cement. Both dolomite and low-Mg calcite are
present in the host rock of the footwall.
4. Stable Isotope Analyses
[30] We use data from stable isotope geochemistry to
reconstruct the occurrence and origin of fluid(s) within the
fault, and the movement of fluid through the fault zone.
4.1. Methodology
[31] Hand samples were selected from the outcrops in the
structural context provided by our detail mapping and
structural analysis. Most samples were cut perpendicular to
the orientation of the main fault at each locality and parallel
to the inferred direction of slip. Powder samples were
obtained by microdrilling on the cut surface or by whole
rock powdering. More than two hundred powder samples
obtained from 139 hand samples were analyzed by stable
isotope geochemistry to determine the isotopic composition
AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS ETG 18 - 7
of (1) survivor clasts, (2) uncemented matrix surrounding the
survivor clasts, (3) nonmineralized slickensides, (4) slick-
enfibers, (5) veins, (6) mineralized slickensides of slope scree
and Quaternary fluvio-lacustrine deposits, and (7) cement
surrounding survivor clasts. The survivor clasts, uncemented
matrix, and nonmineralized slickensides (groups 13) do not
contain mesoscopically observable cement or veins, or any
other evidence for significant fluid-rock interaction. Slick-
enfibers, veins, mineralized slickensides, and cements sur-
rounding survivor clasts (groups 4 7) formed by
precipitation from a fluid. The isotopic values obtained from
the latter set of powder samples should more closely correlate
with the isotopic value(s) of the fluid(s) in the faults.
[32] The stable isotope compositions of the powders were
obtained by two different techniques, both of which
involved phosphoric acid digestion of the carbonates
[McCrea, 1950]. The first technique involved the reaction
of 1020 mg aliquots of sample powder in a conventional
vacuum-extraction line (details are given by Kirschner et al.
[2000]). The extractions were done at 50C for 56 hours.
At least one in-house standard was analyzed with each
series of nine samples. The extracted gas was analyzed on
an isotope ratio, gas source mass spectrometer at Washing-
ton University (St. Louis). The second technique involved
the reaction of 0.5 mg of sample powder in an automated
carbonate extraction device connected to a continuous flow,
isotope ratio, gas source mass spectrometer at Saint Louis
University. These reactions were done at 90C for a mini-
mum of 4 hours. Approximately one standard was analyzed
for every five samples. Values of d
13
C and d
18
O for the in-
house standard were calibrated relative to NBS-19. Accep-
ted NBS-19 values are d
13
C VPDB = 1.95% and d
18
O
VSMOW = 28.64% [Copeland et al., 1983]. For conven-
tional extractions, the standard deviation is 0.05% for d
13
C
and 0.19% for d
18
O for 21 in-house standard analyses. For
automated extractions, the standard deviation is 0.03% for
d
13
C and 0.23%for d
18
O for 21 in-house standard analyses.
4.2. Results
[33] Isotopic data of each fault zone define similar arcuate
trends in d
18
O - d
13
C space with d
18
O values ranging from
22 to 34% and d
13
C values from 8 to +4% (Figure 4).
Figure 4. Isotope data of samples in d
18
O - d
13
C space. Structures are subdivided into different groups;
those denoted by solid symbols are relicts of the host rock (groups 13 in text), while open symbols
represent samples precipitated from a fluid (groups 47). Most samples in groups 13 have isotopic
values typical of Mesozoic marine carbonates. Many samples in groups 47 have lower values,
consistent with these samples having exchanged with, or precipitated from, an externally derived fluid,
probably of meteoric origin.
ETG 18 - 8 AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS
Survivor clasts, uncemented matrix, and nonmineralized
slickensides have the highest isotopic values, similar to
the isotopic values of most Mesozoic and younger marine
limestones and dolostones in central Italy [Corfield et al.,
1991; Jenkyns et al., 1994; Bartolini et al., 1996; Ghisetti et
al., 2001]. Slickenfibers, veins, mineralized slickensides,
and cement surrounding survivor clasts have the lowest
isotopic values, consistent with isotopic exchange and/or
precipitation from a fluid in isotopic disequilibrium with the
host rock carbonates.
5. Discussion
[34] The four normal faults of this study are similar in
their dimensions, displacements, isotopic values, and iso-
topic trends in d
18
O - d
13
C space. In the following dis-
cussion, we assume the processes responsible for the faults
development were similar in order to develop a general
description and model of the faults structure and evolution,
and the spatial-temporal variation in fluid-rock interactions.
We then consider briefly the possible dynamic fluid/fault
interactions occurring at depth in these seismogenic faults.
5.1. Fault Structure
[35] Most of the fault rocks and structures in the four
normal faults can be separated into five spatial domains
(Figure 5a). Within the footwall, farthest from the fault is a
domain (domain 1) of moderately to highly fractured rock
that is many tens of meters thick. Although fractured,
bedding planes and stratigraphy are still recognizable. The
fractures strike parallel and perpendicular to the main fault
surface (Figure 5b). Small, centimeter-thick veins are
present locally in this domain. Domain 1 is commonly
separated from the relatively undeformed host rock by a
subsidiary normal fault that subparallels the main fault and
delimits the outer margin of the faults damage zone.
[36] Domains 25 all reside in the faults core. Domain 2
is composed of fault gouge that is commonly 1 m thick, but
may be locally thicker in areas where two segments of the
normal fault link, intersect, or overlap. Bedding planes and
stratigraphy are very disrupted and generally unrecogniz-
able. The gouge contains angular to subrounded survivor
clasts that are graded according to clast size in horizons that
subparallel the main fault. In general, the clasts decrease in
size toward the fault.
[37] Domain 3 is a centimeter- to decimeter-thick catacla-
site. The rock is made of millimeter- to centimeter-diameter,
angular to rounded survivor clasts in a finer-grained, light
colored friable matrix. Sparse, centimeter-thick veins locally
crosscut the cataclasite. The dominant set of fractures in this
domain strikes parallel to the main fault; a second set of
fractures strikes perpendicular to the main fault (Figure 5b).
[38] Domain 4 contains the primary slip surface(s) of the
normal fault. This surface is usually planar, polished, and
striated with slickensides and some slickenfibers. Sense-of-
shear displacement from slickensides and slickenfibers is
almost always consistent with predominantly normal, dip-
slip movement. Locally, there are centimeter-thick zones of
cement-rich horizons (classified above as mineralized slick-
ensides), each of which has a slickensided planar surface
facing the hanging wall and an irregular, slightly gradational
contact with the underlying cataclasite. We interpret these
horizons to be, or have been, the primary slip surface during
part of the faults history.
[39] Domain 5 contains Quaternary sediments of the
hanging wall. These sediments were deformed by slip of
the fault, and form up to a meter(s)-thick, poorly cemented
breccia along the fault. Locally, centimeter-thick slices of
hanging wall sediments have been accreted onto the foot-
wall by outward migration of the primary fault surface. The
fault trace is commonly covered by, and locally offsets,
slope scree that drapes over the fault.
5.2. Proposed Evolution of Faults
[40] We propose a model for the temporal development of
the shallow portion (12 km depth) of the generalized
Figure 5. (a) Schematic cross section of fault rocks in
carbonate-hosted normal fault. Fault zone contains five
domains from fractured host rock of footwall (domain 1) to
brecciated Quaternary alluvium (domain 5). (b) Rose
diagrams of subsidiary faults and fractures in plan and
cross-sectional orientations for core and damage zones. For
ease of comparison, structures have been rotated so that the
main fault at each locality parallels the Venere-Gioia dei
Marsi fault. Most structures in core zone strike parallel to
faults contacts, while in damage zone, most structures are
either parallel or orthogonal to the faults contacts.
AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS ETG 18 - 9
fault zone presented above based on the current distribution
of the exhumed fault rocks and mesostructures, and few
crosscutting relations. As a working hypothesis, we assume
that deformation began in a broad zone and then localized in
the core of the fault during continued slip. In the earliest
stages of deformation in late Pliocene-early Pleistocene, the
rocks currently exposed in the faults footwall began frac-
turing at depth (Figure 6a). Both the footwall and hanging
wall fractured in tens of meters thick zones, which estab-
lished the approximate overall width of the fault zone in the
footwall. Fractures formed parallel and perpendicular to the
evolving fault contact. Some calcite coatings on minor
faults, veins, and breccias formed at this time from rock-
buffered fluids circulating through the fault rocks.
[41] Deformation progressively localized in the incipient
core of the fault with continued fault slip. Fracturing and
cementation/veining continued in both hanging wall and
footwall (Figure 6b). By this stage, core and damage zones
would have been established and centered about the principal
fault contact. New fractures subparallel to the main faults
continued to form primarily in the fault core. Fault displace-
ment controlledthedevelopment of small intramontane basins
on the hanging wall of these emergent faults. Eventually,
fluvial and lacustrine sediments in the basins were juxtaposed
against the fault rocks of the footwall. Calcite continued to
precipitate in the fault zone from rock-buffered fluids.
[42] The latter stage of deformation was characterized by
complete localization of slip at the hanging wall-footwall
contact, and the episodic outward migration of the principal
fault surface (Figure 6c). Red, pink, and white low-Mg
calcite cements precipitated along the fault surface, and were
sheared. The outer surface of these centimeter-thick
cemented horizons (i.e., the surface facing the hanging wall)
would have been the main slip surface of the fault zone for
one or more slip events. Seismic release during this late
faulting was potentially responsible for the formation of the
ring cracks on these surfaces in that their arcuate shape is
consistent with high tensile stresses behind the trailing edge
of a contacting asperity [cf. Scholz, 1990]. Locally, the
migrating fault surface crosscuts centimeter-thick slivers of
slope scree draping the fault, resulting in the accretion of the
brecciated scree onto the footwall. This occurred primarily in
the nonplanar regions of the fault surface. Two contrasting
processes, smearing and segmentation of the fault surfaces,
characterized the latter stages of deformation: smearing due
to shearing and accretion of slope scree breccia onto the
footwall, and segmentation of fault surface due to the activity
of minor faults laterally offsetting the principal fault surface.
5.3. Source of Syntectonic Fluid
[43] The veins, slickenfibers, and cements attest to the
presence of fluids in the fault zone during deformation. The
primary fluid was probably evolved meteoric water for four
reasons. First, the emergent, active normal faults have been
exhumed fromshallowdepth, and thus the fault rocks formed
in close proximity to near-surface aquifers. Second, the
mineralized slickensides, slickenfibers, and veins have iso-
topic values similar to the cements of the hanging wall sedi-
ments, which clearly precipitated from a fluid of meteoric
origin. Third, meteoric water is one of themore commonfluids
in the earths crust with sufficiently lowd
18
Oand d
13
Cvalues
to account for the lowisotopic values of calcite precipitated at
low temperatures. Fourth, strontium concentrations of the
mineralized slickensides, slickenfibers, veins, and cements
are much lower than the carbonate host rocks (Table 1),
consistent with their precipitation from a low-salinity fluid
[Banner, 1995]. Although not definitive, the incursion of
deeply sourced CO
2
-bearing fluids (e.g., from mantle, mag-
mas, or metamorphic decarbonation of limestones) into the
upper crustal levels of the normal faults is not supported by the
Figure 6. Cartoon showing evolution of normal faults. (a)
Carbonate host rock fractured in tens of meters to hundred-
meter-thick zone during initiation of faulting. Fractures
formed parallel and perpendicular to eventual principal
fault. (b) Deformation localized around main fault during
progressive displacement and exhumation, resulting in
formation of core and damage zones. Small, continental
basins developed on faults hanging wall. (c) Quaternary
continental alluvium eventually juxtaposed against fault
rocks of footwall, resulting in complete localization of slip.
The principal slip surface migrated into hanging wall due to
formation of cement-dominated horizons and accretion of
hanging wall slivers.
ETG 18 - 10 AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS
data. This interpretation is consistent with the analyses
of Chiodini et al. [1999] for CO
2
fluxes in regional aquifers
and Quattrocchi et al. [2000] geochemical data of a thermal
spring in the eastern part of central Apennines.
[44] Assuming constant isotope composition of the mete-
oric water that entered the faults, then the relative position
of individual datum in the d
18
O - d
13
C arcuate trend is due to
variations in the partly evolved nature of the water and/or
quantity of water that formed or exchanged with each
sample (Figure 4). Those samples with isotopic values
similar to the carbonate host rocks either interacted with
small quantities of meteoric water or precipitated from a
meteoric water that had previously equilibrated with the
host rock. Conversely, those samples with lower isotopic
values either interacted with larger quantities of meteoric
water or precipitated from meteoric water that had partly
retained its original isotopic values. It is not possible with
the data to eliminate any of these possibilities.
5.4. Paleohydrogeology
[45] Fault rocks collected at different distance from the
principal fault surfaces are characterized by different stable
isotope values (Figure 7). In domains 1, 2, and 3, all samples
including cements and veins have d
18
O values between 26
and 32%and d
13
C of 2 and +2%, consistent either with a
rock-dominated isotopic system or fluid-rock interaction
with fluid that had equilibrated with the host rocks prior to
entering the investigated segments of the fault zones. In
contrast, most mineralized slickensides, slickenfibers, and
veins fromdomains 4 and 5 have isotopic values that are up to
6%lower, similar to values of the cements in the Quaternary
sediments and consistent with a fluid-dominated isotopic
system. The lower isotopic values of domains 4 and 5 relative
to domains 13 preclude large quantities of fluid having
pervasively entered the faults fromthe footwall side. The data
are most consistent with meteoric water having entered the
faults either from the hanging wall side and/or channeled into
the faults directly at the land surface [cf. Antonellini and
Aydin, 1995]. Both flow paths are compatible with the data.
[46] The differences in isotopic values between domains
13 and domains 4 and 5 are consistent with compartmen-
talization of meteoric water in the fault zone. Meteoric water
was able to move along the fault surface (domains 4 and 5)
and into/out of the hanging wall, but not move laterally into
adjacent footwall (domains 13). We propose that the fault
gouge and cataclasites of domains 2 and 3 were the primary
barriers to fluid flow [cf. Zhang and Tullis, 1998] and that
the fractured carbonates in domain 1 might have accom-
modated fluid flow albeit in a closed, isotopically rock-
buffered system. Such compartmentalization of fluids
would be a variant of the combined conduit barrier perme-
ability structure discussed by Caine et al. [1996] and Caine
and Forster [1999]. In contrast to the other three more
mature faults, the Monte Capo di Serre fault does not have
much gouge and cataclasites (domains 2 and 3) and thus
effectively forms only one fluid conduit, which would be
classified as a localized conduit permeability structure of
Caine et al. [1996] and Caine and Forster [1999].
5.5. Dynamic Fluid/Faults Interactions
[47] Data from paleoseismic studies [Giraudi, 1994; Mic-
chetti et al., 1996; Pantosti et al., 1996] and historical
seismicity are consistent with these faults having been
seismically active during exhumation [Armijo et al., 1985;
Gasparini et al., 1985; Galadini et al., 1995; Basili and
Valensise, 1991]. These west dipping, high-angle normal
faults merge asymptotically with west dipping, low-angle
detachment surfaces at depths of 1015 km [Bally et al.,
1988; Amato et al., 1998]. The zone of earthquake nucleation
is localized at these depths, in the region where the Tyr-
rhenian crust overrides the subducting Adriatic plate (focal
volume region of Figure 1a). There are potentially deep-
seated fluids at these depths due to dehydration reactions
occurring within Triassic evaporites and Paleozoic crystalline
rocks of the Tyrrhenian basement [Quattrocchi, 1999]. The
fluids are thought to play an important role in the earthquake
nucleation/rupture in central Italy [cf. Cello, 2000]. Because
of the load-weakening character of normal faults after
Figure 7. The d
18
O and d
13
C data of fault rocks relative to
their position in fault zones (compare Figure 5). Domain
Q represents data from Quaternary sediments on hanging
walls. A clear isotopic jump occurs between domains 3 and
4, consistent with a rock-buffered system in domains 13,
and potentially fluid-buffered (open) system in domains 4
and 5. Low isotopic values are consistent with meteoric
fluid that entered faults either from hanging wall and/or
directly at land surface. Symbols are identical to those in
Figure 4.
AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS ETG 18 - 11
coseismic slip, models predict dilation and vertical fracturing
of faults just after failure, subsequent incursion of fluids into
the faults with episodes of fluid overpressure development,
and further seismicity [e.g., Quattrocchi, 1999].
[48] Occurrence of coseismic fluid flow is thought to take
place along the study faults, consistent with the centimeter-
thick flame structures in the Monte Capo di Serre fault core.
At least at shallow levels of 12 km, slip localized on these
faults during their progressive exhumation, concurrent with
the flow of meteoric water along the main fault contacts.
There is no isotopic evidence of deep-seated fluids having
passed these faults at shallow levels. Consequently, those
models that envision deep-seated fluids passing up through
shallow levels of these seismogenic faults are not supported
by this study. If there are deep-seated fluids passing up
through the faults, then the fluids are either being diverted
from the fault or isotopically equilibrated with the Meso-
Cenozoic carbonates before they reach 12 km depth.
[49] The corrugated fault surfaces observed in this study
potentially form channels for fluid flow at shallow depths.
These conduits are oriented both parallel and perpendicular
to the faults strikes along the hanging wall-footwall con-
tacts. The along-strike hydraulic communication is pre-
dicted to occur at depth within normal faults [Sibson,
2000]. The observation of multiple aftershock sequence
progressively extending along strike after the 19971998
Umbria-Marche sequence in central Italy [Amato et al.,
1998], and experimental results obtained by Evans et al.
[1997] also support this prediction.
6. Conclusions
[50] Internal structures of four carbonate-hosted normal
faults in central Italy are similar and can be separated into
five structural domains. Farthest from the main fault contact
in the footwall is undamaged host rock and a zone of
fractured host rock (domain 1). The fractures in domain 1
are both parallel and orthogonal to the main fault contact.
Progressively nearer to the fault contact is gouge (domain 2),
cataclasites (domain 3), and cement-dominated horizons
with polished, slickensided fault surfaces (domain 4). A
narrow breccia horizon (domain 5) of hanging wall sedi-
ments is present locally next to domain 4. Small slivers of
this breccia had been accreted onto the footwall, resulting in
laterally migration of the main fault contact toward the
hanging wall. Fractures in the core of the fault (domains
25) are primarily parallel to the main fault contact. In
general, fracture density increases toward the main fault
contact.
[51] Stable isotope data from the four faults form an
arcuate trend in d
18
O - d
13
C space, consistent with variable
fluid-rock interaction and compartmentalization of fluid in
the fault domains. The relatively high d
13
C and d
18
O values
of host rocks, fault rocks, veins, and cements from domains
1 and 3 are consistent with little to no isotopic exchange
with a fluid in isotopic disequilibrium with the host rocks.
In contrast, the relatively low d
13
C and d
18
O values of the
cemented horizons in domain 4 and brecciated hanging wall
of domain 5 are consistent with meteoric water having
infiltrated and isotopically exchanged with these fault rocks.
The meteoric water entered fault domains 4 and 5 either
laterally from the hanging wall sediments or longitudinally
down from the land surface, and were prevented from
passing laterally into the footwall carbonates by the gouge
and cataclastic barriers of domains 2 and 3.
[52] Two of the four study faults have produced large,
historic earthquakes; the other two are presumed to have
been seismogenic due to their proximity and similarities
with seismogenic faults. At least at shallow levels of 12
km, displacement on these seismogenic faults has been very
localized. Meteoric fluids have flowed along the main fault
contacts. There is no isotopic evidence of deep-seated fluids
having passed through these faults. Consequently, any
model of earthquake nucleation/rupture that envisions large
quantities of deep-seated fluids passing up through shallow
levels of seismogenic faults is not supported by our data.
[53] Acknowledgments. We thank F. Ghisetti and L. Vezzani for their
help during the fieldwork and input during the writing of an earlier version
of the manuscript. We thank R. Criss and D. Kremser of Washington
University (St. Louis, Missouri) for providing us access to their labs.
Fieldwork and laboratory expenses were partly supported by Sigma XI to
Agosta and American Chemical Society (Petroleum Research Fund 31943-
GB2) to Kirschner. We gratefully acknowledge an equipment grant from
National Science Foundation (DUE 9952256) to Kirschner that supported
the stable isotope laboratory at Saint Louis University. The reviews of R.
Bruhn, an anonymous reviewer, Rudi Wenk (JGR Associate Editor), and
Andreas Mulch are much appreciated.
References
Amato, A., et al., The 1997 Umbria-Marche, Italy, earthquake sequence: A
first look at the main shocks and aftershocks, Geophys. Res. Lett., 25,
28612864, 1998.
Antonellini, E. M., and A. Aydin, Effect of faulting on fluid flow in porous
sandstone: Petrophysical properties, AAPG Bull., 78, 355377, 1994.
Antonellini, E. M., and A. Aydin, Effect of faulting on fluid flow in porous
sandstones: Geometry and spatial distribution, AAPG Bull., 79, 642671,
1995.
Armijo, R., A. Deshamps, and J. P. Poirie, Carte Sismotectonique Europe et
Bassin Mediterraneen, Inst. de Phys. du Globe de Paris, Paris, 1985.
Bally, A. W., L. Burbi, R. Cooper, and R. Ghelardoni, Balanced sections
and seismic reflection profiles across the central Apennines, Mem. Soc.
Geol. It., 35, 257310, 1988.
Banner, J. L., Application of the trace element and isotope geochemistry of
strontium to studies of carbonate diagenesis, Sedimentology, 42, 805
824, 1995.
Bartolini, A., P. O. Baumgartner, and J. Hunziker, Middle and Late Jurassic
carbon stable isotope-stratigraphy and radiolarite sedimentation of the
Umbria-Marche basin (central Italy), Eclogae Geol. Helv., 89, 811
844, 1996.
Basili, A., and G. Valensise, Contributo alla caratterizzazione della sismi-
cita dellarea marsicano-fucense, in Aree Sismogenetiche e Rischio Sis-
mico in Italia, pp. 197214, Ist. Naz. di Geofis. e Vulcanol., Rome,
1991.
Boschi, E., E. Guidoboni, G. Ferrari, and G. Valensise, Catalogo dei forti
terremoti in Italia dal 461 A.C. al 1990, 644 pp., ING-SGA, 1997.
Bosi, C., F. Galadini, and P. Messina, Stratigrafia Plio-Pleistocenica della
conca del Fucino, Quaternario, 8, 8993, 1995.
Brown, S. R., and R. L. Bruhn, Formation of voids and veins during
faulting, J. Struct. Geol., 18, 657671, 1996.
Bruhn, R. L., W. T. Parry, and T. Thompson, Fracturing and hydrothermal
alteration in normal fault zone, Pure Appl. Geophys., 142, 609644,
1994.
Byerlee, J., Model for episodic flow of high-pressure water in fault zones
before earthquakes, Geology, 21, 303306, 1993.
Caine, J. S., and C. B. Forster, Fault zone architecture and fluid flow:
Insights from field data and numerical modeling, in Faults and Subsur-
face Fluid Flow in the Shallow Crust, Geophys. Monogr. Ser., vol. 133,
edited by W. Haneberg et al., pp. 101127, AGU, Washington, D.C.,
1999.
Caine, J. S., J. P. Evans, and C. B. Forster, Fault zone architecture and
permeability structure, Geology, 24, 10251028, 1996.
Cello, G., A quantitative structural approach to the study of active faults in
the Apennines (peninsular Italy), J. Geodyn., 29, 265292, 2000.
Cello, G., E. Tondi, L. Micarelli, and C. Invernizzi, Fault zone fabrics and
geofluid properties as indicators of rock deformation modes, J. Geodyn.,
32, 543565, 2001.
ETG 18 - 12 AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS
Chester, F. M., and J. M. Logan, Composite planar fabric of gouge from the
Punchbowl fault, California, J. Struct. Geol., 9, 621634, 1986.
Chiodini, G., F. Frondini, D. Kerrick, J. Rogie, F. Parello, L. Peruzzi, and
A. Zanzari, Quantification of deep CO
2
fluxes from central Italy. Exam-
ples of carbon balance for regional aquifers and of soil diffuse degassing,
Chem. Geol., 159, 205222, 1999.
Copeland, T., C. Kendall, and J. Hopple, Comparison of stable isotope
reference samples, Nature, 302, 236238, 1983.
Corfield, R. M., J. E. Cartlidge, S. I. Premoli, and R. A. Housley, Oxygen
and carbon isotope stratigraphy of the Palaeogene and Cretaceous lime-
stone in the Bottaccione, Gorge and Contessa highway sections, Umbria,
Italy, Terra Nova, 3, 414422, 1991.
DAgostino, N., N. Chamot-Rooke, R. Funiciello, L. Jolivet, and F. Sper-
anza, The role of pre-existing thrust faults and topography on the styles of
extension in the Gran Sasso Range (Italy), Tectonophysics, 292, 229
254, 1998.
Doglioni, C., A proposal for kynematic modeling of W-dipping subduc-
tionsPossible applications to the Tyrrhenian-Apennines system, Terra
Nova, 3, 423434, 1991.
Evans, J. P., C. B. Forster, and J. V. Goddard, Permeability of fault-related
rocks, and implications for hydraulic structure of fault zones, J. Struct.
Geol., 19, 13931404, 1997.
Ferrill, D. A., J. Stamatakos, A. John, and D. Sims, Normal fault corruga-
tion; implications for growth and seismicity of active normal faults,
J. Struct. Geol., 21, 10271038, 1999.
Galadini, F., and P. Galli, The Holocene paleo-earthquakes on the 1915
Avezzano earthquake faults (central Italy): Implications for active tec-
tonics in the central Apennines, Tectonophysics, 308, 143170, 1999.
Galadini, F., and P. Messina, Stratigraphy of continental deposits, tectonics
and Quaternary geologic evolution of the Sangro River valley, southern
Abruzzi, Boll. Soc. Geol. It., 34, 877892, 1994.
Galadini, F., P. Galli, C. Giraudi, and D. Molin, The 1915 earthquake and
the seismicity of the Fucino Plain, central Italy, Boll. Soc. Geol. It., 3,
635663, 1995.
Gasparini, C., G. Iannacone, and R. Scarpa, Fault-plane solutions and seis-
micity of the Italian peninsula, Tectonophysics, 117, 5978, 1985.
Ghisetti, F., and L. Vezzani, Depth and modes of Pliocene-Pleistocene
crustal extension of the Apennines (Italy), Terra Nova, 11, 6772, 1999.
Ghisetti, F., U. Fullador, R. Casnedi, and L. Vezzani, Assetto tettonico delle
zone esterne dellAppenino Abruzzese: Elementi di analisi stratigrafo-
strutturali, Atti Ticinesi Sci. Terra, 2, 543, 1994.
Ghisetti, F., D. L. Kirschner, L. Vezzani, and F. Agosta, Stable isotope
evidence for contrasting paleofluid circulation in thrust and seismogenic
normal faults of central Apennines, Italy, J. Geophys. Res., 106, 8811
8825, 2001.
Giraudi, C., Elementi di geologia del Quaternario della piana di Campo
Imperatore, in Atti Ticinesi di Scienze della Terra, vol. 2, pp. 137143,
Pavia Univ. Press, Pavia, Italy, 1994.
Hanemberg, W. C., P. S. Mozley, J. C. Moore, and L. B. Goodwin (Eds.),
Faults and Subsurface Fluid Flow in the Shallow Crust, Geophys.
Monogr. Ser., vol. 113, AGU, Washington, D.C., 1999.
Jenkyns, H. C., A. S. Gale, and R. M. Corfield, Carbon- and oxygen-
isotope stratigraphy of the English chalk and Italian Scaglia, and its
paleoclimate significance, Geol. Mag., 131, 134, 1994.
Jolivet, L., et al., Mid-crustal shear zones in postorogenic extension: Ex-
ample from the northern Tyrrhenian Sea, J. Geophys. Res., 103, 123
12,160, 1998.
Jones, G., , Q. J. Fisher, and R. J. Knipe (Eds.), Faulting, Fault Sealing,
and Fluid Flow in Hydrocarbon Reservoirs, Geol Soc. Spec. Publ., 147,
1998.
Kirschner, D. L., J. P. Encarnacion, and F. Agosta, Incorporating stable-
isotope geochemistry in undergraduate laboratory courses, J. Geosci.
Educ., 48, 209215, 2000.
Knipe, R. J., S. M. Agar, and D. J. Prior, The microstuctural evolution of
fluid flow paths in semi-lithified sediments from subduction complex,
Philos. Trans. R. Soc. London, 335, 261273, 1991.
Locardi, E., The origin of the Apenninic Arcs, Tectonophysics, 146, 105
123, 1988.
Malinverno, A., and W. B. F. Ryan, Extension in the Tyrrhenian Sea and
shortening in the Apennines as result of arc migration driven by sinking
of the lithosphere, Tectonics, 5, 227245, 1986.
Mantovani, E., D. Albarello, C. Tamburelli, and D. Babbucci, Evolution of
the Tyrrhenian basin and surrounding regions as result of the Africa-
Eurasia convergence, J. Geodyn., 21, 3572, 1996.
Marone, C., The effect of loading rate on static friction and the rate of fault
healing, Nature, 391, 6972, 1998.
Matthai, S., A. Aydin, D. Pollard, and S. Roberts, Numerical simulation of
departures from radial drawdown in a faulted sandstone reservoir with
joints and deformation bands, in Faulting, Fault Sealing and Fluid Flow
in Hydrocarbon Reservoir, edited by G. Jones, Q. J. Fisher, and R. J.
Knipe, Geol. Soc. Spec. Publ., 147, 157191, 1998.
McCaig, A. M., Deep fluid circulation in fault zones, Geology, 16, 867
870, 1988.
McCrea, J. M., On the isotope chemistry of carbonates and a paleotempera-
ture scale, J. Chem. Phys., 18, 849857, 1950.
Miccadei, E., R. Barberi, and G. P. Cavinato, La geologia Quaternaria della
conca di Sulmona (Abruzzo, Italia centrale), Geol. Romana, 34, 5986,
1998.
Micchetti, A., F. Brunamonte, L. Serva, and E. Vittori, Trench investiga-
tions of the 1915 earthquake fault scarps (Abruzzo, central Italy): Geo-
logical evidence of large, historical events, J. Geophys. Res., 101, 5921
5936, 1996.
Mostardini, F., and S. Merlini, Appennino centro-meridionale: Sezioni geo-
logiche e proposta di modello strutturale, Mem Soc. Geol. It., 35, 177
202, 1988.
Mozley, P. S., and L. C. Goodwin, Patterns of cementation along a Cen-
ozoic normal fault: A record of paleoflow orientations, Geology, 23,
539542, 1995.
Muir-Wood, R., and G. C. P. King, Hydrological signatures of earthquake
strain, J. Geophys. Res., 98, 22,03522,068, 1993.
Oliver, J., Fluids expelled tectonically from orogenic belts: Their role in
hydrocarbon migration and other geologic phenomena, Geology, 14, 99
102, 1986.
Pantosti, D., G. DAddezio, and F. R. Cinti, Paleoseismicity of the Ovin-
doli-Pezza Fault, central Apennines, Italy: A history including a large
previously unrecorded earthquake in the Middle Ages (8601300 A. D.),
J. Geophys. Res., 101, 59375959, 1996.
Patacca, E., R. Sartori, and P. Scandone, Tyrrhenian Basin and Apenninic
arcs; kinematic relations since late Tortonian times, Mem. Soc. Geol. It.,
45, 425451, 1990.
Piccardi, L., Y. Gaudemer, P. Tapponnier, and M. Boccaletti, Active oblique
extension in the central Apennines (Italy): Evidence from the Fucino
region, Geophys. J. Int., 2, 499530, 1999.
Quattrocchi, F., In search of evidence of deep fluid discharges and pore
pressure evolution in the crust to explain the seismicity style of the
Umbria-Marche 19971998 seismic sequence (central Italy), Ann. Geo-
fis., 42, 609636, 1999.
Quattrocchi, F., et al., Geochemical changes at the Bagni di Triponzo ther-
mal spring during the Umbria-Marche 19971998 seismic sequence,
J. Seismol., 4, 567587, 2000.
Rawling, G. C., L. B. Goodwin, and J. L. Wilson, Internal architecture,
permeability structure, and hydrologic significance of contrasting fault-
zone types, Geology, 29, 4346, 2001.
Royden, L., E. Patacca, and P. Scandone, Segmentation and configuration
of subducted lithosphere in Italy: An important control on thrust-belt and
foredeep-basin evolution, Geology, 15, 714717, 1987.
Scholz, C. H., The Mechanics of Earthquakes and Faulting, 439 pp., Cam-
bridge Univ. Press, New York, 1990.
Scholz, C. H., and M. H. Anders, The permeability of faults, U.S. Geol. Sur.
Open File Rep., OF 94-0228, 247253, 1994.
Scholz, C. H., L. R. Sykes, and Y. P. Aggarwal, Earthquake prediction: A
physical basis, Science, 181, 803810, 1973.
Sibson, R. H., Fault rocks and fault mechanisms, J. Geol. Soc. London, 133,
191213, 1977.
Sibson, R. H., Rupture nucleation on unfavorably oriented faults, Bull.
Seismol. Soc. Am, 80, 15801604, 1990.
Sibson, R. H., Implication of fault-valve behavior for rupture nucleation
and recurrence, Tectonophysics, 211, 283293, 1992.
Sibson, R. H., Fluid involvement in normal faulting, J. Geodyn., 29, 469
499, 2000.
Sibson, R. H., J. M. Moore, and A. H. Rankin, Seismic pumpingA
hydrothermal fluid transport mechanism, J. Geol. Soc. London, 131,
639653, 1975.
Vezzani, L., and F. Ghisetti, Carta Geologica dellAbruzzo, Soc. Elabora-
zioni Cartogr., Firenze, Italy, 1998.
Zhang, S., and T. E. Tullis, The effect of fault slip on permeability and
permeability anisotropy in quartz gouge, Tectonophysics, 295, 4152,
1998.

F. Agosta, Department of Geological and Environmental Sciences,


Stanford University, 450 Serra Mall, Stanford, CA 94305-2115, USA.
(bizio@pangea.stanford.edu)
D. L. Kirschner, Department of Earth and Atmospheric Sciences, Saint
Louis University, 3507 Laclede Ave., Saint Louis, MO 63103, USA.
(dkirschn@eas.slu.edu)
AGOSTA AND KIRSCHNER: FLUID CONDUITS IN SEISMOGENIC NORMAL FAULTS ETG 18 - 13

You might also like