You are on page 1of 19

Ital. J. Geosci., Vol. 134, No. 3 (2015), pp. 423-441, 14 figs., (doi: 10.3301/IJG.2014.

20)
© Società Geologica Italiana, Roma 2015

The role of mechanical stratigraphy on normal fault growth


across a Cretaceous carbonate multi-layer, central Texas (USA)
FABRIZIO AGOSTA (*), CHRISTOPHER WILSON (**), (***) & ATILLA AYDIN (**)

ABSTRACT partitioned deformation is responsible for several types of


geologic structures. Actually, it is documented that this
We present the results of an integrated field and laboratory
investigation of the deformation mechanisms associated with nucle- type of deformation affected fold formation (CURRIE et
ation and development of normal faults across heterogeneously alii, 1962; JOHNSON, 1967; AYDIN et alii, 2010), breccia-
layered Cretaceous carbonates. These results shed a new light on the tion (DHOLAKIA et alii, 1998), fault geometry (WOODWARD
role of partitioning of failure structures in the fault growth across & RUTHERFORD, 1989; FISCHER & JACKSON, 1999), and
such a carbonate multi-layer. Structural data show that the oldest
structural elements consist of low-angle to bedding pressure solution fault growth (WILKINS & GROSS, 2002; ANTONELLINI et
seams, which formed during burial diagenesis of limestone rocks, alii, 2008; AGOSTA et alii, 2012).
such as wackestones, containing a large amount of carbonate mud. Bed-partitioned deformation is due to mechanical
Subsequently, we envision that normal fault nucleated due to shear- stratigraphy, which subdivides stratified rocks into
ing of these burial-related structures. Continuous slip along the
sheared low-angle pressure solution seams localized fault-related discrete intervals, named mechanical units, bounded by
dilation in the surrounding stiffer carbonate beds, which mainly mechanical interfaces (cf. GROSS & ENGELDER, 1995;
consisted of packestones and dolostones. In particular, dilation FERRILL & MORRIS, 2008; RUSTICHELLI et alii, 2013).
localized within releasing jogs, as shown by the high frequency of Mechanical stratigraphy is a product of depositional com-
high-angle to bedding splay joints. The splay joints were eventually
sheared during fault development due to fault-related bed tilting. position, texture and structure as well as of the physical,
Linkage of the sheared elements originally compartmentalized chemical and mechanical changes associated to diagene-
within individual beds, which were characterized by variable cut-off sis. Differently, fracture stratigraphy subdivides rocks
angles, determined the observed scalloped vertical trace of the into intervals according to the vertical extent, density, or
through-going normal faults across the carbonate multi-layer.
some other observed attributes of fractures crosscutting a
KEY WORDS: carbonate rocks, mechanical stratigraphy, layered rock mass (LAUBACH et alii, 2009, and references
normal faulting, fault growth. therein). Commonly, both terms are often used as syno-
nyms in literature, but they do not always coincide.
Layered Cretaceous carbonate rocks of central Texas
INTRODUCTION
nicely record the influence of mechanical stratigraphy on
bed-partitioned deformation associated to growth of
It is well known that the varying material properties small offset normal faults, with throws up to about 1m.
cause the none uniform deformation of sedimentary Previous works carried out in the area documented the
multi-layers when subjected to burial or tectonic stress. A impact of faults and fractures on fluid flow. It has been
cursory glance at measurements from rock mechanics shown how fractures strongly control the reservoir per-
experiments on sedimentary rocks displays a wide vari- meability in the Austin Chalk Formation (COLLINS et alii,
ability of their material properties. In fact, tensile 1992; LAUBACH et alii, 1995; TAYLOR & GRANT, 2004), and
strengths range from 2 to 40 MPa, compressive strengths faults pronouncedly influence groundwater flow in the
from 30 to 350 MPa (BIENIAWKSI, 1984), and the elastic Edwards Aquifer (FERRILL et alii, 2004).
Young’s moduli from about 1 to 100 GPa (ESCHBAC, In this paper, contrasting failure modes are docu-
1961). The confining stress at which the brittle to ductile mented for normal faults that crosscut the Cretaceous
failure occurs range between 30 and 300 MPa (JAEGER et carbonate multi-layer. The results of combined field and
alii, 2007). Coupling these wide ranges of material prop- laboratory analyses show the control exerted by the
erties with the heterogeneity and anisotropy introduced mechanical rock properties on deformation. The two
by bedding interfaces and sedimentary structures such as investigated vertical outcrops expose normal faults with
ripple marks, hummocky cross-bedding stratification, scalloped vertical traces. The working hypothesis is that
bioturbation and tepees, it is of little surprise that bed- the carbonate mechanical stratigraphy influenced the
faulting process and, hence, impacted the resulting fault
architecture. To support this hypothesis, the geometry
and distribution of opening-mode (joints) and closing-
(*) Dipartimento di Scienze, Università degli Studi della mode fractures (pressure solution seams) are investigated
Basilicata, Potenza, Italy. Telephone number 0971-206176. Corre- throughout the outcrops to determine whether certain
sponding author e-mail: fabrizio.agosta@unibas.it lithologies contain predominantly fractures of one type or
(**) Geological and Environmental Science Department, Stan- the other. Then, a conceptual model of normal fault
ford University, Palo Alto (California), USA.
(***) Currently at Chevron Corporate Upstream & Gas, Hous- growth across the carbonate multi-layer is proposed in
ton (Texas), USA. terms of the initial failure structures.
424 F. AGOSTA ET ALII

TERMINOLOGY fractures (RISPOLI, 1981; BATHURST, 1987; PETIT & MAT-


TAUER, 1995).
Since jointing in the carbonates of Texas has been Pressure solution structures exhibit a variety of mor-
well documented in the past (e.g., TAYLOR & GRANT, phologies. Stylolites, perhaps the most recognizable mani-
2004), a greater effort has been devoted to establish the festation of one type of pressure solution often localized
wide spread occurrence of pressure solution seams (pss). along bedding, appear as jagged and serrated bands
Except few notes (e.g., CHANCHANI et alii, 1976) and two defined by insoluble material, mostly clay minerals
recent publications (WILSON et alii, 2011a,b), little is (STOCKDALE, 1922). These features often cut across
known about occurrence of pressure solution in these allochems, cement, and matrix and are commonly ori-
rocks. The role of a pressure solution-based mechanism ented parallel or sub-parallel to bedding (STOCKDALE,
on faulting has been described from other parts of the 1922; BATHURST, 1987; RUSTICHELLI et alii, 2012). Stylo-
World. For instance, the shearing of multiple generations lites are equivalent to what WANLESS (1979) refers to as
of solution seams appears to be responsible for the for- sutured-seams. However, pss do not always appear as
mation of normal faults in the Majella Mt. of central Italy conspicuous as stylolites. They also occur as relatively
(GRAHAM et alii, 2003). In Mesozoic platform carbonates smooth, undulated seams of insoluble clay-rich residue
of central Italy, normal faults initiated by shearing of pre- which lack the obvious sutures or the jagged morphology
existing solution seams and formation of new sets of solu- that characterizes stylolites. In the past, researchers used
tion seams and joints/veins (AGOSTA & AYDIN, 2006). various terms to refer to these smooth, undulated seams;
Strike-slip faults found in shallow marine carbonates such terms include non-sutured seams (WANLESS, 1979;
near Somerset, UK appear to result, in large part, through CHANCHANI et alii. 1996), solution seams (GARRISON &
the coalescence of solution seams and calcite veins KENNEDY, 1977), microstylolites (MARSHAK & ENGELDER,
(WILLEMSE et alii, 1997; PEACOCK & SANDERSON, 1995). 1985), and dissolution seams (BATHURST, 1987). In this
SALVINI et alii (1999) documented that the lateral propa- paper, we will refer to all types of pressure solution-
gation of the Mattinata strike-slip fault in Mesozoic plat- related features simply as pss. However we will note evi-
form carbonate rocks, exposed in southern Italy, occurred dence leading to our interpretation of their origin, such as
thanks to pressure solution processes that localized. stylolitic sutures, clay mineral infillings, or truncated and
ANTONELLINI et alii (2008) reported three types of failure interpenetrating grains or fossils.
modes, pss, deformation bands, and joints in Meso-Ceno- In addition to exhibiting a variety of morphologies,
zoic slope and basinal carbonates of the Majella Mt. of pss may influence the fabric of the rocks that contain
central Italy. Coeval formation of two conjugate sets of them. For instance, pervasive pressure solution accompa-
strike faults and two conjugate sets of normal faults nied by mechanical compaction during late stage diagen-
occurred, in concomitance to fold-related reverse fault- esis may alter carbonate beds into small bodies of rela-
ing, by mean means of pressure-solution based mecha- tively pure carbonate encapsulated by clay-rich undulated
nisms as documented by AYDIN et alii (2010) in the Meso- pss (GARRISON & KENNEDY, 1977). When the encapsu-
zoic platform carbonates of the Majella Mt., central Italy. lated carbonate bodies take on an ellipsoidal or lenticular
One of the premises of this investigation is that pres- shape, the rock is said to exhibit flaser fabric or flaser
sure solution is one of the most dominant failure mecha- structure (GARRISON & KENNEDY, 1977). If, however, the
nisms. For this reason, this process and its structural prod- carbonate bodies appear to be rounded rather than long
ucts require a specific introduction. The term pressure and lenticular, the beds are referred to as a nodular
solution describes a three step process that includes the (SCHOLLE, 1977; WANLESS, 1979; GARRISON & KENNEDY,
dissolution resulting from grain interpenetration at grain 1977; BATHURST, 1987; FLÜGEL, 2004). Though largely
contacts subject to high effective stress, which is followed attributed to pressure solution, some nodular fabrics may
by the diffusion/dispersion of the dissolved material, and also arise from depositional processes (FLÜGEL, 2004).
then the precipitation of the dissolved material in locations The key for identifying the pressure solution genesis is
of lower effective stress (WEYL, 1959; GROSHONG, 1988). the content and morphology of the structure.
Pressure solution may either occur pervasively throughout
a body of rock, characterized by distributed dissolution at
grain contacts (HEALD, 1956) sometimes referred to as “fit- GEOLOGICAL AND STRUCTURAL SETTINGS
ted fabric” (BUXTON & SIBLEY, 1981), or it may localize
along discrete surfaces or bands known as pressure solu- During the Cretaceous period, the shallow seas that
tion seams (STOCKDALE, 1922). The process of pressure extended across central and west Texas, USA, provided a
solution results in a permanent strain of rocks and is depositional environment for the formation of a carbo-
therefore indicative of an inelastic, or creep, deformation nate sequence that crop out across much of the state
(RUTTER, 1983). However, when pressure solution local- (BARKER et alii, 1994; fig. 1a). The Cretaceous, shallow-
izes along discrete tabular bands forming pss, its mecha- water, marine carbonates are made up of alternated lime-
nical failure mode is more accurately described as a stone, dolostone, calcareous shale, chalk and marl units.
closing-mode fracturing (FLETCHER & POLLARD, 1981). The two study outcrops expose a section of the Upper
Conceptualizing pressure solution seams as closing-mode Glen Rose and Buda Formations, respectively (figs. 1b
fractures (also coined as anticracks, localized volume and 1c). These two roadcuts are hundreds of kilometres
reduction structures or mode-IV fractures) proves useful apart, and display Tertiary normal faults formed at
when evaluating their associated stress fields (FLETCHER & crustal depths not greater than 2,3 km (FULLMER &
POLLARD, 1981; KATSMAN et alii, 2006; FOSSEN, 2010; LUCIA, 2006).
ZHOU & AYDIN, 2010). Moreover, the aforementioned fail- The Glen Rose Formation comprises the upper unit of
ure mode interpretation is the key to decipher the relation- the Cretaceous Trinity Group. Many researchers attribute
ships between pss with secondary structures such as splay the deposition of the Glen Rose to a Late Aptian and
THE ROLE OF MECHANICAL STRATIGRAPHY 425

Fig. 1 - Location map of the study areas and stratigraphic sections: (a) Digital elevation map of Texas. Areas where Cretaceous aged rocks
crop out are shaded green. The outline of the subsurface Ouachita orogenic belt is in blue. The Trans-Pecos region of Texas is in brown.
Several regionally mapped normal faults within the Balcones Fault Zone and in the Trans-Pecos region are traced in black. The two study
locations (1 and 2) included in this work are labelled in the order they appear in the text. Selected cities referred to in the text are marked
with black circles; (b) Cretaceous stratigraphic column showing the thickness of the formations in the San Antonio/Austin area (after
COLLINS, 2000); (c) Cretaceous stratigraphic column for the Trans-Pecos region (after BRAND & DEFORD, 1958). Arrows between (b) and
(c) indicate time equivalent correlations.

Early Albian regional high stand (BEHRENS, 1965; KE- 2007). A regionally identified corbula marker separates
RANS, 2005; STRICKLIN & AMSBURY, 1971; YOUNG, 1974). the Glen Rose Formation into upper and lower intervals
Stratigraphic relationships show that the Glen Rose gen- (STRICKLIN & AMSBURY, 1971). Each one of these two
erally overlies and interfingers with the Hensel Sand- intervals can be also subdivided into several lithostrati-
stone, and underlies the Edwards Group (FERRILL & graphic units (WARD & WARD, 2007).
MORRIS, 2008). Based upon the fossil content of specific The other study location is in the Cretaceous Buda
beds (STRICKLIN & AMSBURY, 1971) and distinctive well Formation, which crops out extensively in south, central,
log signatures (BEHRENS, 1965; PITTMAN, 1989), the Glen and western Trans-Pecos Texas as well as northern Chi-
Rose Formation is believed to extend from western cen- huahua and Coahuila, Mexico (REASER & DAWSON,
tral-south Texas (1067 m-thick) to the adjacent territories 1995). It serves as an oil and gas reservoir in south Texas,
of Arkansas and Louisiana (610 m-thick). Its exposed where production is generally associated with intense
thickness of the Glen Rose Formation in the Austin and brecciation and fracturing in the upper part of this for-
San Antonio region, the current study area, varies mation (STAPP, 1977). REASER & DAWSON (1995) docu-
between 225 m (WARD & WARD, 2007) and 270 m mented a thickness variation of the Buda Fm. between
(COLLINS, 2000). The Glen Rose Formation consists of less than 1 m (northeast Texas) and 107 m (western
vertically stacked dolostones, limestones, clays, marls and Trans-Pecos Texas). In the latter area, the Buda Forma-
evaporates, consistent with an open-marine, shallow sub- tion is subdivided into six informal lithofacies, which are
tidal-peritidal, restricted inner shelf depositional environ- sandwiched in between the Boracho and Boquillas For-
ment (STRICKLIN & AMSBURY, 1971; WARD & WARD, mations. These informal lithofacies include fossiliferous
426 F. AGOSTA ET ALII

sparites, sandy micrites, a range of biomicrites, and faults and the general subsidence of the Gulf Coast
interbedded shale intervals (REASER & DAWSON, 1995). (GALLOWAY, 2000); the normal faults of the Balcones
Although the two study outcrops are found in differ- Fault Zone formed at the boundary of the subsiding gulf
ent tectonic settings, they each expose various types of coast and the uplifting central United States craton.
fractures and small normal faults crosscutting Cretaceous Evidence supporting this hypothesis includes reworked
carbonate rocks. One of these outcrops (figs. 2a and 2b) Cretaceous faunas and chert cobbles found within Ter-
lies within the Austin-San Antonio corridor of the Bal- tiary sediments along the Gulf Coast, which suggests that
cones Fault Zone, about 32 km northwest of Austin, the majority of displacement across the Balcones normal
Texas (USA), where the total thickness of the Cretaceous faults occurred during the Oligocene-Early Miocene
section is approximately 1045 meters (COLLINS, 2000). (WEEKS, 1945; EWING, 1987).
The second outcrop is located within the Trans-Pecos The development of normal faults in the Trans-Pecos
region of Texas (figs. 2c and 2d), 8.7 km northwest of the region, situated west of the Edwards Plateau of Texas
city of Balmorhea, Texas (USA). There, the 307 m-thick (cf. fig. 1a), exhibits some differences from those in the
exposed Cretaceous section reflects a depositional Late Balcones Fault Zone to the northeast. As it lies closer to
Triassic-Early Cretaceous hiatus (BRAND & DEFORD, the central part of the North American craton, the
1958; LEHMAN, 1986). Trans-Pecos geology reflects the influence of both
Despite the aforementioned depositional hiatus, the Laramide orogeny and Basin and Range rifting (HENRY
Mid-to-Late Cretaceous carbonates are believed to extend & PRICE, 1986). From the late Cretaceous to the early
from the Austin-San Antonio corridor to the Trans-Pecos Tertiary, the Laramide orogeny mildly compressed rocks
region (about 640 km west) as an essentially continuous in the Trans-Pecos region, leading to the formation of a
body of rock (BRAND & DEFORD, 1958). Several stra- series of mountain ranges and subtle anticlines and syn-
tigraphic contacts are correlated between these two clines (EIFLER, 1951). It has also been surmised that
regions. For instance, the Upper Glen Rose Formation mild Laramide compression affected central Texas dur-
found in central Texas is time equivalent to the Yearwood ing the same period (LAUBACH & JACKSON, 1990). In the
Formation in the Trans-Pecos region. Furthermore, the late Tertiary, Laramide compression gave way to Basin
unconformable contact between the Buda Limestone and and Range extensional tectonics (BARKER, 1987). Many
Eagleford Shale in central Texas is the same as that of the Trans-Pecos region’s normal faults formed during
between the Buda and the Boquillas Formations in the this time, probably in the Oligocene (EIFLER, 1951;
Trans-Pecos region (see arrows connecting figs. 1b to 1c). MUEHLBERGER et alii, 1978). The surface pattern of
As stated above, even if the geologic history of two fault traces reveals both north and northwest to west
study fault zones is quite different, the normal faults in trending segments, which occupy a 100 km-wide belt
the Balcones Fault Zone and those in the Trans-Pecos running roughly parallel to the southern border of
region both formed at similar shallow depths during the Texas. The former fault set is parallel to Late Paleozoic,
Late Tertiary. Additionally, the initiation of each normal Mesozoic and/or Laramide faults, while the latter one
faulting system appears to be influenced by pre-existing trends parallel to Paleozoic and Precambrian faults
structures. The Balcones Fault Zone crosses the state of (KING & FLAWN, 1953).
Texas from the city of Del Rio to the city of Dallas (cf.
fig. 1a). This broad curved belt follows the pre-existing
structural assemblage of compressed Paleozoic rocks that METHODOLOGY
composes remains of the now eroded Ouachita orogenic
belt, which marks the southern margin of the central The present work entailed detailed field mapping as
United States craton (CONDON & DYMAN, 2006). Physio- well as petrographic and scanning electron microscope
graphically, the Balcones Fault Zone separates the analyses of selected samples. In the field, the exposed
Tertiary clastic sediments that overlie the Gulf Coastal fractures and faults were mapped at a variety of scales.
Plane from the predominantly carbonate sediments now The attitudes of these structures were measured and, in
exposed in the Edwards Plateau (Texas, USA). From SW some locations, their spacing was determined by mean of
to NE, the dominant fault strike rotates counter-clock- scanlines oriented perpendicular to their strike direction.
wise from an approximately east-west direction, near Del The twenty-two representative samples collected from
Rio, to a nearly north-south direction, near Dallas. In gen- specific outcrop locations, mainly of site 1, were then
eral, the dip directions of the normal faults throughout impregnated with blue epoxy, cut into 2.5×5 cm thin
the Balcones Fault Zone show presence of a conjugate sections, polished, and then stained with alizarin red
system made up of SE and NW dipping fault sets, respec- solution to distinguish limestone from dolomite. Petro-
tively. At surface, the aforementioned fault zone ranges graphic analysis was completed with a Zeiss Avio Plan 2
in width between 25 and 30 km (WILTSCHKO et alii, 1991; Optical Microscopes (equipped with reflected, transmit-
FERRILL et alii, 2004; TAYLOR & GRANT, 2004). Cumulative ted, and fluorescent light sources). An Axio Cam MRS
fault throws appear to be greatest, more than 500 meters, digital camera captured the petrographic images pre-
in the vicinity of the San Antonio and Austin cities (COLE sented in this study. In addition to the petrographic sam-
& PANTEA, 2004). ples, seventeen samples were impregnated with blue
The mechanisms leading to the initiation and growth epoxy, cut into 2.5×5 cm thin sections, polished with a
of the faults within the Balcones Fault Zone remain diamond paste, carbon coated and analysed using a JEOL
unclear. Some researchers proposed that their formation JXA-733A Super-probe scanning electron microprobe.
is associated with a period of regional extension induced SEM backscatter images were used to calculate rough
by a rapid accumulation of Tertiary sediments along the approximations of rock composition through a three step
Gulf Coast. That is, the rapid accumulation of Tertiary process. First, the elemental compositions of the rock
sediments lead to the formation of salt related growth constituents (various grains, matrix, fossils, and fossil
THE ROLE OF MECHANICAL STRATIGRAPHY 427

Fig. 2 - Locations of the study Glen Rose and Buda roadcuts: (a) Partial digital elevation map of central Texas; (b) Geologic map of the area
surrounding the roadcut of site 1. The roadcut location, indicated by a white circle, lies at UTM 14R coordinates 603771 east, 3363281 north.
The surface geology of the surrounding area is coloured according to the legend below the map (adapted from PROCTOR et alii, 1981);
(c) Partial digital elevation map of Trans-Pecos Texas; (d) Geologic map of the area surrounding Balmorhea (from MCKALIPS & BARNES,
1981). The roadcut location, indicated by a white circle, lies at UTM 13R coordinates 601573 east and 3434604 north.

fragments) were determined. Then the hue, saturation, RESULTS OF FIELD AND LABORATORY ANALYSES
and value (HSV) combinations corresponding to these
constituents were categorized for a given backscatter In this section, the results of combined field mapping
image. Lastly, the surface areas of the constituent cate- and laboratory analyses conducted on carbonates pertain-
gories were calculated. Such calculations provide a first ing to the Upper Glen Rose and the Buda Formations.
order approximation to be used for qualitative and rela- The former rocks are exposed 32 km northwest of down-
tive comparisons. town Austin, along a roadcut of the Route 620 located in
428 F. AGOSTA ET ALII

Fig. 3 - Overview of the Upper Glen Rose Formation. Photo views are to the north: (a) and (b) are mapped photomosaics upon which Fault
Set 1 and Set 2, as well as the boundaries of a jointed dolostone (in white), are traced. Locations of collected samples are indicated with
enumerated white circles. The lower inset shows the relative positions of (a) and (b) on a photo of the roadcut; (c) Equal-angle stereonet plot
of measured planes. The great circles show the average attitude of the two fault sets and joints. See fig. 3a for the photograph location.

proximity to the Lake Travis (site 1, figs. 2a and 2b). The dolostone and wackestone beds exhibit composi-
There, in correspondence of the SW-NE trend of one of tional and microstructural differences. Petrographic and
the regionally mapped normal faults pertaining to the SEM analyses of samples taken from a wackestone bed
Balcones Fault Zone (fig. 2b), the E-W oriented outcrop is (# 2, 3 and 4 in figs. 3 and 4) reveal a wide array of fossils,
up to 13 m-high and stretches for about 140 m. mainly miliolid foraminifera and peloids, interspersed
The second site is located nearly 650 km west of the within a micritic (lime mud) matrix, which may contain
Balcones Fault Zone, in the Trans-Pecos region of Texas some euhedral rhombohedra dolomite crystals. The
(fig. 2c), 8.7 km northwest of the city of Balmorhea results of backscatter image analysis are consistent with
along the interstate I-10 (fig. 2d). The roadcut, which the wackestone being composed of about 56.8% matrix
shows a 180 m-long, gently SE-dipping, less than 10°, (mainly calcium carbonate with some remnants of potas-
portion of the Buda Formation, lies in a horst bounded sium feldspars and clay), 19.5% fossil content, 14.2%
by two NW-trending normal faults (MCKALIPAS & dolomite, 1.5% pyrite and 8% porosity. Both the sutured
BARNES, 1981). boundaries of fossils and the sutured contacts between
adjacent fossil grains (figs. 5a and b) indicate that this
UPPER GLEN ROSE FORMATION rocks underwent pressure solution. Furthermore, under
the microscope the clay seams with flaser fabric truncate
At site 1, series of vertically stacked, alternating partial miliolid foraminifera. In fact, the examples of
lithologies of a few dolostone beds embedded within figs. 5a and b show missing portions of foraminifera
wackestones are exposed (figs. 3 and 4). The friable chambers (FLÜGEL, 2004), providing some evidence that
wackestones exhibit an extensive flaser fabric, which the flaser fabric in the wackestone beds resulted from
imparts a distinctive undulating apparent layering. The pressure solution and the associated seams.
finely crystalline dolostone beds stand apart from the Differently, euhedral rhombohedra dolomite crystals
wackestones due to their finer grain size and, generally, compose the majority of the dolostone beds and appear to

The dolomite rhombohedra are about 10 to 20 μm-long,


absence of sedimentary structures. Much of the dolostone have replaced a carbonate mud matrix (figs. 5c and 5d).
beds are stained brown, inferably through the oxidation
of groundwater flow. Oxidation stains are found primar- and are generally in grain-to-grain contact with each
ily towards the top of the roadcut as well as in proximity other. No identifiable fossils are found in the dolostone
to normal faults. samples. Sparse, small and sub-angular quartz and pyrite
THE ROLE OF MECHANICAL STRATIGRAPHY 429

Fig. 4 - Mapped photograph of the Upper Glen Rose Formation. View is to the north. Fault Set 1 and Set 2 are traced as well as joints (in
blue). The latter are found primarily within the dolostone layers. The boundaries of the most heavily jointed dolostone bed is traced in white.
See fig. 3a for the photograph location.

grains are present. The results of backscatter image that of Set 1 and Set 2 faults, their average dip angle, 88°,
analysis are consistent with the following average compo- is much steeper (fig. 3c).
sition: 55.2% dolomite, 18.3% matrix (calcium carbonate Focussing on the pss, at the meso-scale they exhibit a
with some remnant potassium feldspar), 4.5% quartz, range of morphologies varying from barely visible, thin
0.5% pyrite and 21.5% porosity. wispy lines to clusters up to 3 cm-thick. Their amplitudes
From a structural perspective, several small normal range from 1 to 5 cm, while the wavelengths are com-
faults (figs. 3a and 3b) and fractures (fig. 4) are visible prised between 5 and 10 cm. Length of individual seg-
along the vertical roadcut face. The faults strike about ments of pss is comprised between 2 and 8 cm. Multiple
NE-SW, and dip either SE or NW. Collectively, these pressure solution seam segments may link together form-
faults form an apparent conjugate system made up of ing continuous anastomosing zones sub-parallel to bed-
SE-dipping (Set 1) and NW-dipping faults (Set 2). As ding, similar to what NENNA & AYDIN (2010) documented
measured in the wackestone beds, the average attitude of in the Variscan of SE Ireland. Some of these bed-parallel
Set 1 is N33E and 55°, while that of Set 2 is N20E and 47° zones are traceable for several meters, often anastomos-
(fig. 3c). In addition to the small faults, numerous frac- ing and enclosing, either partially or completely, lens-
tures are documented in the wackestone beds. These frac- shaped and ellipsoidal limestone bodies. In other loca-
ture, which can also be categorized into Set 1 and Set 2 tions, multiple segments of different seams may link
groups based upon their attitudes (fig. 4), do not exhibit a together at low angles to bedding forming composite
plumose morphology. Their traces are associated with the seams that dip, either SE or NW, between 20° and 45°.
clay-rich seams that define lenticular bodies of rock These segments show mutually crosscutting/abutting rela-
(flaser structure). Differently, the dolostone beds contain tionships (fig. 6). Such composite seams exhibit a similar
numerous fractures sub-perpendicular to bedding (fig. 4). attitude of the Set 1 and Set 2 fractures (cf. fig. 4). For
These fractures, joints because characterized by plumose this reason, they are interpreted as linked pss.
structures along their surfaces, are primarily located in Microprobe transects further corroborate the afore-
the relatively thin (less than 40 cm-thick) beds. Although mentioned interpretation. In fact, an SEM backscatter
the average joint strike, N25E, approximately aligns with image reveals a high resolution view of some of the seams
430 F. AGOSTA ET ALII

Fig. 5 - Petrographic and backscatter SEM images of the samples of the Upper Glen Rose cropping out at the 620 roadcut near Austin, Texas,
USA. Images in (a) and (b) are from a wackestone bed while (c) and (d) are from a dolostone bed (sample locations are labeled 2 and 1,
respectively, in fig. 3a): (a) Transmitted light photomicrograph with several labeled items: a peloid (pd), the sutured edge (sb) of a miliolid
foraminifera (mf), a pressure solution seam (pss) that composes part of an outcrop observed flaser fabric, and a sutured grain (sg);
(b) Backscatter electron image showing a fitted contact (fc) between a bivalve test (bv) and a miliolid foraminifera (mf) whose edges are
sutured (sb), a pressure solution seam (pss) truncating the foraminifera, and a dolomite rhomb (dm); (c) Transmitted light photomicrograph
showing the fine crystalline dolomite crystals; (d) Backscatter electron image from a dolostone bed. Several items are pointed out: the matrix
(m), a quartz grain (q), a dolomite rhomb (dm), and a pyrite grain (p).

dark material about 10 μm-thick and up to 100 μm-long.


that comprise the flaser fabric (fig. 7a). The pss appear as results of transects of relative element counts show that
the pss contain a smaller amount of calcium and a higher
The micro-scale image also shows presence of the differ- amount of insoluble clay-related elements, such as silica
ent pressure solution sets, the most common of which and aluminium, than the surrounding rock (figs. 7b
appears as sub-parallel to bedding. However, pss oriented and 7c). This identification is similar to those docu-
like the Set 1 and Set 2 fractures are also visible. The mented in a number of locations worldwide (GARRISON
THE ROLE OF MECHANICAL STRATIGRAPHY 431

Fig. 6 - Mapped photograph of flaser fabric in a wackestone bed


within the Upper Glen Rose Formation. View is to the north. Ellip-
soidal limestone bodies partially or fully encapsulated by clay seams
form flaser fabric (or flaser structure). In this case, parts or
segments of the pressure solution seams exhibit dip directions
consistent with the Set 1 and Set 2 fractures and faults. See fig. 3a
for the photograph location.

& KENNEDY, 1977; WANLESS, 1979; BATHURST, 1987;


AGOSTA et alii, 2012).
Investigating the relationship between the flaser fab-
ric and the normal faults, many of the Set 1 and Set 2 pss
align parallel, and in close proximity, to nearby slip sur-
faces (figs. 4 and 8). Additionally, the tips of some indi-
vidual slip surfaces sole out and terminate into the flaser
fabric (fig. 8). This configuration, which indicates that the
slip surfaces either begin or end at those locations, often
include continuous zones of either Set 1 or Set 2 pss
ahead of the fault tips (fig. 8). Locally, either these contin-
uous zones of pss or the fault segments themselves abut Fig. 7 - Micro-probe backscatter electron image and measurements
into networks of the opposite dipping set of pss, forming of the relative element counts across dark seams in the wackestone
an apparent conjugate system (fig. 8). beds. Sample taken from a wackestone bed at location S2 shown in
While in wackestone beds the slip surfaces sub-paral- fig. 3a: (a) Back scatter electron image. The lines 1 to 1' and 2 to 2',
provide the locations of two microprobe transects; (b) Relative
lel the Set 1 and Set 2 pss, they follow the joint attitude in element counts of transect 1 to 1'; (c) Relative element counts of
the dolostone beds (fig. 9). The dip angle of individual transect 2 to 2'. In both cases, relative element counts are specific to
fault segments therefore changes abruptly along the study each transect (that is, a relative concentration of 1 indicates the
outcrop from moderate, ~45° within the wackestone beds, highest elemental count of that particular transect).
to higher values, ~85° in the dolostone beds. This change
of the dip angle is most pronounced in the ~25 cm-thick
dolostone bed at the base of the roadcut, which contains
km2 area (PARKER & MCDOWELL, 1979) whose occur-
very few Set 1 or Set 2 pss, if any. There, the joint spacing
rence was likely related to Basin and Range extension
rapidly decreases in the vicinity of the slip surfaces (fig. 9).
(EIFLER, 1951; MUEHLBERGER et alii, 1978). The study
section of the Buda Formation shows a sharp lithological
BUDA FORMATION variation (fig. 10a). About 2.5 m up from the base of the
roadcut, the bed colours change from interbedded tans,
The study roadcut is located on the easternmost side browns, and blacks to a whitish-gray (fig. 10b). This
of a NW-trending normal fault zone (cf. figs. 2c and d), colour change marks an abrupt transition from grain-sup-
which is sub-parallel to others that crosscut the Gomez ported beds, mainly comprised of fossiliferous packstones
Tuff, a Tertiary aged peralkalic rhyolite that covers a 4300 with calcite cements, to matrix-supported beds made up
432 F. AGOSTA ET ALII

Fig. 8 - Mapped photograph of a fault in a wackestone bed, view is to Fig. 9 - A 3.5 cm throw normal fault crossing from a wackestone at
the north. The indicated fault surface (semitransparent red) solves a the base of the roadcut into a dolostone and then back into a wacke-
few cm of throw. Set 1 and Set 2 pressure solution seams are traced in stone. Its view is to the north. Both Set 1and Set 2 pressure solution
red and yellow, respectively. An arrangement of Set 2 pressure solution seams are found predominantly in the wackestone beds, while the
seams that form an apparent conjugate set with the fault is shown in bed perpendicular joints in the dolostone bed (outlined in black).
semi-transparent yellow. See fig. 3a for the location of this photograph. See fig. 3a for location.

Fig. 10 - Overview of a roadcut through the Buda Limestone. All photograph views are to the south; (a) and (b) consist of photomosaics upon
which several small faults and fractures are mapped. In (b) the double-pointed arrow spans the 1 m of throw across a normal fault. A black
horizontal line at the right of the photomosaic indicates the bedding contact that divides this Buda Limestone exposure into primarily matrix
supported beds and primarily grain supported beds; (c) Equal-angle stereonet plots of the measured planes, the great circles show the
average attitude of the measurements.

of fossiliferous wackestones. Thus, above this transition ably, these faults formed synchronously with the Tertiary
both fossil content and amount of calcite cementation aged, regionally mapped normal faults. None of the faults
appear to abruptly decrease, while the amount of lime- exhibit throws greater than ~1 m. Similar to what it has
mud matrix sharply increases. been done for the site 1, the outcropping faults are
In addition to the aforementioned lithological varia- grouped into two different sets based upon their dip
tion, a series of SE-NW striking normal faults and frac- direction: Set 1 dips SW, while Set 2 dips NE (fig. 10c).
tures prominently transects the study exposure. Presum- This naming convention does not imply that the sets
THE ROLE OF MECHANICAL STRATIGRAPHY 433

defined here share any genetic relationships with the pre-


viously defined sets (cf. Section 5.1). Set 1 and Set 2
faults shows average dip angles of 71° and 76°, respec-
tively (fig. 10c). The majority of the shallower dipping
faults and fractures are located above the sharp litholo-
gical variation described above.
A closer inspection of one of the wackestone beds
reveals that this fossil-rich, hard limestone contains a
well-developed flaser fabric (figs. 11a and 11b). In contrast
to the pss associated to the flaser fabric identified in the
Upper Glen Rose Formation, the seams present in the
Buda Formation readily present themselves as products of
pressure solution. In fact, not only they contain a conspic-
uous clay-rich, orange and brownish, oxidized material,
but also the stylolitic sutures give them jagged edges easily
identifiable at both meso- and micro-scales (fig. 11c).
As the seams encapsulate irregularly shaped limestone
bodies, they often truncate fossil grains. The trace of
the flaser fabric documented in the wackestone beds is
consistent with most of the pss dipping either SW or NE
(fig. 11a). The results of thin section analysis show that
stylolitic sutures formed perpendicular to bedding,
regardless if the individual seams lie parallel (stylolites)
or at an angle to bedding (slickolites).
Considering the flaser fabric, the ellipsoidal lenses

from less than 1 μm to about 3 cm. Though this was an


often contain calcite filled veins. Vein aperture ranges

unexpected finding, it is not unprecedented to find veins


within flaser or nodular limestone rocks. In fact, veins of
similar appearance crosscut the flaser biomicrite beds of
the Chambersburg Limestone, western Maryland, USA
(WANLESS, 1979), and the Ireton Formation, Alberta,
Canada (MCCROSSAN, 1958). Veins are mainly oriented at
a high-angle to bedding, and commonly abut at right
angles against pss (fig. 11b). Similar orthogonal associa-
tions have been reported by NELSON (1981), AGOSTA &
AYDIN (2006), GRAHAM-WALL et alii (2006), AGOSTA et alii
(2009), KATSMAN et alii (2010), and ZHOU & AYDIN
(2010). Along portions of the investigated outcrop, small
groups of individual veins form sub-vertical en-echelon
arrangements. Within this configuration, the pss link
together sequences of either right or left stepping veins,
forming single composite segments of closing- and open-
ing-mode fractures (fig. 11b). The wackestone beds there-
fore contain three types of fractures, which display a vari-
ety of cut off angles. Pss are at low angles to bedding,
with cut offs smaller than 45°, veins are at high angles to
bedding, with cut offs greater than 70°, and composite
fractures whose cut off angles range between 45° and 70°.
At a larger scale, the traces of the fault segments lie at
orientations either roughly perpendicular to bedding, in
the packstone beds, or at lower angles to it, in the wacke-
stone beds (fig. 12a). Within the packstone beds, other
than a few bed-parallel pss, most of the fractures consist
of calcite-filled veins visible at both thin section (figs. 12b
and 12c) and outcrop scales (fig. 12d). The structural
position of calcite filled releasing jogs, which link Fig. 11 - Images of a wackestone bed within the Buda Limestone
together discrete sheared vein segments, provide evidence (cf. sample 5 in fig. 10). View due to south: (a) Mapped photograph
tracing the two sets of pressure solution seams in the wackestone
that the latter features underwent to, at least, one episode that comprise a flaser fabric: Set 1 and Set 2. Subvertical blue lines
of shearing (fig. 12e). A number of these sheared veins mark traces of veins. View due to south; (b) Photograph showing the
show a similar orientation with nearby fault traces, which flaser fabric defined by pressure solution seams. Arrows mark veins
often exhibit some calcite infilling (fig. 12a). that abut against the bed-parallel (bp pss) and low-angle pressure
solution seams (la pss); (c) Plan polarized light view of a thin section
Observations from a 1 m-throw normal fault, the obtained from a hand specimen of the wackestone bed. The flaser
largest at this exposure, indicate that the individual fault fabric is comprised bed-parallel (bp pss) and low-angle pressure
segments have attitudes similar to either the low-angle solution seams (la pss).
434 F. AGOSTA ET ALII

Fig. 12 - A normal fault and several nearby veins in the Buda Limestone. The upper portion of the cliff is made up of wackestones, the lower
part of packstones: (a) Trace (in yellow) of a 5 cm throw normal fault that vertically transects the Buda exposure. The black line and accom-
panying arrows on the left divide the outcrop into primarily matrix-supported and grain-supported beds. The hammer at the base of the
roadcut for scale. The two double-pointed arrows indicate the beds from which the hand specimens shown as microphotographs in (b) and
(c) were collected; (b) View of a stained thin section under cross-polarized light. The veins crosscut a peloidal packstone filled with sparry
cement. The label points to a vein that splits several peloidal bioclasts. This sample is labelled as S6 in fig. 10; (c) View of a stained thin
section under cross-polarized light showing a sparry cement filled vein within a peloidal-bivalve packstone. This sample is labelled as S7 in
fig. 10; (d) Photograph of sheared joints and associated vein precipitation in their dilational quadrants. The hammer is for scale; (e) Structural
interpretation of the image shown in (d).

bedding planes or the high-angle joints (figs. 13a and b). or within the clay-rich beds. Differently, the fault seg-
The bedding plane-parallel fault segments localize either ments are characterized by dip angles of about 80° in the
at the bed interfaces, which are made up of thin shales, primarily grain-supported beds. There, numerous joints
THE ROLE OF MECHANICAL STRATIGRAPHY 435

Fig. 13 - Mapped photographs of a 1 m throw normal fault crosscutting the Buda Limestone. The upper side of the cliff is made up of wacke-
stones, the lower one of packstones. In both (a) and (b), red and yellow lines mark fault traces that dip consistently with Set 1 and Set 2
faults and pss, respectively, while blue lines follow the traces of veins and joints: (a) Selectively mapped photograph (view due to southeast)
of the normal fault. A yellow field-book for scale; (b) Detailed map of the normal fault near the transition between wackestone (matrix-
supported carbonate) and packstone (grain-supported carbonate) beds.

and veins surround the fault segments. Veins 4 to 6 cm- DISCUSSION


thick are located on either side of the fault segments,
suggesting that much of the shear strain in these rocks Recent works on Cretaceous carbonates of Texas
was accommodated through dilation. The grain-sup- focussed on fracture detection and mapping by using
ported beds also contain sparse, less frequent, Set 1 and Lidar technology (WILSON et alii, 2011a), and on the
Set 2 pss. potential impact of fractures and faults on hydrocarbon
Differently, both Set 1 and Set 2 pss are abundant in flow and migration in an open pit asphalt mine (WILSON
the wackestone bed that lies just above the contact between et alii, 2011b). Previous investigations documented some
the matrix- and grain-supported intervals (figs. 13a and b). expression of the aforementioned mechanical/fracture
Within this bed, the 1 m-throw normal fault abruptly stratigraphy within the same carbonate rocks. CORBETT et
changes its dip angle, down to 64°, and is mainly associ- alii (1987) measured the uniaxial compressive strength
ated to low-angle pss and composite fractures. Specifi- and magnitudes of strain at failure (a measurement of
cally, the fault segment pertaining to the 1 m-throw fault ductility) of Austin Chalk samples collected from out-
is sub-parallel to the Set 1 pss, while some Set 2 pss abut crops, and identified the measurable rock properties that
against this fault segment forming an apparent conjugate appeared to influence the fracture distribution. In fact,
system (fig. 13b). these authors found that samples from heavily jointed
436 F. AGOSTA ET ALII

outcrop intervals tended to be the most porous, contained rants of mode-II propagating shear fractures. To corrobo-
the least amount of clay, and exhibited the least amount of rate this interpretation, it is noted that the number of
strain prior to failure. CHANCHANI et alii (1996) studied joints within and around the releasing jogs tends to rise
thin sections obtained from cores of the Austin Chalk, with increasing fault slip.
which were extracted from 2000-3000 meters below the In light of the aforementioned fault-related jointing of
ground surface, and documented localized presence of the most competent beds, and interpreting the evolving
pressure solution seams within its shale-rich intervals. fault segments as original sheared pss localized in the less
FERRILL & MORRIS (2003) suggested that a material prop- competent beds (wackestones), the latter features are
erty described as competence determines the mode of fail- interpreted as the oldest structural elements for a few rea-
ure in some carbonate rocks. These authors inferred that sons. First, where they connect, joints abut against the
the scalloped vertical trace of normal faults is due to a pss, which implies that the former features are younger.
combination of shearing-mode fracturing of the less com- Also, as stated above, the number of high-angle joints
petent beds, resulting in fault segments with dip angles within the most competent beds increases in proximity of
lower than 80°, and mixed shearing- and opening-mode the Set 1 or Set 2 pss present in adjacent wackestone
(hybrid) fracturing of the more competent beds, causing beds; joint formation appears hence dependent upon pss.
fault segments with dip angles greater than 80°. FERRILL Lastly, Set 1 and Set 2 pss are found within the wacke-
& MORRIS (2008) suggested that the dip angle of the fault stone beds without a significant presence of nearby joints,
segments pertaining to normal faults of the Balcones Fault which implies that both Set 1 and Set 2 pss formed inde-
Zone depend upon clay content of carbonates. Fault seg- pendently of joints.
ments in clay-poor limestones, or dolostones, form at high The documented relative timing of formation of the
angles greater than 70°, while those in clay-rich limestones two different fracture types, respectively joints within the
at low angles smaller than 60°. MORRIS et alii (2009) fur- most competent beds and pss in the less-competent ones,
ther investigated this topic, finding a relationship between is not what a common experience would suggests for
the competence property of beds (measured with a fault zones that crosscut a mechanically layered sequence
Schmidt hammer) and both spacing and displacement of of beds. In fact, when a multi-layer rock is subject to
normal faults. As already mentioned in the text, LAUBACH strain, faults first generally form in the more competent
et alii (2009) investigated both fracture and mechanical beds. Then, a further strain causes fault development
stratigraphy in the carbonates of Texas documenting how across the whole sequence (PEACOCK & SANDERSON,
they coincide in the Travis Peak Formation, but not in the 1991; CHILDS et alii, 1996; WALSH et alii, 1998, 2003;
Austin Chalk Formation. SCHÖPFER et alii, 2006, 2007; ROCHE et alii, 2012). Differ-
Although the two study sites expose different Creta- ently, in the study case we document an earlier formation
ceous carbonates at entirely different regions, they do of pss, a time-dependent slow process, in the less compe-
exhibit some similarities. Both formations comprise some tent beds , and the subsequent jointing, a dynamic brittle
beds (made up of dolostones and packestones) that process, of the most competent ones.
include more joints than pss due to predominant open- As noted in Section 5.1, both Set 1 and Set 2 pss
ing-mode brittle deformation, and others (made up of appear intimately related to the development of the flaser
wackestones) that contain more pss and flaser fabric than fabric within the wackestone beds. In general, flaser
joints due to closing-mode inelastic, or ductile, deforma- structures develop an ellipsoidal fabric during late stage
tion. The latter wackestone beds often include two sets of diagenesis through mechanical compaction and associ-
low-angle pss. In light of these resemblances, in this sec- ated pressure solution (GARRISON & KENNEDY, 1977). For
tion the field and laboratory results shown previously are this reason, we infer that this process took place during
discussed together in terms of faulting processes in a burial diagenesis of the clay-rich limestone rocks, likely
multi-layer comprised of the aforementioned carbonate under a lithostatic load smaller than 50 MPa (FERRILL &
beds. Based upon the assumption that the vertical propa- MORRIS, 2008). Differently, the fault-related jointing,
gation of normal faults was strongly controlled by the which localized in the most competent beds, probably
contrasting properties of contiguous beds, as shown by occurred within an already lithified medium. The modali-
the marked changes in dip angle of individual fault seg- ties of normal fault evolution through competent beds by
ments and presence of releasing jogs accompanying these mean of opening-mode fracturing will be discussed in the
changes, a conceptual model of fault development across next section.
such a multi-layer is then proposed. Even if the field and laboratory data reported in the
previous chapters show many consistencies with the
NORMAL FAULTING OF THE CRETACEOUS CARBONATES OF
results of previous studies, some key differences are high-
CENTRAL TEXAS lighted. The two main study lithologies failed through
two markedly different mechanisms: (i) brittle jointing
The fault architecture documented in this work is (clay-poor limestone and dolostone), and (ii) ductile pres-
consistent with joint localization in the most competent sure solution (clay-rich limestone). These two failure
beds. In fact, most fault segments that transect dolostone mechanisms differ from the shearing-mode and mixed-
and packestone beds form releasing jogs surrounded by mode fracturing proposed in the past, among many others,
joint clusters (cf. figs. 3, 4, 10 and 13) fading away from by TAYLOR & GRANT (2004) and FERRILL & MORRIS
them. The initiation and progressive slip of the normal (2008) to explain the scalloped vertical trace of normal
fault segments within the most competent beds is there- faults crosscutting similar carbonate multi-layers. Fur-
fore associated to pronounced opening-mode fracturing. thermore, since the study normal faults developed by
This process, widely reported in literature along normal linking sheared joints with sheared pss, it looks like that
faults (e.g., DAVATZES & AYDIN, 2003; AGOSTA & AYDIN, dominant control on fault dip is the failure mechanism,
2006), involves localized jointing at the extensional quad- not the rock competency (FERRILL & MORRIS, 2003).
THE ROLE OF MECHANICAL STRATIGRAPHY 437

It is also noted that the results of this work show a bonate rocks (cf. AGOSTA et alii, 2009, and references
positive correlation between clay/carbonate mud content therein), determines formation of splay joints at the
and the occurrence of pressure solution, as previously extensional quadrants of the sheared fractures. Such
suggested by FERRILL & MORRIS (2008). This correlation splay joints are characterized by power-law spacing and
is similar to that recently documented by RUSTICHELLI et length multi-scale distributions (AGOSTA et alii, 2010).
alii (2012) for bed-parallel pss present within the fine- Then, the splay joints dilate and propagate sub-perpen-
grained packstones of the Bolognano Formation (Majella dicular to bedding while slip continues to accumulate
Mt., central Italy). The latter authors documented how along the sheared pss.
bed-parallel pss development was enhanced not only by Upon transecting the entire most competent bed, due
discrete amounts of clayish matrix (2-4% of total rock to bed tilting as a consequence of the aforementioned slip
volume), but also by the degree of sorting and sphericity accumulation in the adjacent sheared pss, the splay joints
of the carbonate grains. COLLINS et alii (1992), on the also begin to shear forming splay pss at their contrac-
other hand, though that the clay distribution (particle tional quadrants where they tip out in the less competent
rimming versus pore filling) was more important than the beds (fig. 14b, step 3). Formation of splay pss due to
overall amount of clay for fracture development within shearing of a parental fracture has been already reported
the Austin Chalk Formation of central Texas. for carbonate rocks by RISPOLI (1981), GRAHAM et alii
(2003), AGOSTA & AYDIN (2006), AGOSTA (2008), and
CONCEPTUAL MODEL OF NORMAL FAULT GROWTH ACROSS
ANTONELLINI et alii (2008). An apparent conjugate set of
A CARBONATE MULTI-LAYER
pss forms, more or less, contemporaneously with this
stage of deformation along the pre-existing zones of
Taking into account the aforementioned discussion of weakness defined by the flaser fabric. Although the exam-
the data, a conceptual model of normal fault growth ple in fig. 14b shows the primary fault as a Set 1 pressure
across a carbonate multi-layer is proposed. Starting from solution seam, an analogous process can also be inferred
a background deformation made up of the flaser fabric for the Set 2 pss. In this case, the Set 2 acts as the pri-
due to burial diagenesis (fig. 14a, steps 1 and 2), it is mary fault segments while the Set 1 pss as the apparent
noted that variations in grain type, such as coarse skeletal conjugate set.
grains, can define positions of the ellipsoids (WANLESS, In the case of a restraining jog (fig. 14c, steps 1 and 2),
1979). It appears that this particular fabric determined no mechanism is readily available for aiding normal
orientations particularly prone to failure along the flanks faults to propagate across the most competent beds.
of the limestone ellipses (fig. 14a, step 3). In the proposed Nevertheless, joints may propagate across these beds in
conceptual model, both Set 1 and Set 2 pss form weak an analogous manner as in the releasing jog configura-
zones with an apparent conjugate geometry. tion. Upon transecting the most competent beds, due to
The most competent beds, dolostones (Upper Glen bed tilting as discussed above, one of the fault segments
Rose Formation) and packestones (Buda Formation), are may instigate a releasing jog forming an additional pss in
characterized by a small amount of pss and lead to fault the adjacent most competent beds (fig. 14c, step 3).
formation through other mechanisms. Being primarily
composed of crystalline dolomite and a very small
amount of clay, and therefore less susceptible to pressure CONCLUSIONS
solution and other types of ductile behaviour (DONATH,
1970), in the proposed model these most competent beds We documented the inner architecture of small-offset
likely fail through opening-mode fracturing producing normal faults that crosscut the Upper Glen Rose and Buda
joints. As previously noted, all the study normal faults Formations of central Texas, USA. These normal faults are
that transect the most competent beds appear to do so by characterized by scalloped vertical traces, and developed
incorporating high-angle joints within releasing jogs. Dif- within a multi-layer comprised of more competent (dolo-
ferently, despite the restraining jogs exhibit splay joints stones and packestones) and less competent (wackestones)
associated with individual normal fault segments, they do carbonate beds. The nature, attitude, geometry, distribu-
not develop into fracture clusters and thus larger offset tion, and relative timing of formation of the several struc-
normal faults. As discussed below, the mechanical con- tural elements present within the fault zones has been
trasts of the lithologies does not facilitate the fault devel- investigated by integrating field structural analysis, optical
opment by mean of restraining jog configurations. microscopy observations and SEM investigations.
Evolutionary fault scenarios offer some explanations The results are consistent with presence of two main
for the development of normal faults across the carbonate structural elements, joints/veins and pressure solution
multi-layer and their resulting jog geometries. Two end- seams. Opening-mode fractures, joints/veins, localized
member evolutionary scenarios are proposed. The pss mainly within the grain-supported or crystalline beds,
in the wackestone beds intercept the most competent which formed the most competent units of the carbonate
ones forming either releasing jogs (fig. 14b, step 1) or multi-layer. Some of the original joints were later filled
restraining jogs (fig. 14c, step 1). Each scenario involves with a calcite cement precipitated by mineralizing fluids.
the sequential formation and shearing of splay fractures Pressure solution, differently, focussed within the matrix-
(figs. 14b and c, steps 2 and 3) resulting in the jog confi- supported carbonate beds forming two different sets of
guration observed in the field (figs. 14b and c, step 4). pressure solution seams, which formed an apparent con-
In the case of an initial releasing jog (fig. 14b, step 1), jugate geometry. Based upon the relative timing of forma-
the sheared pss form splay joints when intercepting the tion of the aforementioned structural elements, assessed
most competent beds (fig. 14b, step 2). This process, by means of detailed mapping of key outcrops, a concep-
which invokes a parent sheared fracture-splay relation- tual model of normal fault growth across a carbonate
ship widely documented for normal faults in layered car- multi-layer was then proposed.
438 F. AGOSTA ET ALII

Fig. 14 - Conceptual model of fault growth across a carbonate multi-layer: (a) Map-based conceptualization of the formation of Set 1 and Set 2
pressure solution seams in the Upper Glen Rose; (a1) Formation of initial, primarily bed parallel pressure solution seams with locally sinu-
soidal geometry in a wackestone bed; (a2) Continued pressure solution compartmentalizes the wackestone bed into limestone ellipsoids and
pressure solution seams, forming flaser fabric. This arrangement of the pressure solution seams creates weak zones prone to slip; (a3) Set 1
and Set 2 sheared pressure solution seams localize along the weak zones. A photograph of the traced flaser fabric along with the two shear
zones is shown for comparison; (b, c) Modalities of normal fault growth in correspondence of a releasing and restraining jog geometry,
respectively. See text for details.
THE ROLE OF MECHANICAL STRATIGRAPHY 439

Based upon the jog geometries formed by the pre- AGOSTA F., RUANO P., RUSTICHELLI A., TONDI E., GALINDO-ZALDIVAR J.
existing pressure solution seams, the proposed model & SANZ DE GALDEANO C. (2012) - Inner structure and deforma-
tion mechanisms of normal faults in conglomerates and car-
considered two end-members evolutionary scenarios. bonate grainstones (Granada Basin, Betic Cordillera, Spain):
Each scenario involved the sequential shearing and for- inferences on fault permeability. Journal of Structural Geology,
mation of splay fractures. The pressure solution seams, 45, 4-20.
mainly present within the wackestone beds, intercepted ANTONELLINI M., TONDI E., AGOSTA F., AYDIN A. & CELLO G. (2008) -
the most competent ones forming either releasing or Failure modes in basin carbonates and their impact on fault
restraining jogs. In the first case, the sheared pressure development, Majella Mountain, central Italy. Marine and Petro-
solution seams formed splay joints when intercepting the leum Geology, 25, 1074-1096.
most competent beds. The latter structural elements, AYDIN A., ANTONELLINI M., TONDI E. & AGOSTA F. (2010) - Deforma-
tion along the leading edge of the Majella thrust sheet in central
could eventually dilate and propagate sub-perpendicular Italy. Journal of Structural Geology, 32, 1291-1304.
to bedding while slip accumulated along the sheared BARKER D.S. (1987) - Tertiary alkaline magmatism in Trans-Pecos
parental structure. Upon transecting the entire most com- Texas. Geological Society of America Special Publication, 30,
petent bed, due to bed tilting, the splay joints could also 415-431.
undergo to shear forming splay pressure solution seams BARKER R.A., BUSH P.W. & BAKER T.E. JR. (1994) - Geologic history
at their contractional quadrants when tipping out in the and hydrogeologic setting of the Edwards-Trinity aquifer system,
less competent beds. In the case of restraining jogs, no West-Central Texas. U.S. Geological Survey Water Resources
Investigation Report 94-4039, 50 pp.
mechanism was available for aiding normal faults to
propagate across the most competent beds. BATHURST R.G. (1987) - Diagenetically enhanced bedding in argilla-
ceous platform limestones: stratified cementation and selective
We annotate that a good understanding of the geome- compaction. Sedimentology, 34, 749-778.
try, dimension, aperture distribution and overall architec- BEHRENS E.W. (1965) - Environment reconstruction for a part of the
ture of normal faults that crosscut a carbonate multi- Glen Rose Limestone, central Texas. Sedimentology, 4, 65-111.
layer could aid in the exploration as well as the BIENIAWKSI Z.T. (1984) - Rock Mechanics Design in Mining and Tun-
management of subsurface geofluids. In fact, all geolo- neling. Boston, A.A. Balkema, 212 pp.
gists dealing with oil, gas and water resources would ben- BRAND J.P. & DEFORD R.K. (1958) - Comanchean stratigraphy of
efit of the knowledge derived from detailed studies of the Kent quadrangle, Trans-Pecos Texas. American Association of
processes associated to fault growth. In particular, for a Petroleum Geologists Bulletin, 42, 371-387.
carbonate multi-layer characterized by a pronounced BUXTON T.M. & SIBLEY D.F. (1981) - Pressure solution features in a
lithological variation, the knowledge gained from this shallow buried limestone. Journal of Sedimentary Petrology, 51,
19-26.
work may help to better assess the discrete volumes in
CHANCHANI J., BERG R.R. & CHUNG-I.D.F. (1996) - Pressure solution
which opening-mode deformation may have clustered and microfracturing in primary oil migration, Upper Cretaceous
during the fault growth. In fact, a good prediction of the Austin Chalk, Texas Gulf Coast. Gulf Coast Association of Geo-
subsurface distribution of releasing jogs would be benefi- logical Societies Transactions, 46, 71-78.
cial in order to target spots characterized by high values CHILDS C., NICOL A., WALSH J.J. & WATTERSON J. (1996) - Growth
of splay joint frequency and, hence, connected fracture of vertically segmented normal faults. Journal of Structural Geo-
porosity. logy, 18, 1389-1397
COLLINS E., LAUBACH S. & HOVORKA S. (1992) - Fracture systems in
the Austin Chalk, north-central Texas. In: Schmoker J., Coalson E.
ACKNOWLEDGMENTS & Brown C. (eds.), Geological studies relevant to horizontal
drilling: examples from western North America. Denver, Colo-
This work was initially supported by a grant provided by
rado, Rocky Mountain Association of Geologists, 129-142 pp.
Anadarko Petroleum Corporation and the Stanford Rock Project.
We thank Andrea Brogi and an anonymous referees for the careful COLLINS E.W. (2000) - Geologic map of the New Braunfels, Texas,
revision of the text, David Ferrill and Alan Morris for introducing us 30×60 minute quadrangle: Geologic framework of an urban-
to the Buda Limestone, and Jim Grant for the help provided during growth corridor along the Edwards aquifer, south-central Texas.
the first phase of the project. Bob Jones provided tremendously The University of Texas at Austin, Bureau of Economic Geology,
helpful instruction in using scanning electron microprobe. Bill Miscellaneous Map 39.
Ward graciously discussed some carbonate nomenclature and fossil COLE J. & PANTEA M. (2004) - Three-dimensional structural modeling
identification. of the displacement of the Edwards aquifer within the Balcones
Fault Zone in Northern Bexar County, Texas. Gulf Coast Associa-
tion of Geological Societies Transactions, 54, 121-131.
REFERENCES CONDON S.M. & DYMAN T.S. (2006) - 2003 geologic assessment of
undiscovered conventional oil and gas resources in the Upper
AGOSTA A. & AYDIN A. (2006) - Architecture and deformation mecha- Cretaceous Navarro and Taylor Groups, Western Gulf Province,
nism of a basin-bounding normal fault in Mesozoic platform Texas. U.S. Geological Survey Digital Data Series DDS-69-H,
carbonates, central Italy. Journal of Structural Geology, 28, 2, 42.
1445-1467. CORBETT K., FRIEDMAN M. & SPANG J. (1987) - Fracture development
AGOSTA A. (2008) - Fluid flow properties of large, basin-bounding nor- and mechanical stratigraphy of Austin Chalk, Texas. American
mal fault zones in massive carbonates. In: Wibberly C., Kurz W., Association of Petroleum Geologists Bulletin, 71, 17-28.
Imber J., Holdsworth R. & Collettini C., “The Internal Structure CURRIE J.B., PATNODE H.W. & TRUMP R.P. (1962) - Development of
of Fault Zones: Implications for Mechanical and Fluid-Flow folds in sedimentary strata. Geological Society of America Bul-
Properties”. Geological Society of London Special Publication, letin, 73, 655-673.
299, 277-291. DAVATZES N.C. & AYDIN A. (2003) - The formation of conjugate nor-
AGOSTA F., ALESSANDRONI M., TONDI E. & AYDIN A. (2009) - Oblique mal fault systems in folded sandstone by sequential jointing and
normal faulting along the northern edge of the Majella Anticline, shearing, Waterpocket Monocline, Utah. Journal of Geophysical
central Italy. Inferences on hydrocarbon migration and accumu- Research, 108, 2478. doi: 10.1029/2002JB002289.
lation, 31, 674-690. DHOLAKIA S.K., AYDIN A., POLLARD D.D. & ZOBACK M.D. (1998) -
AGOSTA F., ALESSANDRONI M., ANTONELLINI M. & TONDI E. (2010) - Fault-controlled hydrocarbon pathways in the Monterey Forma-
From fractures to flow: a field-based analysis of an outcropping tion, California. American Association of Petroleum Geologists
carbonate reservoir. Tectonophysics, 490, 197-213. Bulletin, 82, 1551-1574.
440 F. AGOSTA ET ALII

DONATH F.A. (1970) - Some information squeezed out of rock. Ameri- JAEGER J.C., COOK N.G.W. & ZIMMERMAN R.W. (2007) - Funda-
can Scientist, 58, 54-72. mentals of rock mechanics, 4th edition. Malden, Blackwell Pub-
DRUGUET E., ALSOP G.I. & CARRERAS J. (2009) - Coeval brittle and lishing, 475 pp.
ductile structures associated with extreme deformation partition- KATSMAN R., AHARONOV E. & SCHER H. (2006) - A numerical study
ing in a multilayer sequence. Journal of Structural Geology, 31, on localized volume reduction in elastic media: Some insights on
498-511. the mechanics of anticracks. Journal of Geophysical Research,
DUNHAM R.J. (1962) - Classification of carbonate rocks according to 111, B03204. doi: 10.1029/2004JB003607.
depositional texture. American Association of Petroleum Geolo- KING P.B. & FLAWN P.T. (1953) - Geology and mineral deposits of
gists Memoirs, 1, 108-121. Precambrian rocks of the Van Horn area, Texas. The University of
EIFLER G.K. (1951) - Geology of the Barrilla Mountains, Texas. Geo- Texas Publication, 530, 281 pp.
logical Society of America Bulletin, 62, 339-353. KERANS C. (2005) - Field Trip Stop 4: Lower Glen Rose rudist biohemrs,
EVANS M.A., LEWCHUK M.T. & ELMORE R.D. (2003) - Strain parti- Pipe Creek, Bandera County. In: Filkhorn H.F., Johnson C.C.,
tioning of deformation mechanisms in limestones: examining the Molineux A. & Scott R.W. (eds.), Abstracts and Post-Congress
relationship of strain and anisotropy of magnetic susceptibility Field Guide. Seventh International Congress on Rudists, Austin,
(AMS). Journal of Structural Geology, 25, 1525-1549. Texas, 156-166.
ESCHBAC O.W. (1961) - Handbook of Engineering Fundamentals. LAUBACH S.E. & JACKSON M.L.W. (1990) - Origin of arches in the
Handbook Series, New York. John Wiley and Sons, 319 pp. northwestern Gulf of Mexico basin. Geology, 7, 595-598.
EWING T.E. (1987) - The Frio River Line in south Texas-Transition LAUBACH S.E., MACE R.E. & NANCE H.S. (1995) - Fault and joint
from Cordilleran to Northern Gulf tectonic regimes. Gulf Coast swarms in a normal fault zone. Proceedings of the second inter-
Association of Geological Societies Transactions, 37, 87-94. nal conference on the mechanics of jointed and faulted rock, 2,
FERRILL D.A. & MORRIS A.P. (2003) - Dilational Normal Faults. Jour- 305-309.
nal of Structural Geology, 25, 183-196. LAUBACH S.E., OLSON J.E. & GROSS M.R. (2009) - Mechanical and
FERRILL D.A., SIMS D.W., WAITING D.J., MORRIS A.P., FRANKLIN N.M. fracture stratigraphy. American Association of Petroleum Geolo-
& SCHULTZ A.L. (2004) - Structural framework of the Edwards gists Bulletin, 93, 1413-1426.
Aquifer recharge zone in south-central Texas. Geological Society LEHMAN T.M. (1986) - Late Cretaceous sedimentation in Trans-Pecos
of America Bulletin, 116, 407-418. Texas. In: Pause P.H. & Spears R.G. (eds.), Geology of the Big
FERRILL D.A. & MORRIS A.P. (2008) - Fault zone deformation Bend area and Solitario dome, Texas. West Texas Geological
controlled by carbonate mechanical stratigraphy, Balcones fault Society 1986, Field Trip Guidebook, 105-110.
system, Texas. American Association of Petroleum Geologists MARSHAK S. & ENGELDER T. (1985) - Development of cleavage in
Bulletin, 92, 359-380. limestone of a fold-thrust belt in eastern New York. Journal of
FISCHER M.P. & JACKSON P.B. (1999) - Stratigraphic controls on Structural Geology, 7, 345-359.
deformation patterns in fault-related folds: A detachment fold MCCROSSAN R.G. (1958) - Sedimentary “boudinage” structures in the
example from the Sierra Madre Oriental, northeast Mexico. Jour- Upper Devonian Ireton Formation of Alberta. Journal of Sedi-
nal of Structural Geology, 21, 613-633. mentary Petrology, 28, 316-320.
FLETCHER R.C. & POLLARD D.D. (1981) - Anticrack model for pres- MCKALIPS D. & BARNES V.E. (1981) - Sonora Sheet: UTBEG Geologic
sure solution surfaces. Geology, 9, 419-424. Atlas of Texas. Roy Thorpe Hazzard Memorial Edition, scale
FLÜGEL E. (2004) - Microfacies of Carbonate Rocks. Berlin, Springer- 1:250,000.
Verlag, 976 pp. MORRIS A.P., FERRILL D.A. & MCGINNIS R.N. (2009) - Mechanical
FOSSEN H. (2010) - Structural Geology. Cambridge University Press, Stratigraphy and Faulting in Cretaceous Carbonates. American
463 pp. Association of Petroleum Geologists Bulletin, 93, 1459-1470.
FULLMER S. & LUCIA F.J. (2006) - Burial history of Central Texas MUEHLBERGER W.R., BELCHER R.C. & GOETZ L.K. (1978) - Quater-
Cretaceous carbonates. Gulf Coast Association of Geological nary faulting in Trans-Pecos Texas. Geology, 6, 337-340.
Societies Transactions, 55, 225-232. NELSON R.A. (1981) - Significance of fracture sets associated with
GALLOWAY W.E., GANEY-CURRY P.E., LI X. & BUFFLER R.T. (2000) - stylolite zones. AAPG Bulletin, 65, 2417-2425.
Cenozoic depositional history of the Gulf of Mexico basin. NENNA F.A. & AYDIN A. (2011) - The formation and growth of pres-
American Association of Petroleum Geologists Bulletin, 84, sure solution seams in clastic rocks: A field and analytical study.
1743-1774. Journal of Structural Geology, 33, 1595-1610.
GARRISON R.E. & KENNEDY W.J. (1977) - Origin of solution seams PITTMAN J.G. (1989) - Stratigraphy of the Glen Rose Formation, West-
and flaser fabric in Upper Cretaceous Chalks of Southern ern Gulf Coastal plain. Gulf Coast Association of Geological
England. Sedimentary Geology, 19, 107-137. Societies Transactions, 39, 247-264.
GRAHAM B., ANTONELLINI M. & AYDIN A. (2003) - Formation and PARKER D.F. & MCDOWELL F.W. (1979) - K-Ar geochronology of
growth of normal faults in carbonates within a compressive envi- Oligocene volcanic rocks, Davis and Barrilla Mountains, Texas.
ronment. Geology, 31, 11-14. Geological Society of America Bulletin, 90, 1100-1110.
GRAHAM-WALL B.R., GIRBACES R., MESONJESI A. & AYDIN A. (2006) - PEACOCK D.C.P. & SANDERSON D.J. (1991) - Displacements, segment
Evolution of fracture and fault-controlled fluid pathways in linkage and relay ramps in normal fault zones. Journal of Struc-
carbonates of the Albanides fold-thrust belt. AAPG Bulletin, 90, tural Geology, 13, 721-733.
1227-1249. PEACOCK D.C.P. & SANDERSON D.J. (1995) - Pull-aparts, shear frac-
GROSHONG R.H. (1988) - Low-temperature deformation mechanisms tures and pressure solution. Tectonophysics, 241, 1-13.
and their interpretation. GSA Bulletin, 100, 1329-1360. PETIT J.P. & MATTAUER M. (1995) - Paleostress superimposition
GROSS M.R., FISCHER M.P., ENGELDER T. & GREENFIELD R.J. deduced from mesoscale structures in limestone: the Matelles
(1995) - Factors controlling joint spacing in interbedded sedimen- exposures. Languedoc, France. Journal of Structural Geology,
tary rocks: interpreting numerical models with field observations 17, 245-256.
from the Monterey Formation, USA. In: Ameen M.S. (ed.), Frac- PROCTOR C.V., BROWN T.E., MCGOWEN J.H., WAECHTER N.B. &
tography: Fracture Topography as a Tool in Fracture Mechanics BARNES V.E. (1981) - Austin Sheet: UTBEG Geologic Atlas
and Stress Analysis. Geological Society of America Special Pub- of Texas. Francis Luther Whitney Memorial Edition, scale
lication, 92, 215-233. 1:250.000.
HEALD M.T. (1956) - Cementation of Simpson and St. Peter sand- RAMSAY J.G. & LISLE R.J. (2000) - Techniques of Modern Structural
stones in parts of Oklahoma, Arkansas, and Missouri. Journal of Geology. Volume III. Applications of Continuum Mechanics in
Geology, 64,16-30. Structural Geology. Academic Press, 1061 pp.
HENRY C.D. & PRICE J.G. (1986) - Early Basin and Range develop- REASER D.F & DAWSON W.C. (1995) - Geologic Study of Upper Creta-
ment in Trans-Pecos Texas and adjacent Chihuahua: Magmatism ceous (Cenomanian) Buda Limestone in Northeast Texas with
and orientation, timing, and style of extension. Journal of Geo- analysis of some Regional Implications. Gulf Coast Association
physical Research, 91, 6213-6224. of Geological Societies Transactions, 45, 495-502.
THE ROLE OF MECHANICAL STRATIGRAPHY 441

RISPOLI R. (1981) - Stress fields about strike-slip faults inferred from WALSH J.J., WATTERSON J., HEATH A.E. & CHILDS C. (1998) - Repre-
stylolites and tension gashes. Tectonophysics, 75, T29-T36. sentation and scaling of faults in fluid flow models. Petroleum
ROCHE V., HOMBERG C. & ROCHER M. (2012) - Architecture and Geoscience, 4, 241-251.
growth of normal fault zones in multilayer systems: a 3D field WALSH J.J., BAILEY W.R., CHILDS C., NICOL A. & BONSON C.G.
analysis in the South-Eastern Basin, France. Journal of Structural (2003) - Formation of segmented normal faults: a 3D perspective.
Geology, 37, 19-35. Journal of Structural Geology, 25, 1251-1262.
RUSTICHELLI A., TONDI E., AGOSTA F., CILONA A. & GIORGIONI M. WANLESS (1979) - Limestone response to stress: pressure solution
(2012) - Development and distribution of bed-parallel compaction and dolomitization. Journal of Sedimentary Petrology, 49,
bands and pressure solution seams in carbonates (Bolognano For- 437-462.
mation, Majella Mountain, Italy). Journal of Structural Geology, WARD W.C. & WARD W.B. (2007) - Stratigraphy of middle part of Glen
37, 181-199. Rose Formation (Lower Albian), Canyon Lake Gorge, Central
RUSTICHELLI A., AGOSTA F., TONDI E. & SPINA V. (2013a) - Spacing Texas, U.S.A. SEPM Special Publication, 87, 193-210.
and distribution of bed-perpendicular joints throughout layered,
WEEKS A.W. (1945) - Balcones, Luling, and Mexia fault zones in
shallow-marine carbonates (Granada Basin, southern Spain).
Texas. American Association of Petroleum Geologists Bulletin,
Tectonophysics, 582, 188-204.
29, 1733-1737.
RUSTICHELLI A., TONDI E., AGOSTA F., DI CELMA C. & GIORGIONI M.
(2013b) - Sedimentologic and diagenetic controls on pore-network WEYL P.K. (1959) - Pressure solution and porosity. Society of Eco-
characteristics of Oligocene-Miocene ramp carbonates (Majella nomic Paleontologists and Mineralogists Special Publication, 7,
Mountain, central Italy). American Association of Petroleum 92-110.
Geologists Bulletin, 97 (3), 487-524. WILKINS S.J. & GROSS M.R. (2002) - Normal fault growth in layered
RUTTER E.H. (1983) - Pressure solution in nature, theory, and experi- rocks at Split Mountain, Utah: Influence of mechanical stratigra-
ment. Journal of the Geological Society (London), 140, 725-740. phy on dip linkage, fault restriction and fault scaling. Journal of
Structural Geology, 24, 1413-1429.
SALVINI F., BILLI A. & WISE D.U. (1999) - Strike-slip fault-propagation
cleavage in carbonate rocks: the Mattinata fault zone, Southern WILSON C.E., AYDIN A., DURLOFSKY L.J., BOUCHER A. & BROWNLOW D.
Apennines, Italy. Journal of Structural Geology, 21, 1731-1749. (2011a) - Use of outcrop observations, geostatistical analysis, and
flow simulation to investigate structural controls on secondary
SCHOLLE P.A. (1977) - Chalk diagenesis and its relation to petroleum
hydrocarbon migration in the Anacacho Limestone, Uvalde, TX.
exploration: Oil from chalks, a modern miracle? American Asso-
American Association of Petroleum Geologists Bulletin, 95,
ciation of Petroleum Geologists Bulletin, 61, 982-1009.
1181-1206.
SCHÖPFER M.P.J., CHILDS C. & WALSH J.J. (2006) - Localisation
of normal faults in multilayer sequences. Journal of Structural WILSON C.E., AYDIN A., KARIMI-FARD M., DURLOFSKY L.J., SAGY A.,
Geology, 28, 816-833. BRODSKY E.E., KREYLOS O. & KELLOG L.H. (2011b) - From out-
crop to flow simulation: constructing discrete fracture models
SCHÖPFER M.P.J., CHILDS C., WALSH J.J., MANZOCCHI T. & KOYI H.A. from a LiDAR Survey. American Association of Petroleum Geo-
(2007) - Geometrical analysis of the refraction and segmentation logists Bulletin, 95, 1883-1905.
of normal faults in periodically layered sequences. Journal of
Structural Geology, 29, 318-335. WILTSCHKO D., CORBETT K., FRIEDMAN M. & HUNG J. (1991) - Pre-
dicting fracture connectivity, intensity within the Austin Chalk
STAPP W.L. (1977) - The geology of the fractured Austin and Buda for-
from outcrop fracture maps and scanline data. Gulf Coast Asso-
mations in the subsurface of south Texas. Gulf Coast Association
ciation Geological Society Transactions, 41, 702-717.
of Geological Societies Transactions, 27, 209-229.
STOCKDALE P.B. (1922) - Styolites: Their nature and origin. Indiana WILLEMSE E.J.M., PEACOCK D.C.P. & AYDIN A. (1997) - Nucleation
University, Bloomington, Indiana, 97 pp. and growth of strike-slip faults in limestones from Somerset, U.K.
Journal of Structural Geology, 19, 1461-1477.
STRICKLIN F.L. & AMSBURY D.L. (1971) - Stratigraphy of Lower Creta-
ceous Trinity deposits of Central Texas. University of Texas at WOODWARD N.B. & RUTHERFORD R. JR. (1989) - Structural lithic
Austin, Bureau of Economic Geology, Report of Investigations, units in external orogenic zones. Tectonophysics, 158, 247-267.
71, 63 pp. YOUNG K. (1974) - Lower Albian and Aptian (Cretaceous) Ammonites
TAYLOR L. & GRANT J. (2004) - Carbonate deformation: outcrop of Texas. Geosciences and Man, 8, 77-102.
analogs for fractured reservoirs. Guidebook for American Associ- ZHOU X. & AYDIN A. (2010) - Mechanics of pressure solution seam
ation of Petroleum Geologists Annual Convention Field trip, growth and evolution. Journal of Geophysical Research, 115,
Dallas, April 16-18. B12207. doi: 10.1029/2010JB007614.

Manuscript received 13 March 2014; accepted 10 June 2014; published online 11 May 2015; editorial responsability and handling by S. Mazzoli.

You might also like