You are on page 1of 12

Marine and Petroleum Geology 97 (2018) 556–567

Contents lists available at ScienceDirect

Marine and Petroleum Geology


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

Research paper

Quantifying aperture, spacing and fracture intensity in a carbonate reservoir T


analogue: Crato Formation, NE Brazil
T.S. Mirandaa,∗, R.F. Santosb, J.A. Barbosaa, I.F. Gomesb, M.L. Alencarc, O.J. Correiac,
T.C. Falcãod, J.F.W. Galee, V.H. Neumanna
a
Department of Geology, Federal University of Pernambuco, Recife, PE, 50740-530, Brazil
b
Civil Engineering Graduate Program, Federal University of Pernambuco, Recife, PE, 50740-530, Brazil
c
Geosciences Graduate Program, Federal University of Pernambuco, Recife, PE, 50740-530, Brazil
d
Petrobras Research Center (CENPES), Rio de Janeiro, 21941-915, Brazil
e
Bureau of Economic Geology, The University of Texas at Austin, Austin, TX, 78712-1101, USA

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

Keywords: The characterization of fracture networks in carbonate reservoirs represents a primary approach to understand
Veins the processes of fluid flow. The study of outcrop analogues has been used to enhance the understanding of the
Cluster relative contribution of fracture networks to reservoir permeability. The outcrop data can overcome scale lim-
Power-law itations and/or lack of consistency of subsurface datasets (e.g. seismic, well logs, and cores). This paper presents
Laminites
the fracture characterization of the Aptian-Albian lacustrine laminites of Crato Formation (Araripe Basin, NE
Scanline
Brazil), which have been investigated as an analogue to some of the observed carbonate facies present in the pre-
Araripe basin
salt reservoir sequence of the Brazilian marginal basins. In order to provide fracture attributes for Discrete
Fracture Network (DFN) modelling we investigate spacing and aperture distribution, length and intensity of
fractures based on the use of traditional scanline (macro and microscale) method. The main fractures identified
in the Crato Formation are shear and extensional fractures and with stylolites also present. In this study, we
focused on vertical veins, which strike in two principal directions, NNW-SSE (Set 1), NE-SW (Set 2), and are
filled or partially filled by calcite or gypsum. The relationship between length and aperture of vertical veins is
linear. Apertures and spacings of fractures in both sets 1 and 2 follow power law and lognormal cumulative
frequency distribution functions, respectively. Fractures of Set 2 are more likely to be clustered than are fractures
of Set 1, which have a wider kinematic aperture. This structural database can be used to populate computational
models that consider the widespread fracture system in the fluid flow simulation of carbonate reservoirs.

1. Introduction lack of information regarding sub seismic fractures is the study of


outcrop reservoir analogues (Bisdom et al., 2017). Outcrop studies
Fracture characterization has motivated an increasing number of make it possible to construct realistic 3D models (geological and nu-
investigations regarding reservoir numerical modelling due to its im- merical) using data accessible at the Earth's surface, which cannot
pact on fluid flow prediction of ground water and hydrocarbon re- otherwise be observed in cores or seismic surveys (Hardebol and
servoirs (Guerriero et al., 2013; Massaro et al., 2018; Laubach et al., Bertotti, 2013; Watkins et al., 2015; Healy et al., 2017).
2018a). Fracture network permeability is governed by the occurrence of Analogue reservoir modelling using outcrop datasets can be applied
fractures, their aperture, spacing, length, topology, and whether or not to numerical simulations using equivalent fine scale and upscaling to
they are filled with cement (National Research Council, 1996). How- relevant properties (e.g. porosity and permeability) taking into account
ever, a large portion of fractures are below the limit of seismic re- the resultant structural geometries from DFN models (Falcão et al.,
solution and these fractures are under sampled due to the limited area 2018).
investigated by subsurface probes and the consequent small size of This study focuses on fracture characterization of a carbonate re-
subsurface datasets (Bonnet et al., 2001). One way to correct for this servoir analogue from the laminites of the Crato Formation, Lower


Corresponding author.
E-mail addresses: tiago.smiranda@ufpe.br (T.S. Miranda), rafaelfvcs@hotmail.com (R.F. Santos), jose.antoniob@ufpe.br (J.A. Barbosa),
gomes@ufpe.br (I.F. Gomes), marciolimaalencar@gmail.com (M.L. Alencar), osv.correia@gmail.com (O.J. Correia), tfalcao@petrobras.com.br (T.C. Falcão),
julia.gale@beg.utexas.edu (J.F.W. Gale), neumann@ufpe.br (V.H. Neumann).

https://doi.org/10.1016/j.marpetgeo.2018.07.019
Received 16 May 2018; Received in revised form 10 July 2018; Accepted 18 July 2018
Available online 01 August 2018
0264-8172/ © 2018 Elsevier Ltd. All rights reserved.
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Fig. 1. A) Location of the Araripe Basin in NE Brazil


(yellow star) and the geological map of the study
area, which is located in the 3 Irmãos quarry (black
star), in the region of Nova Olinda, Ceará State; b)
Chronostratigraphic section of the Post-rift sequence
of Araripe Basin. Gp = Gypsum, Ipubi Formation.
(For interpretation of the references to colour in this
figure legend, the reader is referred to the Web ver-
sion of this article.)

Cretaceous lacustrine section of the Araripe Basin (AB), NE Brazil analogue. This data will allow us to better parametrize future numerical
(Fig. 1). The laminites of the Crato Formation are analogous (strati- simulations of carbonate reservoir analogues.
graphic equivalent) to one of the carbonate facies encountered in the
pre-salt reservoir of the Brazilian marginal basins (Catto et al., 2016;
2. Geological settings
Terra et al., 2010).
We collected data that provide parameters for the multi-scale in-
The study area is located in the AB, the largest interior basin of
tegration (thin sections, outcrops) of reservoir models, supplying
Northeast Brazil. The Crato Formation belongs to the post-rift tectono-
structural data for improved representation of geological hetero-
stratigraphic sequence and represents the main lacustrine phase of the
geneities, especially extensional fractures. To develop a fracture system
AB (Assine, 2007; Neumann, 1999) (Fig. 1b). The AB North Rim is
database that can be used in subsequent reservoir analogue modelling,
mostly well known for its pterosaur bone beds and the Crato Formation
we applied the scanline technique to the laminites, focusing on the
represents a famous lagerstatte deposit that preserves pristine soft body
following fracture attributes: orientation, crosscutting relationships,
fossils (Barling et al., 2015; Field and Martill, 2017). The origin of the
composition, texture of fracture fill, fracture aperture-size and spacing
laminites is still a matter of debate, with some arguing for an inorganic
(Gomez and Laubach, 2006; Gale et al., 2007; Guerriero et al., 2010;
authigenic origin and others that point to a for microbial influence as
Hooker et al., 2009; Marrett, 1996; Ortega et al., 2006; Priest, 1993;
the main source (Heimhofer et al., 2010 and Catto et al., 2016, re-
Sanderson and Nixon, 2015). We also quantified fracture intensity with
spectively). Our study was carried out at the best-expressed level, which
the Normalized Correlation Count (NCC) computer program (Corr-
is exposed in 3 Irmãos complex quarry, northern region of the basin,
Count), which provides a spatial arrangement by normalizing to ex-
Nova Olinda, Ceará state (Fig. 1). The 3 Irmãos quarry contains tridi-
pected values for randomly arranged fractures (Marrett et al., 2018).
mensional exposures providing a valuable opportunity to collect in-
The NCC describes fracture positions in space, indicating the occurrence
formation on the contribution of fracture networks.
of fracture clusters, which is a crucial element needed to support DFN
This study takes the laminites of the Crato Formation as a potential
models and numerical simulations.
analogue of low permeability carbonate facies of pre-salt reservoirs of
Our study develops a conceptual structural evolution model of
Barra Velha Formation, Santos Basin (Fig. 2). This is mainly due to their
fractures in the Crato Formation's laminites. Furthermore, this research
correspondence in depositional age, and similarities in lithology, pet-
is the first to undertake quantification of the aperture, spacing and
rophysical and geomechanical properties (Terra et al., 2010; Miranda
fracture intensity of vertical veins from a Brazilian pre-salt reservoir
et al., 2016; Catto et al., 2016; Menezes et al., 2016; Zihms et al., 2017).

Fig. 2. Samples of planar laminites from (A) Barra Velha Formation, pre-salt reservoir of Santos Basin (SE Brazil) (Terra et al., 2010) and (B) Crato Formation,
outcrop analogue of Araripe Basin (NE Brazil).

557
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

data has great value, 1D scanline data is most readily compared to


subsurface observations. Although our quantitative data is 1D, we also
describe the 2D topology of the fracture patterns.
The following features were recorded through the linear scanline (1-
D) readings: (1) the spacing (perpendicular distance) between fractures
in the same set, (2) the kinematic aperture, (3) the fracture orientation,
(4) the morphology, (5) the crosscutting relationships between frac-
tures and (6) the composition of the fracture fill (Fig. 3a). To measure
the fracture apertures in the macroscanline, we used a fracture-width
comparator with a collection of apertures measuring 0.05–5 mm
(Fig. 3b) (Ortega et al., 2006). The comparator shows logarithmically
graduated lines such that differ by approximately constant increments
when plotted on a logarithmic axis. In order to have greater scale range
in surveys the total lengths of the scanlines made along pavements and
thin section were normalized and compared for each set of extension
fractures. In this paper, the microscanlines were performed according to
the methods in Gomez and Laubach (2006). We selected vertical veins
that show no field evidence of shear displacement, subsequent de-
formation, any of which might indicate alteration of the original
aperture (Vermilye and Scholz, 1995).
We also investigated the distribution of aperture and spacing of
opening-mode fractures by comparing goodness-of-fit for power law,
exponential or lognormal distributions using the log-log plots of aper-
ture (b) and fracture spacing (s) versus cumulative frequency (F)
(Hooker et al., 2014). Following Santos et al. (2015) we established
95% confidence intervals for each of the best fit distribution to elim-
inate sampling biases in fracture data acquisition (scanline). These
biases are recognized as censoring and truncation artifacts, caused by
geometrical restrictions and the sampling dimensions (Pickering et al.,
1995; Ortega et al., 2006). Censoring artifacts are common because
large structures may be undersampled due to the limited length of
scanlines across the outcrop. Since the outcrop surveying procedures
include naked eye and hand lens observations, truncation normally
occurs owing to the difficulty of measurement of small aperture frac-
tures (0.05 mm). In this study we performed microscanlines to resolve
Fig. 3. (A) General view of the laminites outcrop located in the Três Irmãos fracture apertures down to the micron scale using high resolution
Quarry showing an example of the linear scanline positioned on the pavement imaging methods (Milliken and Laubach, 2000; Hooker et al., 2014).
(red line) and sites (green polygon) where laminites samples were collected for After eliminating artifacts due to censoring and truncation we selected
microscanlines. (b) Details of aperture and fracture spacing, and the comparator a representative series to have a better distribution fit and with fewer
(a logarithmically graduated ruler for fracture aperture measurement devel- errors and uncertainties.
oped by Ortega et al. (2006) for use in the field. (c) Example of microscanline Using the CorrCount, NCC computer program the spatial arrange-
using petrographic thin section under crossed nicols (5x). (For interpretation of ment of bed-normal fractures was preliminary quantified (Marrett et al.,
the references to colour in this figure legend, the reader is referred to the Web
2018). The NCC method provides a quantitative analysis of the degree
version of this article.)
to which fracture sets are clustered and can distinguish between even
spacing (periodic or anticlustered), clusters arising due to random ar-
Although sharing similar physical parameters, the subsurface and out- rangement, and clustering that is stronger than a random signal
crop datasets may have some differences regarding diagenetic condi- (Marrett et al., 2018; Laubach et al., 2018b; Li et al., 2018). The
tions and burial history. Such differences can affect fracture patterns CorrCount provides analytical and Monte Carlo solutions for rando-
(e.g., Laubach et al., 2009; Li et al., 2018). mized input spacings and allows a 95% confidence interval to be con-
structed for randomized sequences. If fracture intensity falls either
3. Materials and methods above or below the upper or lower confidence limits, the corresponding
fracture spacing is statistically significant (Marrett et al., 2018; Laubach
To understand the main tectonic stages and to distinguish the nat- et al., 2018b). In this work the NCC settings for both fracture sets were:
ural fracture set distributions, we first collected structural data from the a) uncertainty estimates = Monte Carlo; b) linear windowing = 0; and
Crato Formation's laminites. After the definition of the main sets of c) normalized. Furthermore, the NCC quantifies the coefficient of var-
natural fractures, a series of linear surveys were performed on the pa- iation (CV) of the population of inter-fracture spacings, CV = σ/μ,
vements of outcrops (macroscanline) and thin sections (microscanline) where σ is the standard deviation of the population of spacings and μ is
(Fig. 3). The surveys captured the attributes of the fractures along the the arithmetic mean. For randomly arranged fractures CV is near 1;
macro and microscanlines of the same set following the methods de- CV > 1 if fractures are more clustered than random; and CV < 1 for
veloped by over the past 20 years (Priest, 1993; Marrett, 1996; Marrett those fractures having less clustering than random and it indicates anti-
et al., 1999; Gomez and Laubach, 2006; Ortega et al., 2006; Gale et al., clustering (Gillespie et al., 2001; Hooker et al., 2013; Marrett et al.,
2007; Guerriero et al., 2010; Lu et al., 2017). Measurements in linear 2018).
scanlines that were not perpendicular to the fracture sets were cor-
rected using Terzaghi corrections (Terzaghi, 1965). This allowed us to
maximize the amount of information concerning fracture orientation,
aperture, and spacing. Although higher-dimensional fracture pattern

558
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Fig. 4. Main types of structures of the laminites of Crato Formation. Horizontal veins (blue dashed line) filled by calcite (a) and gypsum (b). c) Stereogram showing
characteristic bedding parallel vein orientation. Shear fracture with normal (d) and inverse (e) displacement and (f) stereogram showing characteristic of shear
fracture orientation. Centimetric vertical stylolites (g and h) and (i) rose diagram showing stylolite orientation. (For interpretation of the references to colour in this
figure legend, the reader is referred to the Web version of this article.)

Fig. 5. Outcrop view showing the Crato Formation's karstic features (a). Rose diagram (b) and stereogram (c) showing the orientation and characteristics of the vuggy
fractures.

4. Results principally shear and extensional fractures. We also found stylolites,


horizontal veins and vuggy fractures. The shear fracture planes were
4.1. Structural characterization identified in conjugate pairs with a preferential N5E direction and
moderate dip (between 20 and 75°) towards ESE and WNW (Fig. 4d, e-
Fractures identified in the laminites of Crato Formation were f). These structures usually occur in association with drag folds formed

559
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

in the enclosing rock. These evidences suggest that during the genera- measured in four macroscanlines, totalling 59 m long, striking at N38W.
tion of these fractures, the laminite deposits had both semi-consolidated The scanlines intersect 173 fractures with apertures ranging from 0.05
and unconsolidated parts. This fact explains the variation of fracture to 66 mm. We measured only three microfractures in the microscanlines
angles (20–75°), considering that some fractures were developed when of Fracture Set 2 where apertures range from 0.01 to 0.005.
the rock was consolidated (brittle regime) and other when the rock was Both fracture sets aperture distributions are best described and re-
more plastic (ductile regime). The angular variation between the planes presented by a power-law rather than lognormal or exponential dis-
of the conjugate pairs of shear fractures depends also on the confining tributions (Fig. 8a, d). The observed power law in a fracture distribution
pressure (extent of deformation), the temperature and the pore fluid. log-log graph (sets 1 and 2) shows that the data form a drop-off at the
Thus, the shear fractures of laminites can be interpreted as syn-sedi- large-size limit of the distribution (i.e., the part of the distribution re-
mentary structures generated during the eodiagenetic phase. Further- presenting macrofractures, indicating a deviation from the power-law
more, due to mechanism deformation observed in macroscopic and scaling trend) (Fig. 8b, e). This deviation may be either real or the effect
microscopic samples, such as granular flow and sliding and its trunca- of sampling biases, such as censoring artifacts (Ortega et al., 2006;
tion by younger structures, the observed shear fractures are classified as Hooker et al., 2013). In the part of the distribution representing small
the earliest structures being formed in the laminites. Vertical stylolites fracture apertures (b < 0.05 mm), it is also possible to observe the
were identified normal to the bedding striking preferentially N50-60E effect of truncation artifacts (Ortega et al., 2006). The power laws of
(Fig. 4g, h-i). The horizontal veins are parallel to bedding, which dips sets one and two indicate a 95% confidence interval. Consequently, the
uniformly at 2°–6° to the ESE (Fig. 4a,b-c). fracture aperture values, which fall within this confidence interval,
The vuggy fractures we identified are directly related to the karst represent the most reliable data set (effective series - best fit function).
features observed in the laminites of Crato Formation. These features Otherwise, the outliers should be interpreted as the data related to
principally strike N60E, N90E and N30W, and are associated with sampling biases. Thus, the censoring biases were represented by large
joints, shear fractures and horizontal veins (bedding), possibly through fractures with b > 3 mm (Set 1), b > 60 mm (Set 2) and the trunca-
the epigenetic processes of dissolution. The fact that the dissolution tion biases were represented by values of b < 0.01 mm for both sets
occurred along a disperse group of pre-existing structures explains the (Fig. 8b,e).
aleatory observed in their rose diagram (Fig. 5). The vuggy fractures are The cumulative frequency (effective series) of fracture sets 1 and 2
a result of the intense dissolution process in some outcrops, with the follow a power-law aperture-size having a scaling exponent of about -
potential to reach the scale of small caverns (Fig. 5). 0.44 and - 0.66, respectively (Fig. 8c). After eliminating the artifacts
(censoring and truncation) we identified that there is a consistent im-
4.1.1. Extensional fractures provement in the R2 (coefficient of determination) values from 0.93 to
The extensional fractures were classified as vertical joints and veins. 0.97 in Set 1 and 0.97–0.99 in Set 2.
The joints occur without macroscopically visible fill or are partially The cumulative frequency of fracture spacings (s) of sets 1 and 2
filled with calcite and/or gypsum (Fig. 6). In some outcrops dissolution follow a lognormal distribution better than power law or exponential
features are visible and filling of calcite and or gypsum occur on the fits (Fig. 9a, d). Considering the 95% confidence interval of the log-
walls of the joints, indicating that these processes can be related to the normal regression, it was possible to remove the outliers interpreted as
telodiagenetic processes, by the effect of infiltration of meteoric water artifacts of fracture spacings (Fig. 9b, e). This process allowed us to
due to the exhumation of the deposits and subsequent erosion of the have an effective series with a greater determination coefficient (R2),
strata. The observed veins are totally filled with calcite and fill textures where truncation biases were represented by values of s < 1 mm (Set
can be classified as syntaxial (Fig. 6). Furthermore, these structures 1), s < 2 mm (Set 2) and censoring biases were represented by large
essentially indicate that the calcite fill was precipitated in the subsur- spacing values of s > 7000 mm and s > 2000 m, of sets 1 and 2, re-
face as the fractures evolved (Urai et al., 1991; Ukar and Laubach, spectively (Fig. 9c, f).
2016). Fig. 10 shows the linear adjustments of the fracture sizes 'l (m)' as a
Vertical extensional fractures strike in two main directions: a) Set 1 - function of the apertures 'b (mm)' for sets 1 and 2. In this case, the
NNW-SSE (mean vector 335 Az); b) Set 2 - NE-SW (mean vector 48 Az) general equation of a line can be defined as follows:
(Fig. 6a–b). These fracture sets were defined in the field by their or- (1)
Y= A*X + B
ientation. And based on crosscutting relationship, both Set 1 and Set 2
are contemporaneous in age of deformation. Where Y is the dependent variable and X is the independent variable. 'A′
The vertical vein networks of Crato Formation laminites have three is called the angular coefficient and 'B′ is the linear coefficient. The
types of intersections: 1) I-node; 2) Y-node, which has three branches; angular coefficient refers to the slope of the line adjusted from the least
3) X-node that has four branches (Fig. 7) (Sanderson and Nixon, 2015). squares method. In this work, we have the independent variable aper-
Although all three nodal types were present, the interaction between X ture according to the dependent variable length. After the adjustment
and I-nodes branch structures was the most frequently observed. with the least squares method, we obtain the equations of the lines
shown in Fig. 10. Therefore, if two straight lines have the same angular
4.2. Fracture aperture, spacing and length coefficient, it means that the rate of change, i.e. the first derivative, of
one variable (aperture) relative to another (length) is equal.
In this study, shear-mode fractures, horizontal veins and stylolites Between the two sets of fractures, we noticed that there is practi-
were not included in the scanline surveys due to the difficulty of cally the same angular coefficient (b1 = 0.95 and b2 = 0.91) for both
measuring the exact displacement of the faults, and because the stylo- adjustments. This means that the ratio of fracture size to aperture size
lites were undersampled in many of the scanlines due to their or- has the same rate of change. Accordingly, the variation behavior of the
ientation and due to the low abundance in the study area. Our statistical aperture as a function of length is practically the same. Fractures of Set
study focused only on vertical extensional fractures (veins). 1 have an average of length and aperture values of 1.15 m and 0.5 mm,
We measured 75 kinematic apertures of Set 1 vertical veins along respectively. A mean value of 0.32 m (length) and 0.095 mm (aperture)
four scanlines measured on the outcrop, totaling approximately 90 m was recorded for Set 2.
long with a strike of N65E. The microscanline of Fracture Set 1 inter-
sected 20 microfractures. The macro and micro scanline data processing 4.3. Fracture intensity
allowed us to describe the aperture-size distribution in the log-log plot
of Fig. 8 (cumulative frequency versus kinematic aperture) where Here we present the fracture spatial arrangement based on NCC
apertures range from 0.004 to 30 mm. Set 2 extensional fractures were aimed at assessing whether the vertical vein datasets can be quantified

560
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Fig. 6. Extensional fractures of Crato Formations. (a–b) Rose diagram and stereogram contoured lower hemisphere projections of poles to joints and veins of the
laminites. (c) Joint of Crato Formation completely open. (d) Joint partially filled by gypsum. (e–f) Vertical veins filled by calcite.

and distinguished from random clustering (Marrett et al., 2018). between clusters comparing to fractures of Set 1. These results agree
Fracture Set 1 has a CV value of 1.76, indicating a more clustered with a CV of 1.76 and 2.16, set 1 and 2 (most clustered), respectively
than random spacing. Fracture Set 2 has a CV value of 2.16. Note that (Table 1).
CV values for Set 1 are lower than those from Set 2, which could in-
dicate that Set 2 fractures are more clustered, or irregularly spaced. The 5. Discussion
average fracture intensity (number of fractures across scanline length)
values of sets 1 and 2 are 1.01 and 2.98, respectively (Table 1). 5.1. Chronological relationship between fractures
Fig. 11 shows the normalized fracture intensity plot with the hor-
izontal axis representing the distance in meters along the scanlines The chronologic relationship between the structures of the laminites
performed on the outcrop pavement across fractures sets 1 and 2, re- of Crato Formation, are mainly based on their crosscut relationship and
spectively. The vertical axis shows fracture intensity. The calculated can be summarized, from oldest to youngest by: 1) shear fractures; 2)
intensity values are normalized by expected intensity for randomly vertical stylolites; 3) vertical veins; 4) bed-parallel veins and 5) joints
spaced fractures (Marrett et al., 2018). and vuggy fractures (Fig. 12).
The normalized fracture intensity of Set 1 shows five peaks above It is possible to infer that the first deformational cause that impacted
the upper 95% confidence limit: a) 41–42 m; b) 43–47 m; c) 48–50 m; the laminated carbonates was a horizontal extension in a WNW - ESE
d) 60–66 m and e) 84.5–90 m, with distances between clusters of 1 m, direction. This regime generated the conjugate pairs of shear fractures
1 m, 10 m and 18.5 m, respectively (Fig. 11). This suggests that these (normal faults), typical of an extensional environment. These are
intervals have statistically significant values and deviate from random. characterized by syn-sedimentary deformation based on granular flow
Fractures Set 2 shows six intervals with statistically positive anomalies: deformation mechanism developed in an early and pre-consolidated
a) 10–11 m; b) 13–15 m; c) 44–46 m; d) 46.5–49 m; e) 49.5–50 m and f) phase when sediments porous where full of fluid. Conforming to the
58 m (Fig. 11a). Distances between clusters of fracture Set 2 are 1 m, extensional environment model, within which the Crato Formation is
2 m, 29 m, 0.5 m, 0.5 m and 8 m. Fractures of Set 2 have less distance positioned, represented by the post rift interval of the AB, it would be

561
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Fig. 7. Examples of joints and veins from the Crato


Formation laminites showing different topological
relationships in pavement of outcrops. (a) Y-nodes
created by joints (dashed line) that curves towards an
earlier joint and abuts it at approximately 90° at the
tips of branch. (b) X-nodes created by apparently
mutually crosscutting joints. (c) I-node at the tips of
veins that approach each other and interact. Note
that the vertical linear, fracture-like traces are in fact
marks from the cutting machine.

normal to encounter depositional stylolites in the horizontal position, fractures and mostly release joints planes; vuggy fractures were gen-
concordant with the characteristic of the shear and extension fractures erated. These structures were formed due to the uplift of the Crato
of the formation. Instead, we observe vertical stylolites. One hypothesis Formation and evolved preferentially along preexisting planes of
for the occurrence of these structures with vertical planes, perpendi- weakness (shear, extensional and bedding parallel fractures), gen-
cular to bedding, is the action of a local field of horizontal shortening. erating karstic features in the laminites.
This local horizontal compression could have been generated by dis-
placement of the conjugate shear fracture pairs, described above. The
5.2. Implications for numerical modelling
shear fractures and vertical stylolites are cut by extensional fractures
(joints and veins), indicating that the shear fractures were formed in an
Discrete fracture network models (DFN) are critical to the devel-
earlier deformation event.
opment of effective reservoir models (e.g. Dershowitz and Einstein,
According to the classification of fractures proposed by Nelson
1988; Cacas et al., 1990), especially in frequently difficult to access
(2001), the opening-mode fractures mapped in the laminites can be
ultra-deep petroleum plays. Our data suggest a multiplicity of specific
classified as regional extension fractures. Some aspects of the regional
factors generated the observed fracture distributions in the laminite
geologic context of the AB strongly control the orientation of these
pre-salt analogues from the Crato Formation, NE Brazil. Although dif-
fractures. These extensional fractures in the laminites occur parallel to
ferences between our study area and the corresponding subsurface
the preferential direction of the structural alignments observed on a
analogue due to regional tectonic history and local diagenetic condi-
regional scale in the AB (NW-SE and NE-SW) (Miranda, 2015).
tions are likely to exist, useful up-scalable parameters can nevertheless
Considering that the veins and joints present a difference in the type
be extrapolated from the results acquired surveying the chosen exposed
and nature of filling, it is possible to hypothesize that the veins are older
reservoir analogue.
and occurred under burial conditions (subsurface) when compared with
Instead of trying to investigate 2-D or 3-D fracture patterns, we
joints. The formation of the laminites's vertical veins may be is asso-
carried out a robust 1-D multiscale approach (macro and micro-
ciated with the late stages of the AB's burial. Followed by the uplift
scanlines) to quantify opening-mode fracture attributes more accu-
phase of AB during the Cenozoic (Gurgel et a. 2013), the joints possibly
rately. We found that kinematic aperture data can be fit using re-
were originated from the veins, by erosion of the overlying formations
presentative power-law and log-normal distribution. Likewise, we
due to the process of uplift-related thermoelastic contraction or ex-
documented patterns of normalized fracture intensity and the re-
posure. This process would explain the occurrence of joints with a
lationship between kinematic aperture and fracture length. Our field
significant aperture enlargement due to aerial exposure. Locally, these
data allow us to derive an accurate representation of the power-law
structures also were partially filled with gypsum-rich material, possibly
fracture aperture distribution (fracture sets 1 and 2) over a wide size
derived from the Ipubi Formation. Due to the effect of continuum ex-
range, since data was obtained from two data collection scales (micro
humation process, also pressure release fractures were formed sub-
and macroscale), as shown in Fig. 8. The spacing frequency of fracture
horizontal and parallel to bedding.
sets 1 and 2 indicate that the fracture population fits a log-normal
Finally, due to the dissolution of shear fractures, extensional
distribution (Fig. 9). The log-normality of the distribution implies that

562
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Fig. 8. Log-log plots of cumulative frequency versus fracture aperture showing the trendlines corresponding to the best fit of the scanline data to power law,
lognormal and exponential equations for fracture sets 1 (a) and 2 (d). The scanline correlation with the complete data sets 1 (b) and 2 (e). The effective series (Set 1 - c
and Set 2 - f) after the exclusion of data interpreted as artifacts (truncation and censoring bias). Circles = truncation and censoring artifacts.

the average spacing values is less than the standard deviation, in- The relationship between fracture aperture versus length implies
dicating that the fractures are clustered around low spacing values that all fractures from sets 1 and 2 have the same angular coefficients.
(CV > 1) (Gillespie et al., 2001; Marrett et al., 2018). This is congruent This might suggest the same driving stress at the time of fracture filling
with a CV of 1.76 and 2.16, set 1 and 2, respectively (Table 1). (Olson, 2003). These data are consistent with crosscutting relations

563
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Fig. 9. Log-log plots of cumulative frequency versus fracture spacing showing (a–c) the trendlines corresponding to the best fit of the scanline data to power law,
lognormal, and exponential equations, in which the lognormal provides the best fit for the data; (b–d) the scanline correlation with the complete data set for both
sets, and (c–f) the effective series after the exclusion of data interpreted as artifacts (truncation and censoring bias). Circles = truncation and censoring artifacts.

observed in the field, which suggest that sets 1 and 2 have a con- (Ortega et al., 2003) is controlled by regional weakness of basement
temporaneous deformation timing. rocks that underlying the Crato Formation (Miranda, 2015).
The applied NCC computer program (CorrCount) (Marrett et al., These results can be used to understand the impact of fracture
2018) rigorously quantifies what constitutes a cluster in the normalized aperture and clustering on permeability, and their implications for
fracture intensity. This approach was used along 1D scanlines to com- subsurface reservoirs. Furthermore, fracture clusters have been identi-
pare fracture intensity for different fracture sets. Our results demon- fied as widespread features in reservoir rocks and need to be included in
strate that fracture Set 2 is more clustered than Set 1. Possibly, the reservoir models (Laubach et al., 1995; Questiaux et al., 2010; Belgrano
dominant influence over fracture intensity or spacing, highlighting the et al., 2016; Li et al., 2018).
potential importance of diagenetic and mechanical-property history Massaro et al. (2018) performed DFN modelling based on a

564
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Fig. 10. Linear scaling of aperture versus length data for extensional fractures
(set 1 and 2) of Crato Formation's laminites. Note that both the “b1” and “b2”
angular coefficients are almost equal.

multiscale approach considering mesoscale (outcrop) and large scale


(aerial photos), analyzing a fractured carbonate succession located in
the Sorrento Peninsula, an analogue of buried reservoir units of
southern Italy's major oil fields. In these analyses the authors carried
out 1-D scanlines and observed that the final model provided useful
insights related to fluid flow response in carbonate reservoirs. Blessent
et al. (2009) applied a finite element simulation comparing the DFN
models and equivalent models for a contaminant transport in fractured
media based on field-scale fractured aquitard overlying an aquifer.
The tectonic history behind the fracture distribution and intensity of
the Crato Formation laminites provides a useful basis upon which to
plan and process data obtained from Brazilian Pre-Salt carbonates,
contributing to building a workflow of a robust conceptual DFN mod- Fig. 11. Spatial analysis of Crato Formation vertical veins. Fracture intensity
eling. Beyond the importance of obtaining fractures attributes to help showing statistically significant values - above the upper 95% confidence limit -
parameterize DFN models, our data provide a regional background (red arrows) of (a) fracture Set 1 (NNW-SSE) and (b) Set 2 (NE-SW). (c)
deformation, which in most of the cases is neglected in the models, Schematic diagrams showing fracture spacing heterogeneity based on coeffi-
where users generally associate high-intensity fracture values with cient of variation (CV) - ratio of standard deviation to mean. CV < 1, regularly
large-scale structures. At the stage of this study, NCC results show that spaced fractures; CV = 1, fracture are randomly arranged; CV > 1, fractures
fracture related structure, in Crato Formation, is clustered and related are clustered. White areas = spacing; black line segments = fractures. (For in-
terpretation of the references to colour in this figure legend, the reader is re-
overall to a regional stress field and possibly with the mechanical
ferred to the Web version of this article.)
stratigraphical thickness at the moment of deformation then to the
seismic scale structure. However, this deserves more and more in-
vestigations, with combining of subsurface data such as seismic, important.
gravimetric and magnetometric data. In addition, this could corrobo- In summary, this research shows that their role is evidently critical
rate the indicative of background fractures tend to occur in clusters in to understanding and enhancing the DFN models and fluid flow simu-
the subsurface, with larger spacing and smaller aperture as it gets lations of naturally fractured carbonate reservoirs.
deeper based on geothermal data in Iceland (Gudmundsson, 2011).
Likewise, the clusters spacing measurements can be used to populate 6. Conclusions
the DFN model reducing some uncertainties related to fractures dis-
tribution in the stochastic simulation, where the fractures can be as- The principal structures identified in the laminated carbonates of
sumed randomly located and with homogeneous spatial distribution. Crato Formation are shear and extensional fractures (joints and veins).
Also, it can influence the strategies of well installing, especially in cases Secondary structures include horizontal veins, stylolites and vuggy
in which are applied hydraulic fracturing procedures where the control fractures. Most of the extensional fractures documented in outcrops are
of interaction between hydraulic induced and natural fractures is very classified as vertical veins filled by calcite or gypsum. We describe two

Table 1
Scanline descriptions. b = aperture; CV = coefficient of variation.
Fracture set Dip (mean) Strike Scanline length Number of macro bmin bmax Power law Length vs Aperture Intensity (m-¹) CV spacing
(mean Az) (m) fractures (mm) (mm) (F = a*bˆk)

a k

1 88° 335.6 90 95 0.05 30 0.23 −0.44 b1 = 0.95 l + 0.76 1.05 1.76


2 87° 48.2 59 176 0.05 66 0.36 −0.58 b2 = 0.91 l + 0.34 2.98 2.16

565
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Cacas, M.C., Ledoux, E., de Marsilly, G., Tillie, B., Barbreau, A., Durand, E., 1990.
Modeling fracture flow with a stochastic discrete fracture network: calibration and
validation 1: the flow model. Water Resour. Res. 26 (3), 491–500. https://doi.org/10.
1029/WR026i003p00491.
Catto, B., Jahnert, R.J., Warren, L.V., Varejao, F.G., Assine, M.L., 2016. The microbial
nature of laminated limestones: lessons from the upper aptian, Araripe Basin, Brazil.
Sediment. Geol. 341, 304–315. https://doi.org/10.1016/j.sedgeo.2016.05.007.
Dershowitz, W.S., Einstein, H.H., 1988. Characterizing rock joint geometry with joint
system models. J. Rock Mech. Rock Eng. 21 (1), 21–51. https://doi.org/10.1007/
BF01019674.
Falcão, F., Barroso, J., Murad, M., Pereira, P., Vargas, E., Juvinao, A., Muller, A., Roehl,
D., Quevedo, R., Mejia, C., Guimarães, L., Beserra, L., Alvarez, L., Cleto, P., Manzoli,
O., 2018. Synthetic benchmark for the computation of equivalent properties in
coupled flow and geomechanics conditions for a fractured carbonate rock. In: ARMA
Symposium, DFNE 19-0151_0228_000496.
Field, G.J., Martill, D.M., 2017. Unusual soft tissue preservation in the Early Cretaceous
(Aptian) crocodile cf. Susisuchus from the Crato Formation of north east Brazil.
Cretac. Res. 75, 179–192. https://doi.org/10.1016/j.cretres.2017.04.001.
Gale, J.F.W., Reed, R.M., Holder, J., 2007. Natural fractures in the Barnett Shale and their
importance for hydraulic fracture treatments. Am. Assoc. Petrol. Geol. Bull. 91,
603–622. https://doi.org/10.1306/11010606061.
Gillespie, P.A., Walsh, J.J., Watterson, J., Bonson, C.G., Manzocchi, T., 2001. Scaling
relationships of joint and vein arrays from the Burren, Co. Clare, Ireland. J. Struct.
Geol. 23, 183–201. https://doi.org/10.1016/S0191-8141(00)00090-0.
Gomez, L.A., Laubach, S.E., 2006. Rapid digital quantification of microfracture popula-
tions. J. Struct. Geol. 28, 408–420. https://doi.org/10.1016/j.jsg.2005.12.006.
Gudmundsson, A., 2011. Rock Fractures on Geological Processes. Cambridge University
Fig. 12. Block diagram showing conceptual model of structural features of the
Press, Cambridge 572p.
laminites of Crato Formation. Sh - shear fracture; St - stylolite; Vv - vertical Guerriero, V., Iannace, A., Mazzoli, S., Parente, M., Vitale, S., Giorgioni, M., 2010.
vein; J - joint; Hv - horizontal vein; Vg - vuggy fracture. Quantifying uncertainties in multi-scale studies of fractured reservoir analogues:
implemented statistical analysis of scan line data from carbonate rocks. J. Struct.
Geol. 32, 1271–1278. https://doi.org/10.1016/j.jsg.2009.04.016.
main populations of extensional fractures: a) Set 1, NNW-SSE; Set 2, Guerriero, V., Mazzoli, S., Iannace, A., Vitale, S., Carravetta, A., Strauss, C., 2013. A
NE-SW. The fracture of sets 1 and 2 have a cumulative frequency of permeability model for naturally fractured carbonate reservoirs. Mar. Petrol. Geol.
40, 115–134. https://doi.org/10.1016/j.marpetgeo.2012.11.002.
aperture and spacing follow a power law and log-normal distribution,
Gurgel, S.P.P., Bezerra, F.H.R., Corréa, A.C.B., Marques, F.O., Maia, R.P., 2013. Cenozoic
respectively. Using CorrCount analysis of fracture intensities it was uplift and erosion of structural landforms in NE Brazil. Geomorphology 186, 68–84.
possible to identify that fractures of Set 1 have a higher intensity and Hardebol, N.J., Bertotti, G., 2013. DigiFract: a software and data model implementation
for flexible acquisition and processing of fracture data from outcrops. Comput.
are less likely to be clustered than are fractures of Set 2, which have a
Geosci. 54, 326–336. https://doi.org/10.1016/j.cageo.2012.10.021.
wider kinematic aperture. Healy, D., Rizzo, R.E., Cornwell, D.G., Farrell, N.J.C., Watkins, H., Timms, N.E., Gomez-
Our results show that these methods (macro and microscanline) Rivas, E., Smith, M., 2017. FracPaQ: a MATLABTM toolbox for the quantification of
integrated in the suggested workflow, are an important tool for col- fracture patterns. J. Struct. Geol. 95, 1–16. https://doi.org/10.1016/j.jsg.2016.12.
003.
lecting fracture attributes, which can be used to model a more reliable Heimhofer, U., Ariztegui, D., Lenniger, M., Hesselbo, S.P., Martill, D.M., Rios-Netto, A.M.,
DFN, providing important equivalent properties, after upscaling it, for 2010. Deciphering the depositional environment of the laminated Crato fossil beds
upcoming fluid flow simulations of carbonate reservoirs. (early cretaceous, Araripe Basin, north-eastern Brazil). Sedimentology 57, 677–694.
https://doi.org/10.1111/j.1365-3091.2009.01114.x.
Hooker, J.N., Gale, J.F.W., Gomez, L.A., Laubach, S.E., Marrett, R., Reed, R.M., 2009.
Acknowledgements Aperture-size scaling variations in a low-strain opening-mode fracture set, Cozzette
Sandstone, Colorado. J. Struct. Geol. 31, 707–718. https://doi.org/10.1016/j.jsg.
2009.04.001.
Thanks to Dr. Stephen Laubach for comments on an early version of Hooker, J.N., Laubach, S.E., Marrett, R., 2014. A universal power-law scaling exponent
this paper. We thank Gabriel Matos, Lizzy Micknnon, Jorge André Braz for fracture apertures in sandstones. Bull. Geol. Soc. Am. 126, 1340–1362. https://
and Randy Marrett for the field work and fruitful discussions. This study doi.org/10.1130/B30945.1.
Hooker, J.N., Laubach, S.E., Marrett, R., 2013. Fracture-aperture size-frequency, spatial
was carried out with financial support provided by a grant from the
distribution, and growth processes in strata-bounded and non-strata-bounded frac-
Projects Turing and Crato – Petrobras/Federal University of tures, Cambrian Mesón Group, NW Argentina. J. Struct. Geol. 54, 54–71. https://doi.
Pernambuco (UFPE). We are grateful to Fracture Research and org/10.1016/j.jsg.2013.06.011.
Laubach, S.E., Lamarche, J., Gauthier, B.D.M., Dunne, W.M., Sanderson, D.J., 2018a.
Application Consortium, University of Texas at Austin, Petrobras,
Spatial arrangement of faults and opening-mode fractures. J. Struct. Geol. 108, 2–15.
National Agency of Petroleum, Natural Gas and Biofuels of Brazil https://doi.org/10.1016/j.jsg.2017.08.008.
(ANP), Science without Borders Program, and Department of Geology- Laubach, S.E., Hundley, T.H., Hooker, J.N., Marrett, R.A., 2018b. Spatial arrangement
UFPE. We thank Haydon Mort (Geologize) for useful scientific and and size distribution of normal faults, Buckskin detachment upper plate, Western
Arizona. J. Struct. Geol. 108, 230–242. https://doi.org/10.1016/j.jsg.2017.10.001.
English considerations. We gratefully acknowledge the anonymous re- Laubach, S.E., Olson, J.E., Gross, M.R., 2009. Mechanical and fracture stratigraphy. AAPG
viewers, whose constructive reviews improved the manuscript. (Am. Assoc. Pet. Geol.) Bull. 93 (11), 1413–1426. https://doi.org/10.1306/
07270909094.
Laubach, S.E., Mace, R.E., Nance, H.S., 1995. Fault and joint swarms in a normal fault
References zone. In: Rossmanith, H.-P. (Ed.), Mechanics of Jointed and Faulted Rock. Balkema,
Rotterdam, pp. 305–309.
Assine, M.L., 2007. Bacia do Araripe. Bol. Geociencias Petrobras 15, 371–389. Li, J.Z., Laubach, S.E., Gale, J.F.W., Marrett, R.A., 2018. Quantifying opening-mode
Barling, N., Martill, D.M., Heads, S.W., Gallien, F., 2015. High fidelity preservation of fracture spatial organization in horizontal wellbore image logs, core and outcrop:
fossil insects from the Crato Formation (lower cretaceous) of Brazil. Cretac. Res. 52, application to Upper Cretaceous Frontier Formation tight gas sandstones, USA. J.
605–622. https://doi.org/10.1016/j.cretres.2014.05.007. Struct. Geol. 108. https://doi.org/10.1016/j.jsg.2017.07.005.
Belgrano, T.M., Herwegh, M., Berger, A., 2016. Inherited structural controls on fault Lu, Y.C., Tien, Y.M., Juang, C.H., 2017. Uncertainty of 1-D fracture intensity measure-
geometry, architecture and hydrothermal activity: an example from Grimsel Pass, ments. J. Geophys. Res.: Solid Earth 122, 9344–9358. https://doi.org/10.1002/
Switzerland. Swiss J. Geosci. 109 (3), 345–364. 2016JB013620.
Bisdom, K., Bertotti, G., Bezerra, F.H., 2017. Inter-well scale natural fracture geometry Marrett, R., 1996. Aggregate properties of fracture populations. J. Struct. Geol. 18,
and permeability variations in low-deformation carbonate rocks. J. Struct. Geol. 97, 169–178. https://doi.org/10.1016/S0191-8141(96)80042-3.
23–36. https://doi.org/10.1016/j.jsg.2017.02.011. Marrett, R., Gale, J.F.W., Gómez, L.A., Laubach, S.E., 2018. Correlation analysis of
Blessent, D., Therrien, R., MacQuarrie, K., 2009. Coupling geological and numerical fracture arrangement in space. J. Struct. Geol. 108, 16–33. https://doi.org/10.1016/
models to simulate groundwater flow and contaminant transport in fractured media. j.jsg.2017.06.012.
Comput. Geosci. 35, 1897–1906. https://doi.org/10.1016/j.cageo.2008.12.008. Marrett, R., Ortega, O.J., Kelsey, C.M., 1999. Extent of power-law scaling for natural
Bonnet, E., Bour, O., Odling, N.E., Davy, P., Main, I., Cowie, P., Berkowitz, B., 2001. fractures in rock. Geology 27, 799–802. < 0799:EOPLSF > 2.3.CO. https://doi.org/
Scaling of fracture systems in geological media. Rev. Geophys. 39 (3), 347–383. 10.1130/0091-7613(1999)027.

566
T.S. Miranda et al. Marine and Petroleum Geology 97 (2018) 556–567

Massaro, L., Corradetti, A., Vinci, F., Tavani, S., Iannace, A., Parente, M., Mazzoli, S., of methods to detect mechanically recovered meat in meat products—IV:
2018. Multiscale fracture analysis in a reservoir-scale carbonate platform exposure Immunology. Meat Sci. 40, 327–336.
(Sorrento Peninsula, Italy): implications for fluid flow. Geofluids 2018. https://doi. Priest, S.D., 1993. Discontinuity Analysis for Rock Engineering. Chapman and Hall,
org/10.1155/2018/7526425. London 473p.
Menezes De Jesus, C., Luiz, A., Compan, M., Surmas, R., 2016. Permeability estimation Questiaux, J.M., Couples, G.D., Ruby, N., 2010. Fractured reservoirs with fracture cor-
using ultrasonic borehole image logs in dual-porosity carbonate reservoirs. ridors. Geophys. Prospect. 58, 279–295.
Petrophysics 57, 620–637. Sanderson, D.J., Nixon, C.W., 2015. The use of topology in fracture network character-
Milliken, K.L., Laubach, S.E., 2000. Brittle deformation in sandstone diagenesis revealed ization. J. Struct. Geol. 72, 55–66. https://doi.org/10.1016/j.jsg.2015.01.005.
by scanned cathodoluminescence imaging with application to characterization of Santos, R.F.V.C., Miranda, T.S., Barbosa, J.A., Gomes, I.F., Matos, G.C., Gale, J.,
fractured reservoirs. In: Pagel, M., Barbin, V., Blanc, P., Ohnestetter, D. (Eds.), Neumann, V.H., Guimarães, L.J., 2015. Characterization of natural fracture systems:
Cathodoluminescence in Geosciences. Springer Verlag, Berlin, pp. 225–243. analysis of uncertainty effects in linear scanline results. AAPG (Am. Assoc. Pet. Geol.)
Miranda, T., Barbosa, J.A., Gomes, I.F., Soares, A., Santos, R.F.V.C., Matos, G.C., Bull. 99 (12), 2203–2219. https://doi.org/10.1306/05211514104.
McKinnon, E.A., Neumann, V.H.M.L., Marrett, R.A., 2016. Petrophysics and petro- Terra, J.G.S., Spadini, A.R., França, A.B., Leite, C., Zambonato, E.E., Costa, L., Arienti,
graphy of aptian tight carbonate reservoir, Araripe Basin, NE Brazil. https://doi.org/ L.M., Erthal, M.M., Blauth, M., Franco, M.P., Matsuda, N.S., Goulart, N., Augusto, P.,
10.3997/2214-4609.201600890. Junior, M., Francisco, R.S., Avila, D., Souza, R.S., De, Tonietto, S.N., Maria, S., 2010.
Miranda, T.,S., 2015. Caracterização Geológica e Geomecânica dos Depósitos Classificações Clássicas De Rochas Carbonáticas, vol 18. B. Geoci. Petrobras, Rio
Carbonáticos e Evaporíticos da bacia do Araripe. Tese de Doutorado, NE Brasil 259p. Janeiro, pp. 9–29.
National Research Council, 1996. Rock Fractures and Fluid Flow: Contemporary Terzaghi, R.D., 1965. Sources of error in joint surveys. Geotechnique 15, 287–304.
Understanding and Applications. Committee on Fracture Characterization and Fluid https://doi.org/10.1680/geot.1965.15.3.287.
Flow. National Academy Press, Washington DC 551 p. Ukar, E., Laubach, S.E., 2016. Syn- and postkinematic cement textures in fractured car-
Nelson, R.A., 2001. Geologic Analysis of Naturally Fractures Reservoirs, second ed. Gulf bonate rocks: insights from advanced cathodoluminescence imaging. Tectonophysics
Professional Publishing Co. 2001. 690, 190–205. https://doi.org/10.1016/j.tecto.2016.05.001.
Neumann, V.H., 1999. Estratigrafía, Sedimentología, Geoquímica y Diagénesis de los Urai, J.L., Williams, P.F., van Roermund, H.L.M., 1991. Kinematics of crystal growth in
sistemas lacustres Aptiense – Albienses de la Cuenca de Araripe (NE Brasil). PhD syntectonic fibrous veins. J. Struct. Geol. 13, 823–836.
Thesis. Barcelona University 233 p. Vermilye, J.M., Scholz, C.H., 1995. Relation between vein length and aperture. J. Struct.
Olson, J.E., 2003. Sublinear scaling of fracture aperture versus length: an exception or the Geol. 17 (3), 423–434 1995.
rule? J. Geophys. Res. Solid Earth 108. https://doi.org/10.1029/2001JB000419. Watkins, H., Bond, C.E., Healy, D., Butler, R.W.H., 2015. Appraisal of fracture sampling
Ortega, O.J., Gale, J.F.W., Marrett, R., 2013. Quantifying diagenetic and stratigraphic methods and a new workflow to characterise heterogeneous fracture networks at
controls on fracture intensity in platform carbonates: an example from the Sierra outcrop. J. Struct. Geol. 72, 67–82. https://doi.org/10.1016/j.jsg.2015.02.001.
Madre Oriental, northeast Mexico. J. Struct. Geol. 32, 1943–1959. Zihms, S.G., Miranda, T., Lewis, H., Barbosa, J.A., Neumann, V.H., Souza, J.A.B., Cruz
Ortega, O.J., Marrett, R.A., Laubach, S.E., 2006. A scale-independent approach to fracture Falcão, T., 2017. Crato Formation laminites - a representative geomechanical pre-salt
intensity and average spacing measurement. Am. Assoc. Petrol. Geol. Bull. 90, analogue? representative geomechanical. In: 5th Atlantic Conjugate Margins
193–208. https://doi.org/10.1306/08250505059. Conference.
Pickering, K., Griffin, M., Smethurst, P., Hargin, K.D., Stewart, C. a., 1995. Investigation

567

You might also like