Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Structural Geology 74 (2015) 31e44

Contents lists available at ScienceDirect

Journal of Structural Geology


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

Non-linear growth kinematics of opening-mode fractures


Yaser Alzayer, Peter Eichhubl*, Stephen E. Laubach
Bureau of Economic Geology, Jackson School of Geosciences, The University of Texas at Austin, TX 78758, USA

a r t i c l e i n f o a b s t r a c t

Article history: Fracture aperture is commonly assumed to be a linear function of fracture length, stress, and elastic
Received 11 September 2014 material properties. Under constant stress, fracture opening displacement may change without con-
Received in revised form current length or height growth if the material effectively weakens after initial linear elastic fracture
2 February 2015
growth by changes in elastic properties or by non-elastic deformation processes. To investigate the ki-
Accepted 6 February 2015
nematics of fracture opening and its dependence on length and height growth, we reconstructed the
Available online 5 March 2015
opening history of three opening-mode fractures that are bridged by crack-seal quartz cement. Similar
crack-seal cement bridges had been interpreted to form by repeated incremental fracture opening and
Keywords:
Fracture
subsequent precipitation of quartz cement. Using scanning electron microscope cathodoluminescence
Diagenesis imaging, we determined the thickness and number of crack-seal cement layers as a function of position
Crack-seal along fracture length and height. Observed trends in crack-seal cement layer thickness and number of
Creep fracture opening increments are consistent with non-linear fracture growth kinematics, consisting of an
Sandstone initial stage of fast fracture propagation relative to aperture growth, followed by a stage of slow prop-
Quartz agation and pronounced aperture growth. Consistent with earlier fluid inclusion observations indicating
fracture opening and propagation occurring over 40e50 m.y., we interpret the second stage of pro-
nounced aperture growth and slow propagation to result from fracture opening strain accommodated by
solution-precipitation creep and concurrent slow, possibly subcritical, fracture propagation.
© 2015 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction kinematic aperture growth to fracture length and height growth,


where length and height designate horizontal and vertical fracture
Fractures and fracture networks control fluid flow and mass and dimension, respectively.
heat transfer in the upper crust, and significantly govern its me- Guided by the successful application of engineering linear elastic
chanical properties (National Research Council, 1996). Properties fracture mechanics in geologic systems, the aperture of opening-
that control fluid flow in single fractures include aperture, length mode or extension fractures is generally assumed to follow a
and height, connectivity with other fractures, fracture wall rough- linear function of fracture length, rock elastic properties, pore fluid
ness, connectivity of the fracture space to the pore space in the host pressure, and total normal stress acting on the fracture, the last two
rock, and host rock permeability (Lee and Farmer, 1993; Klimczak parameters combined into an effective loading stress (Secor, 1965;
et al., 2010; Tokan-Lawal et al., 2014; Landry et al., 2014). Where Segall, 1984; Pollard and Aydin, 1988; Engelder et al., 1993). Field
not occluded by fracture mineralization, fracture aperture provides and lab measurements of opening displacement profiles measured
primary control on permeability of open, i.e. well connected, frac- along the trace length of opening-mode fractures have found a first-
tures (Snow, 1968). For partially mineralized fractures, flow prop- order agreement with the predicted elliptical fracture profile for
erties are governed by the hydraulic or open fracture aperture. In elastic fractures (Delaney and Pollard, 1981; Pollard and Segall,
this study, we are concerned with growth in kinematic aperture, i.e. 1987; Gudmundsson, 2011). Fracture aperture, however, is
the kinematic separation of both fracture walls regardless of con- frequently observed to exceed the predicted magnitude for elastic
current or subsequent mineral cement infill, and the relation in fractures for elastic and fracture toughness parameters represen-
tative of the host rock. For instance, for elastic and fracture prop-
erties typical of well-cemented sedimentary rock (Young's
modulus: 50 GPa, Poisson ratio 0.25, fracture toughness
* Corresponding author. Bureau of Economic Geology, The University of Texas at 1e2 MPa m½; Atkinson and Meredith, 1987), the predicted aperture
Austin, University Station Box X, Austin, TX 78713-8924, USA. Tel.: þ1 512 475 8829.
of a 10 cm-long opening-mode fracture is between 10 and 30 mm
E-mail address: peter.eichhubl@beg.utexas.edu (P. Eichhubl).

http://dx.doi.org/10.1016/j.jsg.2015.02.003
0191-8141/© 2015 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
32 Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44

(after Pollard and Segall, 1987; see also Section 4.2). Apertures of the 1980's. The well is located in the East Texas Basin, on the SW
opening-mode fractures of similar length observed in sandstone flank of the Sabine Arch, a low-amplitude north-trending anticline
reservoirs frequently exceed these values by a factor of 100e1000 (Fig. 1) (Laubach and Jackson, 1990). The well is located in gently
(Laubach, 1989; Vermilye and Scholz, 1995; Stowell et al., 1999). dipping strata, distant from normal faults with small displacement,
Based on observations of fracture opening occurring concur- and open folds.
rently with diagenetic dissolution-precipitation reactions in Fractures for this study were collected from the Lower Creta-
adjoining host rock, Eichhubl et al. (2001), Eichhubl and Aydin ceous Travis Peak Formation, a gas-producing unit consisting of
(2003), and Eichhubl (2004) proposed that the opening displace- mainly quartzarenite and subarkose interbedded with mudstone
ment of such wide-aperture fractures is accommodated by (Dutton and Land, 1988). Significant porosity and permeability loss
dissolution-precipitation creep in the host rock. In this case, it is is caused by pervasive quartz cementation, resulting in average
conceivable that fracture aperture growth relates to length or porosity of 8.8% and core plug permeability of less than 0.1 mD
height growth in a non-linear way. For example, fractures may (Holditch et al., 1985; Dutton, 1986; Dutton and Land, 1988). Quartz
reach an initial length or height to aperture aspect ratio by elastic cementation was noted to increase with increasing present-day
fracture growth, with subsequent solution-precipitation creep depth (Dutton and Diggs, 1990; Soeder and Chowdiah, 1990).
increasing fracture aperture without additional length growth. Quartz cementation gradually increases with burial depth within
Solution-precipitation creep would effectively weaken the rock, the Travis Peak and ranges from 16% at the top of the formation to
equivalent to an increase in elastic compliance subsequent to 19% at the bottom (Dutton and Diggs, 1990).
fracture growth in purely elastic media. Alternatively, fracture The sand-rich intervals of the Travis Peak Formation contain
opening displacement may be accommodated entirely by non- cemented and partially cemented vertical opening-mode fractures
elastic creep processes. striking east-northeast (Laubach, 1988). Fracture mineralization is
This study was designed to test such non-linear rock fracture dominated by crack-seal and euhedral quartz with small amount of
growth models by comparing the fracture opening displacement late ankerite and clay. Ratios in natural fracture height to aperture
recorded in crack-seal fracture cement along fracture length and in the Travis Peak Formation range from 30 to 6000 (Laubach,
height. Crack-seal cement forms by repeated episodes or stages of 1989). Fractures with apertures >0.1 mm tend to be only partially
fracture opening and subsequent fracture cementation resulting in mineralized with remaining porosity between quartz cement
banded cement textures (Taber, 1916; Hulin, 1929; Ramsay, 1980). bridges (Laubach, 2003). Fluid inclusion microthermometry corre-
Crack-seal cement is considered synkinematic relative to the frac- lated to a burial curve for the Travis Peak Formation at the nearby
ture opening. The textural and compositional record contained in SFE2 well (Fig. 1) suggested that fractures initiated near maximum
crack-seal cement thus represents a record of fracture opening ki- burial depth around 48 Ma and continued to grow to present day
nematics (Durney and Ramsay, 1973; Cox and Etheridge, 1983; Cox, (Becker et al., 2010). The maximum burial depth of the Travis Peak
1987; Urai et al., 1991; Hilgers and Urai, 2002; Renard et al., 2005; at the S.F.O.T. 1 well was 3.2 km corresponding to a maximum
Bons et al., 2012). Microthermometric and Raman probe analyses of temperature of 145 C at the top of the formation. Currently, the top
fluid inclusions trapped during the crack-seal cement growth allow of the Travis Peak Formation is at a depth of 2.7 km (8957 ft) (CER,
reconstruction of pore fluid chemical, temperature, and pressure 1985). The fractures in this study come from a depth of approxi-
conditions during fracture opening (Parris et al., 2003; Hanks et al., mately 3 km (~10,100 ft).
2006; Becker et al., 2010; Van Noten et al., 2011; Duncan et al.,
2012; Fall et al., 2012, 2015). In quartz fracture cement, crack-seal 3. Methods
textures are best observed in scanning electron microscope-
cathodoluminescence (SEM-CL), with individual crack-seal To allow a comparison between bridges in the center of fractures
cement layers, also referred to as gap deposits (Laubach, 2004a), and at the tip, we collected 7 samples of fractures with preserved
highlighted by variations in CL intensity and color (Dietrich and fracture tips. We sampled only fractures that were observed under
Grant, 1985; Laubach et al., 2004b). a binocular microscope to be partially or completely filled with
In sandstone that has experienced burial temperatures in excess quartz cement thus avoiding barren fractures formed during core
of ~80  C, crack-seal quartz cement in partially cemented fractures extraction and handling. Because most macroscopic fractures
forms isolated cement bridges that connect both fracture walls observed in this core have lengths greater than the 4-inch (10 cm)
surrounded by residual fracture porosity (Laubach, 1988, 2004b; core diameter, all fractures sampled for this study contained only
Lander and Laubach, 2014). Based on microthermometric and one tip (Fig. 2A). Thus, the total length of these fractures is not
Raman microprobe studies of fluid-inclusions trapped during known. Crack-seal bridges were initially identified in transmitted
crack-seal cement growth we have found that crack-seal bridges light petrography by their characteristic fluid inclusion trails
track a protracted history of fracture opening extending over mil- (Fig. 2B). Bridge crack-seal textures were imaged using a Phillips
lions to tens of millions of years (Becker et al., 2010; Fall et al., 2012, XL30 SEM equipped with an Oxford Instruments MonoCL system
2015). operated at 15 kV. Panchromatic CL images were obtained for each
To investigate the fracture opening and length and height bridge by taking several images at 500 magnification with no
growth kinematics of opening-mode fractures, we performed a color filter at 400 ms dwell time following techniques described by
SEM-CL textural analysis of crack-seal cement bridges sampled Reed and Milliken (2003) and Laubach et al. (2004b). Grayscale
along the length and height from the fracture center to the tip. photo mosaics for ̊ each bridge were created using Adobe Photoshop
Fractures for this analysis were collected from core in the Travis (Fig. 2C). Images were digitally sharpened and corrected for optimal
Peak Formation in East Texas. contrast and exposure.
For each of the 38 bridges that we imaged, we measured the
2. Geologic background thickness of 3e44 crack-seal cement layers (10 or more cement
layers for 35 bridges) that we determined to be intact, i.e. cement
Fracture analyzed in this study were collected from core of the layers that were not split by subsequent crack-seal. This determi-
Ashland Exploration, S. F. O. T. No. 1 well, Nacogdoches, N. W. Field, nation was based on the symmetry of the cement texture within
Nacogdoches County, Texas (Fig. 1). This well was drilled as part of each increment around its medial line or suture as marked by fluid
the Gas Research Institute extensive tight gas drilling program in inclusion trails. In CL-SEM, the cement layers are separated by dark
Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44 33

Fig. 1. Location of the SFOT-1 and SFE-2 wells located southwest of the Sabine Arch in east Texas.

lines (Fig. 2E). Measurements of increment thickness were obtained and Becker et al. (2010). These techniques utilize cross-cutting and
using a digital ruler calibrated to the SEM image scale. The mea- overlapping relationships between crack-seal cement and
surement error is ±1 mm. We also measured the kinematic aperture, bordering lateral euhedral cement to determine the relative
i.e. the distance from fracture wall to wall orthogonal to the fracture sequence of crack-seal cement deposits (Fig. 2). In this recon-
trace, at the location of each bridge. The number of fracture opening struction, we attempted to identify and account for split crack-seal
increments at each bridge location was estimated by dividing the cement layers. Because relative timing difference is difficult to
kinematic aperture by the average crack-seal cement layer discern for adjacent crack-seal cement layers, groups of roughly
thickness. contemporaneous cement layers were lumped into larger bridge
A direct count of the number of crack-seal cement layers would growth segments (Fig. 2D). For these two bridges, we measured the
tend to provide an inaccurate count of the number of fracture reconstructed thicknesses of all cement layers (69 and 80 in each
opening increments for two reasons: (1) The crack-seal texture of bridge) and compared this record to the calculated number of
some bridges is only incompletely imaged where the long axis of fracture opening increments based on the average crack-seal
the bridge is oriented oblique to the thin section and thus imaged cement layer thickness.
only partway across the fracture width. In this case, the imaged
crack-seal cement layer count underestimates the total number of 4. Results
crack-seal opening increments across the width of the fracture. (2)
Continued crack-seal process breaks earlier-formed crack-seal Of seven fracture samples collected, three partially cemented
cement. Thus, younger crack-seal opening increments frequently vertical fractures of Travis Peak Formation the SFOT-1 core were
split earlier crack-seal cement layers into what appears to be two or found to contain a sufficient number of bridges suitable for fracture
more cement layers. Without accounting for split cement layers, a growth analysis. Two of these three fractures are closely-spaced en
direct count of crack-seal cement layers would overestimate the chelon fractures sampled for thin sections along their length or
e
number of fracture opening increments. The number of fracture horizontal extent (perpendicular to the vertical core axis and par-
opening increments at each bridge location was thus estimated by allel to bedding); one fracture was sampled along its height or
dividing the kinematic aperture by the average size of complete vertical extent (parallel to core axis and perpendicular to bedding).
crack-seal cement layers. For these three fractures, we imaged 38 crack-seal cement bridges
In addition to determining the thickness of crack-seal cement and calculated the number of fracture opening increments based on
layers and the number of fracture opening increments, we pal- the kinematic aperture and average crack-seal cement layer thick-
inspastically reconstructed two bridges that provided a complete ness. Bridges were selected to sample the fractures along transects
crack-seal cement record over the entire aperture of the fracture from the fracture tips to a position approximately in the center of
using restoration techniques summarized by Laubach et al. (2004a) the fractures.
34
Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44
Fig. 2. (A) Core image of fractures in Travis Peak Formation, SFOT-1 well, measured depth of 10,107 feet (3080.6 m). (B) Transmitted light photomicrograph (plane polarized) of partially cemented fracture. Quartz fracture cement bridge
indicated with red rectangle. Blue: Residual fracture porosity filled with epoxy. (C) SEM-cathodoluminescence image of quartz cement bridge outlined in B. (D) Interpreted map of the quartz bridge indicating stages of bridge cement
growth, numbered from oldest (1) to youngest (11). (E) Detail of C illustrating the symmetry of the internal cement texture of intact crack-seal cement layers around their medial line, in contrast to asymmetric layers interpreted to be
split by later crack-seal cement growth.
Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44
chelon fractures imaged along strike (SFOT-1 well: depth 10,108.3 feet, 3081.0 m). Blue is epoxy filling remaining fracture porosity. Red rectangles indicate
Fig. 3. (A) Photomicrograph (plane polarized) of two partially cemented en e
analyzed crack-seal cement bridges. (B) Average crack-seal cement layer thickness for positions along fracture dip indicated by red rectangles in A. Error bars indicate standard error of the mean. (C) Calculated number of fracture
opening increments for positions along fracture dip indicated by red rectangles in A. Error bars indicate predicted range in number of fracture opening increments for standard error given in B.

35
36 Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44

All three transects revealed an increase in average crack-seal The calculated number of fracture opening increments ranges be-
cement thickness and in the number of fracture opening in- tween 16 and 105, with the majority of cement bridges away from
crements for a limited distance away from fracture tips before fracture tips containing 70e95 fracture opening increments.
reaching a roughly constant value of both parameters for the For the two cement bridges that were reconstructed, crack-seal
remainder of the fracture (Figs. 3 and 4). Average crack-seal cement cement layer thickness is variable over the growth history of the
layer thickness for cement bridges ranges between 3 and 14 mm. bridges, ranging between 1 and 30 mm (Figs. 2, 5 and 6). No

Fig. 4. (A) Photomicrograph (plane polarized) of a partially cemented fracture imaged parallel to the fracture dip (parallel to the core axis) (SFOT-1 well, depth of 10,107 feet,
3080.6 m). Red rectangles indicate analyzed crack-seal cement bridges. (B) Average crack-seal cement layer thickness for positions along fracture dip indicated by red rectangles in
A. Error bars indicate standard error of the mean. (C) Calculated number of fracture opening increments for positions along fracture dip indicated by red rectangles in A. Error bars
indicate predicted range in number of fracture opening increments for standard error given in B.
Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44 37

Fig. 5. (A) Reconstructed crack-seal cement sequence based on SEM-cathodoluminescence image on bridge 4855 mm from tip in Fig. 4 (sample SFOT-1-10107 feet, 3080.6 m).
Selected stages of bridge reconstruction labeled by number of increments. See Fig. 2 for color scheme of interpreted bridge. (B) Crack-seal cement layer thicknesses plotted in order
of relative time of formation.
38 Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44

systematic trend toward wider or narrower cement layer thickness critically loaded linear-elastic fracture and assuming quasi-static
can be discerned from older to younger increment across the fracture growth, the width of fracture opening increments and
bridges (Figs. 5B and 6B). thus crack-seal cement layer thickness within a cement bridge is
expected to decrease over time as the driving stress decreases with
5. Discussion increasing fracture length (Engelder et al., 1993; see also Section
4.2) (Fig. 8A). (2) The second model assumes that fracture growth
5.1. Geometric analysis of fracture aperture growth occurs by initial tip propagation and concurrent aperture growth to
a finite fracture length and height (stage 1), followed by a second
Four geometric growth models for opening-mode fractures stage of aperture growth without further propagation (Fig. 7B). If
were tested against the crack-seal cement layer thickness and the stage of initial fracture propagation resulted in only few or, as
fracture opening increment data obtained from the SFOT-1 crack- shown in Fig. 7B, in a single aperture growth increment, with the
seal cement bridges (Fig. 7): (1) The first model considers fracture majority of aperture growth increments occurring during the sec-
length and height growth occurring by tip propagation with con- ond stage, this type of fracture would result in a roughly constant
current aperture growth as expected for critical fracture propaga- number of fracture opening increments along the length and height
tion in a linear elastic material (Fig. 7A). In this model, the number of the fracture, with narrower increments and thinner crack-seal
of crack-seal fracture opening increments, shown in Fig. 7 as out- cement layers near the tip and wider increments and thicker
lines of successive stages of fracture propagation, is expected to cement layers near the center of the fracture (Fig. 8B). Aperture
decrease linearly toward the fracture tip while the average width of growth without propagation, referred to as a stationary fracture,
opening increments increases toward the tip (Fig. 8A). For a may be envisioned as reloading of a previously formed, then

Fig. 6. (A) Reconstructed crack-seal cement sequence based on SEM-cathodoluminescence image of bridge 4187 mm from tip in Fig. 4 (sample SFOT-1-10107 feet, 3080.6 m).
Selected stages of bridge reconstruction labeled by number of increments. See Fig. 2 for color scheme of interpreted bridge. (B) Crack-seal cement layer thicknesses plotted in order
of relative time of formation.
Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44 39

unloaded fracture, with reloading stress conditions remaining opening increments along the fracture would be roughly constant
below the critical stress for further propagation. Alternatively, a along the fracture length and height. Opening increments and
stationary fracture could increase in aperture under constant stress crack-seal cement layers would be thinner at the fracture tip
with a change in elastic material properties or accommodation of compared to the fracture center (Fig. 8C). (4) The fourth model
fracture opening displacement by non-elastic deformation mech- assumes fracture propagation to a finite fracture height and
anisms such as solution-precipitation creep (see discussion sec- continued growth in length. This model reflects fracture growth in
tion). (3) The third model, a variant of model 2, includes initial rapid layered sedimentary rock where bedding interfaces limit height
length and height propagation with concurrent aperture growth growth (Fig. 7D) (Corbett et al., 1987; Helgeson and Aydin, 1991;
(stage 1), followed by a stage 2 of slow propagation and pro- Cooke and Underwood, 2001; Rijken and Cooke, 2001;
nounced aperture growth (Fig. 7C). As in model 2, the number of Underwood et al., 2003; Laubach et al., 2009). Because the short

Fig. 7. Conceptual models for fracture growth kinematics. (A) Linear elastic circular or penny-shaped fracture propagates both in height and length with concurrent proportional
aperture growth. (B) Stationary circular fracture with aperture growth without tip propagation. (C) Non-linear two-stage growth of a circular fracture with initially fast propagation
with concurrent lesser aperture growth followed by slow propagation and pronounced aperture growth. (D) Fracture height growth limited by mechanical layer boundaries while
horizontal propagation continues resulting in an elongate fracture shape; aperture growth in center of fracture is limited by restricted height growth. Opening kinematics of Travis
Peak fractures resemble non-linear growth model in C.
40 Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44
Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44 41

fracture dimension controls the fracture aperture (Gudmundsson, opening displacement for the two reconstructed bridges against
2000), the aperture at the fracture center will stop growing once theoretical predictions based on linear elastic fracture mechanics
the final fracture height has been reached provided the elastic rock (LEFM) (Fig. 9). Assuming constant elastic material properties, two
properties remain constant. For such a blade- or tunnel-shaped scenarios of aperture growth are distinguished for this analysis: (1)
fracture, the number of fracture opening increments would be Fracture growth by tip propagation with concurrent aperture
constant along the fracture length or strike direction with excep- opening (propagating fracture), here assumed to propagate at
tion of the near-tip region. critical loading conditions with KI ¼ KIc where KI is the mode-I
Observed trends in number of fracture opening increments in the stress concentration at the fracture tip, and KIc the mode-I frac-
SFOT-1 fractures decreasing toward the fracture tips (Figs. 3 and 4) ture toughness, a material property that quantifies the material
are compatible with fracture propagation and thus with the first and resistance to fracture propagation; this scenario is equivalent to
third fracture growth models but inconsistent with the second model 1 above. (2) Fracture aperture opening without tip propa-
model which assumes a stationary tip during the second stage of gation (stationary fracture). In this case, the fracture is assumed to
aperture growth (Fig. 7). The lack of a trend toward decreasing crack- have initially propagated to its finite length by either subcritical
seal cement layer thickness with time for any position along the growth, or by critical growth followed by elastic unloading of the
fracture transect as observed in the two bridge reconstructions fracture. Any further loading that results in aperture growth
(Figs. 5B and 6B) is inconsistent with the first model which predicts remained below the critical loading stress for propagation. This
opening increments at any position along the fracture that decrease scenario is equivalent to model 2 above but can also approximate
in size over time. Model 1 also invokes wider fracture opening in- conditions during stage 2 of fracture growth in model 3.
crements in the tip region than in the fracture interior which is not The maximum opening displacement umax, equivalent to half
consistent with the observed decrease in average crack-seal cement the maximum aperture, of a stationary circular or penny-shaped
layer thickness toward tips. Instead, the observed decrease in both mode-I fracture is given by
the number of fracture opening increments and in average crack-seal  
cement layer thickness at fracture tips is consistent with slow 4 1  n2 sc
umax ¼ (4-1)
propagation and enhanced aperture growth as predicted in model 3. pE
Model 4 of height-limited fracture growth fails to account for the
observed decrease in the average crack-seal cement layer thickness where s is the effective loading stress (effective tension positive), y
toward the fracture tip in the vertical fracture transect which we is Poisson's ratio, c is the fracture half-length, and E is Young's
consider evidence of aperture growth with no or only slow vertical modulus (Gudmundsson, 2011, p. 266e267).
fracture propagation. This is because the height-limited fracture For an opening-mode fracture that is critically loaded,
would stop growing in aperture once the bed height is reached.
K
However, height-limited fracture growth may be invoked for the s ¼ sIc ¼ pIcffiffiffiffiffiffi (4-2)
second fracture growth stage in model 3. The wide range in fracture Y pc
height to aperture ratio of 30e6000 observed in the Travis Peak
Formation (Laubach, 1989) can be explained by model 3, with the where sIc is the critical loading stress for mode I fracture propa-
growth history of narrow fractures limited to stage 1 of initial fast gation, KIc is the mode-I fracture toughness, and Y is a shape factor
propagation with concurrent aperture growth and no or only limited (2/p for a penny-shaped fracture).
stage 2 fracture growth of slow propagation and pronounced aper- The opening displacement profile can be obtained through
ture growth. Fractures with high aspect ratios represent both growth rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
x2
stages. We thus consider the two-stage fracture growth model 3, ux ¼ umax 1 (4-3)
with an initial stage of fast propagation followed by a stage of slow c
propagation relative to the rate of aperture growth to account for the
observed fracture opening displacement data. where ux is the opening displacement at position x measured from
The apparently random and fluctuating crack-seal cement layer the center of the fracture, i.e. from the location of maximum
thickness in the reconstructed bridges (Figs. 5B and 6B) may be aperture.
evidence for the role of changing pore pressure over time in frac- From Eqs. (4-2) it follows that, for fracture aperture growth
ture growth. A similar fluctuating crack-seal cement layer thickness concurrent with critical but stable quasi-static tip propagation, the
was also reported in the Travis Peak and other formations (Laubach fracture driving stress required for propagation decreases with
et al., 2004a). Variation in pore pressure has been interpreted increasing fracture length. Increments in fracture opening
previously to have caused cyclic fracture growth (Bahat and displacement per unit length growth at any position x along the
Engelder, 1984; Lacazette and Engelder, 1992), and records of fracture thus decrease in size with increasing fracture length
cyclically variable pore fluid pressure have been obtained through (Fig. 8A). Consequently, the slope of the cumulative opening
fluid inclusion analyses in crack-seal fracture cements (Fall et al., displacement at any position x decreases with time (Fig. 9, curves
2012, 2015). labeled propagating fracture). In contrast, for the stationary frac-
ture, the fracture opening displacement scales linearly with the
loading stress. At any position x, the slope in cumulative opening
5.2. Comparison to linear elastic fracture mechanics predictions displacement over time is constant (Fig. 9, curves labeled stationary
fracture).
For a quantitative analysis of the SFOT-1 fracture opening The observed cumulative fracture opening displacement for the
displacement results we compared measured cumulative fracture two reconstructed cement bridges, showing a near-linear increase

Fig. 8. Predicted trends in width and number of fracture opening displacement increments for fracture kinematic models. Aperture profiles (left diagrams) shown for half-fracture
only. (A) Proportional tip propagation and aperture growth characteristic of linear elastic fracture growth propagating under critical quasi-static loading conditions. Compare to
Fig. 7A. (B) Aperture opening with no propagation (stationary fracture). Compare to Fig. 7B. (C) Two-stage growth model consisting of fast initial propagation with concurrent lesser
aperture growth followed by slow propagation and pronounced aperture growth (non-linear fracture growth). Compare to Fig. 7C. Fracture is envisioned to propagate subcritically
during stage 2.
42 Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44

Fig. 9. Cumulative fracture opening displacement for the reconstructed bridge in Fig. 5. Shown for comparison are predicted cumulative opening displacements for propagating and
stationary fractures calculated using material properties in Table 1.

in fracture opening displacement with time, follows the trend fractures thus providing maximum apertures as a function of dis-
predicted by the stationary fracture model but is inconsistent with tance from the fracture tip along the observed fracture profiles. To
the propagating fracture (Fig. 9). This suggests that the initial stage match the observed fracture apertures, a second set of curves was
of critical fracture propagation is either not recorded in the two calculated with an “effective” best-fit Young's modulus
reconstructed crack-seal cement bridges, or in too few crack-seal Eeff ¼ 470 MPa, using the Poisson ratio as measured by Jizba (1991)
increments to be recognized in the cumulative fracture opening (Table 1). Theoretical lower and upper bounds in Poisson ratio of
displacement plot. This result is consistent with the two-stage 0 and 0.5 limit the possible range in “effective” Young's modulus to
geometric fracture growth model 3 (Section 4.1) of fast initial 350 and 500 MPa, respectively.
propagation relative to the rate of aperture growth, followed by a The low “effective” Young's modulus needed to account for the
stage of pronounced aperture growth and slow propagation. The observed fracture opening displacement, two orders of magnitude
two reconstructed bridges, collected away from tips, thus reflect lower than the measured value in core plugs, appears inconsistent
predominantly the second stage of fracture growth in model 3. with a linear elastic fracture opening process. Instead, we attribute
Alternatively, or in addition, the linear slope of the cumulative the differences in lab-measured and “effective” Young's moduli to
fracture opening displacement curve may reflect subcritical frac- fracture strain accommodated by inelastic deformation processes
ture propagation under constant loading stress. Subcritical propa- such as solution-precipitation creep as proposed by Eichhubl
gation is consistent with the slow rate of fracture opening of about et al..(2001), Eichhubl and Aydin (2003), and Eichhubl (2004) for
23 mm/m.y. based on the duration of fracture opening of 48 m.y. for fracture growth in chemically reactive subsurface environments.
a 1.1 mm wide bridged fracture in the Travis Peak Formation in the The timing of fracture opening in the Travis Peak Formation as
SFE-2 well (Becker et al., 2010). The distinction between slow deduced by Becker at al. (2010) based on fluid inclusion micro-
subcritical fracture growth and aperture growth in a near- thermometry, commencing during deep burial and lasting
stationary fracture cannot be made based on the results shown in throughout maximum burial, coincides with silicate diagenetic
Fig. 9. However, slow propagation is again consistent with the reactions and quartz pore cementation in the host sandstone
second growth stage of the two-stage geometric fracture opening (Dutton and Diggs, 1990). The slow fracture opening rates proposed
model outlined above and found most consistent with the fracture by Becker et al. (2010) in the Travis Peak Formation and by Fall et al.
opening displacement data (Figs. 3 and 4). (2012, 2015) in similar burial-diagenetic environments elsewhere,
Cumulative opening displacement curves in Fig. 9 are shown for and synkinematic quartz precipitation resulting in bridge forma-
elastic properties of E ¼ 40,000 MPa and y ¼ 0.18, elastic properties tion are consistent with fracture opening accommodated by slow
that were reported by Jizba (1991) for the Travis Peak Formation in solution-precipitation creep and concurrent silica mass transfer
the SFOT-1 core. These calculations conservatively assume that the from sites of dissolution along grain contacts in the host rock to
measured aperture profiles are in the equatorial plane of circular quartz precipitation sites in the fracture. While inelastic deforma-
tion processes account for the finite fracture opening displacement,
individual fracture opening increments likely represent episodic
Table 1 elastic strain events that alternate with periods of dissolution-
Fracture and material properties used for calculation of cumulative fracture opening
precipitation creep converting elastic strain into permanent, non-
displacement (Fig. 9).
recoverable fracture opening displacement.
Distance from location of maximum aperture [x] (m) 0.0064
Fracture half-length [c] (m) 0.0417a
Young's modulus [E] (MPa) 40,000b 6. Conclusions
470c
 pffiffiffiffiffi
Fracture toughness [KIc] MPa m 1.5
Fracture opening displacement history is recorded in crack-seal
Shape factor [Y] 0.637
Poisson ratio [ʋ] 0.18b cement textures of three opening-mode fractures in the Travis Peak
a
Formation. Trends in average crack-seal cement layer thickness and
c ¼ 0e0.0417 m for propagating fracture.
b
Measured value from Jizba (1991).
in the number of fracture opening increments along fracture length
c
Best-fit Young's modulus to match observed fracture aperture and Poisson ratio and height suggest that fractures grew by initial fast propagation
of 0.18. relative to the rate in aperture growth, followed by a stage of slow
Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44 43

propagation and pronounced aperture growth. Based on the cu- Eichhubl, P., Aydin, A., 2003. Ductile opening-mode fracture by pore growth and
coalescence during combustion alteration of siliceous mudstone. J. Struct. Geol.
mulative opening increment distribution obtained through pal-
25, 121e134.
inspastic reconstruction of two cement bridges that span the entire Eichhubl, P., Aydin, A., Lore, J., 2001. Opening-mode fracture in siliceous mudstone
fracture aperture, we find that the crack-seal record reflects largely at high homologous temperature; effect of surface forces. Geophys. Res. Lett. 28,
the second fracture growth stage of dominant aperture growth and 1299e1302.
Engelder, T., Fischer, M., Gross, M.R., 1993. Geological aspects of fracture mechanics,
subcritical fracture propagation under constant loading stress. geological society of America short course notes. In: Presented on the Occasion
These findings are consistent with earlier fluid inclusion micro- of the Annual Meeting of the Geological Society of America, 1993, Boston,
thermometric studies indicating that fracture cement bridges track Massachusetts, p. 281.
Fall, A., Eichhubl, P., Bodnar, R.J., Laubach, S.E., Davis, J.S., 2015. Natural hydraulic
fracture opening in the Travis Peak Formation spanning up to fracturing of tight-gas sandstone reservoirs, Piceance Basin, Colorado. Geol. Soc.
48 m y. In combination, these results support models of slow Am. Bull. 127, 61e75. http://dx.doi.org/10.1130/B31021.1.
fracture growth in diagenetically reactive environments where Fall, A., Eichhubl, P., Cumella, S.P., Bodnar, R.J., Laubach, S.E., Becker, S.P., 2012.
Testing the basin-centered gas accumulation model using fluid inclusion ob-
fracture opening resulted from repeated episodes of elastic fracture servations: Southern Piceance Basin, Colorado. AAPG Bull. 96, 2297e2318.
opening, with the finite fracture opening accommodated by Gudmundsson, A., 2000. Fracture dimensions, displacements and fluid transport.
solution-precipitation creep. J. Struct. Geol. 22, 1221e1231.
Gudmundsson, A., 2011. Rock Fractures in Geological Processes. Cambridge Uni-
versity Press, Cambridge, United Kingdom.
Hanks, C.L., Parris, T.M., Wallace, W.K., 2006. Fracture paragenesis and micro-
Acknowledgments
thermometry in lisburne group detachment folds: implications for the thermal
and structural evolution of the northeastern Brooks Range, Alaska. AAPG Bull.
This work was supported by the industrial associates of the 90, 1e20.
Helgeson, D.E., Aydin, A., 1991. Characteristics of joint propagation across layer
Fracture Research and Application Consortium at the Bureau of
interfaces in sedimentary rocks. J. Struct. Geol. 13, 897e911.
Economic Geology and by grant DE-FG 02-03ER15430 from Hilgers, C., Urai, J.L., 2002. Microstructural observations on natural syntectonic
Chemical Sciences, Geosciences and Biosciences Division, Office of fibrous veins: implications for the growth process. Tectonophysics 352,
Basic Energy Sciences, Office of Science, U.S. Department of Energy. 257e274.
Holditch, S.A., Robinson, B.M., Whitehead, W.S., 1985. Operating GRI's Mobile
YA acknowledges scholarship support through Saudi Aramco. We Testing and Control Facility. Gas Research Institute Annual Report. S. A. Holditch
thank Christoph Hilgers and Michele Cooke for constructive re- & Associates Inc, College Station, Texas.
views. Publication authorized by the Director, Bureau of Economic Hulin, C.D., 1929. Structural control of ore deposition. Econ. Geol. 24, 15e49.
Jizba, D.L., 1991. Mechanical and Acoustical Properties of Sandstones and Shales.
Geology. PhD Dissertation. Stanford University, Stanford, p. 260.
Klimczak, C., Schultz, R., Parashar, R., Reeves, D., 2010. Cubic law with aperture-
length correlation: implications for network scale fluid flow. Hydrogeol. J. 18,
References 851e862.
Lacazette, A., Engelder, T., 1992. Chapter 12 fluid-driven cyclic propagation of a joint
Atkinson, B.K., Meredith, P.G., 1987. Experimental fracture mechanics data for rocks in the Ithaca siltstone, Appalachian Basin, New York. In: Brian, E., Teng-fong, W.
and minerals. In: Atkinson, B.K. (Ed.), Fracture Mechanics of Rock. Acad. Press, (Eds.), Fault Mechanics and Transport Properties of Rocks. Academic Press, San
London, United Kingdom, pp. 477e525. Diego, pp. 297e323.
Bahat, D., Engelder, T., 1984. Surface-morphology on cross-fold joints of the Appa- Lander, R.H., Laubach, S.E., 2014. Insights into rates of fracture growth and sealing
lachian Plateau, New-York and Pennsylvania. Tectonophysics 104, 299e313. from a model for quartz cementation in fractured sandstones. Geol. Soc. Am.
Becker, S.P., Eichhubl, P., Laubach, S.E., Reed, R.M., Lander, R.H., Bodnar, R.J., 2010. Bull. http://dx.doi.org/10.1130/B31092.1.
A 48 m.y. history of fracture opening, temperature, and fluid pressure: creta- Landry, C., Eichhubl, P., Prodanovic, M., Tokan-Lawal, A., 2014. Matrix-fracture
ceous Travis Peak Formation, East Texas basin. Geol. Soc. Am. Bull. 122, connectivity in Eagle Ford shale. In: Unconventional Resources Technology
1081e1093. Conference (URTEC). Society of Petroleum Engineers, Denver, Colorado,
Bons, P.D., Elburg, M.A., Gomez-Rivas, E., 2012. A review of the formation of tectonic pp. 1e10. http://dx.doi.org/10.15530/urtec-2014-1922708.
veins and their microstructures. J. Struct. Geol. 43, 33e62. Laubach, S.E., 1988. Subsurface fractures and their relationship to stress history in
CER, 1985. Coring and Logging Operations Summary, Ashland Exploration Soil East Texas basin sandstone. Tectonophysics 156, 37e49.
Fertility of Texas No. 1, Nacogdoches County, Texas. Gas Research Institute, Las Laubach, S.E., 1989. Fracture Analysis of the Travis Peak Formation, Western Flank of
Vegas, NV, p. 103. the Sabine Arch, East Texas, Report of Investigations 185. Bureau of Economic
Corbett, K., Friedman, M., Spang, J., 1987. Fracture development and mechanical Geology, The University of Texas at Austin, Austin, TX, pp. 1e55.
stratigraphy of Austin Chalk, Texas. AAPG Bull. 71, 17e28. Laubach, S.E., 2003. Practical approaches to identifying sealed and open fractures.
Cooke, M.L., Underwood, C.A., 2001. Fracture termination and step-over at bedding AAPG Bull. 87, 561e579.
interfaces due to frictional slip and interface opening. J. Struct. Geol. 23, Laubach, S.E., Jackson, M.L.W., 1990. Origin of arches in the northwestern Gulf of
223e238. Mexico basin. Geology 18, 595e598.
Cox, S.F., 1987. Antitaxial crack-seal vein microstructures and their relationship to Laubach, S.E., Lander, R.H., Bonnell, L.M., Olson, J.E., Reed, R.M., 2004a. Opening
displacement paths. J. Struct. Geol. 9, 779e787. Histories of Fractures in Sandstone, vol. 231. Geological Society, London,
Cox, S.F., Etheridge, M.A., 1983. Crack-seal fiber growth mechanisms and their sig- pp. 1e9. Special Publications.
nificance in the development of oriented layer silicate microstructures. Tecto- Laubach, S.E., Reed, R.M., Olson, J.E., Lander, R.H., Bonnell, L.M., 2004b. Coevolution
nophysics 92, 147e170. of crack-seal texture and fracture porosity in sedimentary rocks: cath-
Delaney, P.T., Pollard, D.D., 1981. Deformation of host rocks and flow of magma odoluminescence observations of regional fractures. J. Struct. Geol. 26,
during growth of minette dikes and breccia-bearing intrusions near ship Rock, 967e982.
New Mexico. U. S. Geol. Surv. Prof. Pap. 1202, 1e61. Laubach, S.E., Olson, J.E., Gross, M.R., 2009. Mechanical and fracture stratigraphy.
Dietrich, D., Grant, P.R., 1985. Cathodoluminescence petrography of syntectonic AAPG Bull. 93, 1413e1426.
quartz fibres. J. Struct. Geol. 7, 541e553. Lee, C.H., Farmer, I.W., 1993. Fluid Flow in Discontinuous Rocks. Chapman & Hall,
Duncan, A., Hanks, C., Wallace, W.K., O'Sullivan, P.B., Parris, T.M., 2012. An integrated London.
model of the structural evolution of the central Brooks range foothills, Alaska, National Research Council, 1996. Rock Fractures and Fluid Flow: Contemporary
using structural geometry, fracture distribution, geochronology, and micro- Understanding and Applications. National Academy Press, Washington, D.C.
thermometry. AAPG Bull. 96, 2245e2274. Parris, T.M., Burruss, R.C., O'Sullivan, P.B., 2003. Deformation and the timing of gas
Durney, D., Ramsay, J., 1973. Incremental strains measured by syntectonic crystal generation and migration in the eastern Brooks range foothills, Arctic National
growths. In: De Jong, K.A., Scholten, R. (Eds.), Gravity and Tectonics. Wiley, New Wildlife Refuge, Alaska. AAPG Bull. 87, 1823e1846.
York, pp. 67e96. Pollard, D., Aydin, A., 1988. Progress in understanding jointing over the past Cen-
Dutton, S.P., 1986. Diagenesis and Burial History of the Lower Cretaceous Travis Peak tury. Geol. Soc. Am. Bull. 100, 1181e1204.
Formation, East Texas. PhD Dissertation. The University of Texas at Austin, Pollard, D., Segall, P., 1987. Theoretical displacements and stresses near fractures in
Texas. rock: with applications to faults, joints, veins, dikes, and solution cavities. In:
Dutton, S.P., Diggs, T.N., 1990. History of quartz cementation in the lower cretaceous Atkinson, B.K. (Ed.), Fracture Mechanics of Rock. Academic Press, London,
travis peak formation, East Texas. J. Sediment. Res. 60, 191e202. pp. 277e350.
Dutton, S.P., Land, L.S., 1988. Cementation and burial history of a low-permeability Ramsay, J.G., 1980. The crack-seal mechanism of rock deformation. Nature 284,
quartzarenite, lower Cretaceous Travis Peak Formation, East Texas. Geol. Soc. 135e139.
Am. Bull. 100, 1271e1282. Reed, R.M., Milliken, K.L., 2003. How to overcome imaging problems associated
Eichhubl, P., 2004. Growth of Ductile Opening-mode Fractures in Geomaterials, 231. with carbonate minerals on SEM-based cathodoluminescence systems.
Geological Society, London, pp. 11e24. Special Publications. J. Sediment. Res. 73, 328e332.
44 Y. Alzayer et al. / Journal of Structural Geology 74 (2015) 31e44

Renard, F., Andre ani, M., Boullier, A.-M., Labaume, P., 2005. Crack-seal patterns: re- Taber, S., 1916. The origin of veins of the asbestiform minerals. Proc. Natl. Acad. Sci.
cords of uncorrelated stress release variations in crustal rocks. In: Gapais, D., U. S. A. 2, 659e664.
Brun, J.P., Cobbold, P.R. (Eds.), Deformation Mechanisms, Rheology and Tectonics, Tokan-Lawal, A., Prodanovic, M., Landry, C.J., Eichhubl, P., 2014. Understanding
pp. 67e79. From Minerals to the Lithosphere Geological Society, London, London. tortuosity and permeability variations in naturally fractured reservoirs: Nio-
Rijken, P., Cooke, M.L., 2001. Role of shale thickness on vertical connectivity of brara formation. In: Unconventional Resources Technology Conference (URTEC).
fractures: application of crack-bridging theory to the Austin Chalk, Texas. Tec- Society of Petroleum Engineers, Denver, pp. 1e13. http://dx.doi.org/10.15530/
tonophysics 337, 117e133. urtec-2014-1922870.
Secor, D.T., 1965. Role of fluid pressure in jointing. Am. J. Sci. 263, 633e646. Underwood, C.A., Cooke, M.L., Simo, J.A., Muldoon, M.A., 2003. Stratigraphic con-
Segall, P., 1984. Formation and growth of extensional fracture sets. Geol. Soc. Am. trols on vertical fracture patterns in Silurian dolomite, northeastern Wisconsin.
Bull. 95, 454e462. AAPG Bull. 87, 121e142.
Snow, D.T., 1968. Rock fracture spacings, openings, and porosities. J. Soil Mech. Urai, J.L., Williams, P.F., Van Roermund, H.L.M., 1991. Kinematics of crystal growth in
Found. Div. Am. Soc. Civ. Eng. 94, 73e91. syntectonic fibrous veins. J. Struct. Geol. 13, 823e836.
Soeder, D.J., Chowdiah, P., 1990. Pore geometry in high- and low-permeability Van Noten, K., Muchez, P., Sintubin, M., 2011. Stress-state evolution of the brittle
sandstones, travis peak formation, east Texas. In: Society of Petroleum Engi- upper crust during compressional tectonic inversion as defined by successive
neers Symposium on Gas Technology. Society of Petroleum Engineers, Dallas, quartz vein types (High-Ardenne slate belt, Germany). J. Geol. Soc. 168,
Texas, pp. 239e252. 407e422.
Stowell, J.F.W., Watson, A.P., Hudson, N.F.C., 1999. Geometry and population sys- Vermilye, J.M., Scholz, C.H., 1995. Relation between vein length and aperture.
tematics of a quartz vein set, Holy Island, Anglesey, North Wales. Geol. Soc. J. Struct. Geol. 17, 423e434.
Spec. Publ. 155, 17e33.

You might also like