You are on page 1of 7

special topic

first break volume 30, February 2012

Unconventionals & CCS

Enhancing shale gas reservoir characterization


using hydraulic fracture microseismic data
Shawn Maxwell1* and Mark Norton2 present a case study showing how microseismic data can
be a useful part of the mix of geotechnical information sources that, when combined, provide
valuable information about shale gas reservoirs.

hale gas has more than doubled the natural gas reserves
and resources in North America according to estimates
by IHS CERA, adding 1200 tcf of shale gas in the US
and 500 tcf in Canada. Knowledge gained in North
America during the last few decades is now being used to
benefit the increased interest in shale gas around the world. In
Europe, shale gas is being explored in several countries, including Poland, Austria, England, Germany, Hungary, and Sweden.
Mudstones and shales are generally considered to be
the source rock for most hydrocarbon reservoirs. Many still
contain large volumes of gas but, because of their low permeability, they are rarely economic to produce using conventional
drilling and well completion technologies. Natural fractures
can improve permeability, but in gas shales they usually do not
provide adequate pathways for sufficient flow of hydrocarbons
into a well. Most gas shales require long horizontal wells, positioned accurately to maximize exposure to the most productive
zones, and reservoir stimulation particularly hydraulic
fracturing to increase the rate at which hydrocarbons can
be produced from the rock formation. Fluids are pumped
into multiple isolated stages along the well at high pressure
causing the formation to crack. The composition of the fluid
used varies depending on criteria such as formation mineralogy
and permeability. Solid proppant materials, typically sieved
round sand grains or man-made ceramic spheres, are added
to the fracture fluid. These solids hold the fractures open after
injection stops, so the propped hydraulic fracture becomes a
conductive path through which fluids can flow from the rock
formation to the well.
Experience in North America has shown that gas production and recovery rates vary considerably, not only from one
well to another but also between fracturing stages along a
single well. The material properties of shales exhibit high levels
of vertical and lateral heterogeneity resulting from depositional and post-depositional processes that include diagenesis,
interaction with organisms, thermally activated geochemistry,
and movement of mineral-laden fluids. In addition, existing
fractures and local stress fields have a major impact on the
effectiveness of hydraulic fracturing processes. Generating

enough surface area for gas to flow requires a thorough


understanding of the mineralogy and stress regimes in the rock.

Integrating information
Collecting and integrating different sources of knowledge about
shale reservoirs is the key to gaining competitive advantage by
avoiding non-productive acreage, selectively drilling the sweet
spots and effectively stimulating the wellbore. Surface geological studies and high quality 3D seismic data provide valuable
field-wide data before development drilling. Important information acquired during and/or after drilling includes resistivity,
gamma ray, neutron density measurements, as well as sonic
logs, rock cores, downhole spectroscopy measurements, and
production data. Integrating these sources of information
helps in the understanding of reservoir heterogeneity in terms
of mineralogy and organic content, as does learning how stress
varies horizontally and vertically near a well bore and across
the reservoir. Bringing all the information together requires
advanced software to enable model-centric analysis in a shared
subsurface representation. Integrated seismic-to-simulation
software allows companies to incorporate all the data elements
in one subsurface model.

Microseismic data
Hydraulic fracturing results in microseismic events that can
be recorded at the surface or from monitor wells to locate
and help characterize the fractures. Microseismic imaging can
provide valuable information about a reservoir helps to design
the placement and completion of wells. In some cases, the
microseismic activity induced by the stimulation is consistent
with relatively simple, planar fractures while, in other cases,
interaction with pre-existing fracture networks results in
complex fracture networks growing in multiple orientations.
Behaviour ranging from simple to complex fractures has been
observed in multi-stage stimulations within single horizontal
wells (Cipolla et al., 2010). The variability in fracture complexity depends on the geometry of the pre-existing network as well
as the in-situ stress state. Fracture complexity is thought to be
enhanced in the presence of pre-existing fractures at an angle

Schlumberger.
Progress Energy.
*
Corresponding author, E-mail: smaxwell@slb.com
2

2012 EAGE www.firstbreak.org

95

special topic

first break volume 30, February 2012

Unconventionals & CCS


to the maximum stress direction, with low stress anisotropy to
allow fractures in multiple orientations to open (Sayers and Le
Calvez, 2010).
Adding microseismic hydraulic fracture imaging to other
sources of geologic information about fractures and stresses
in the reservoir is advancing our understanding of hydraulic
fracture complexity in shale gas reservoirs. Integration of
microseismic hydraulic fracture responses with seismic reflection reservoir characterization is beginning to show interesting correlations (Rich and Ammerman, 2010). This article
describes an investigation into potential reasons for variations
in microseismic response observed in measurements made during hydraulic fracture stimulations of three horizontal wells.

Case Study: Montney Formation


Norton et al. (2010) described the integration of microseismic data with reservoir conditions identified from seismic
reflection data for the Upper Montney formation in NE
British Columbia, Canada. The investigation has been
extended by examining other aspects of the microseismic
response recorded during multi-stage hydraulic fracturing
of three horizontal wells, and investigating aspects of the
stimulation treatment parameters that relate to reservoir
conditions and heterogeneity.
The Montney is a fine-grained siltstone with matrix porosities between 3% and 6% and permeabilities in the 0.001 to
0.05 mD range. The target interval is approximately 1750
m true vertical depth (TVD), and each well was stimulated
with between six to eight stages of N2 energized slick water
pumped into single perforation clusters over the length of the
approximately 1500 m horizontal well sections at rates of
approximately 10 m3/min. Microseismic monitoring used eight
three-component seismic sensors, spaced at 30 m and placed
in a single vertical observation well located close to one end of
each of the horizontal wells being treated. Offset to the furthest
stages was approximately 1600 m.
Figure 1 shows a map view of the microseismicity recorded
for all the stages. The distribution of the microseismicity follows a zone trending to the NW and SE, with very few events
occurring to the NE. Norton et al. (2010) described how surface
seismic amplitude-versus-offset (AVO) data was used to invert
for variations in Poissons ratio (PR). This exhibits lateral variations between 0.1 and 0.3 in the study region, indicative of
changes in material properties. The seismic data was also run
through an enhanced edge detection (ant-tracking) algorithm
to identify lateral discontinuities, which indicated a number
of NE-SW trending fractures. Figure 2 shows microseismic
events combined with PR and ant-tracking, indicating that
stress variations associated with both the material properties
and pre-existing faults affect how the reservoir responded to
hydraulic fracture stimulations. However, detailed study of the
location and magnitude of the microseismic events prompts
questions regarding the fracture response, including:

96

Figure 1 Map view of microseismic events, scaled by magnitude, for all fracturing stages.

Why are the microseismic events large along Well-C?


Why are there no microseismic events towards the NE?
Why do the microseismic events tend to go towards
the SW?
Why is the geometry of the microseismic events different
near Well-A?

Analysis of the microseismic data


Different distances between the open perforations and the
monitoring well are a cause of variation in the quality and
quantity of the microseismic data between treatment stages.
To lessen this impact, analysis was restricted to only data
from fracturing stages closest to the monitoring well.
Analysis of the magnitude of microseismic events versus
distance to the monitoring array showed how smaller
magnitude microseismic events are recorded close to the
observation well. The analysis was used to define a lower
magnitude limit in this case a magnitude limit of -2.2
above which events would be recorded regardless of
where they were located within a distance of 1000 m, thus
removing the detection-distance bias. This subset is shown
in Figure 3.

www.firstbreak.org 2012 EAGE

first break volume 30, February 2012

special topic

Unconventionals & CCS


Microseismic events recorded during the Well-C fracture
treatments were larger in magnitude than those recorded while
treating the other two wells, suggesting interaction with preexisting faults. There is a distinct lack of seismic response NE
of wells B and C within the detection limits of 1000 m around
the observation well, even though events above magnitude -2.2
would have been detected if they had occurred.
Seismic moment is a robust measure of the source strength
of the microseismic event and provides insight into the
reservoir deformation (Maxwell et al., 2006). Figure 4 shows

contours of the logarithm of seismic moment density. The


seismic moment or deformation density along wells B and C is
relatively high, which is consistent with fault activation. While
the number of events is also high around Well-A, the deformation level is smaller than near the other wells. The deformation
SW of Well-C is low, indicating relatively few small magnitude
events in this region. The few events that did occur happened
late in the associated stages.
The frequency-magnitude relationship was also investigated. The number of events typically follows an exponential

Figure 2 Microseismic events combined with ant-tracking results and Poissons


Ratio computed from surface seismic data.

Figure 4 Contours of logarithm of seismic moment density.

Figure 3 Microseismic events closest to monitoring well with magnitude >-2.2.

2012 EAGE www.firstbreak.org

Figure 5 Contours of b-value slope of microseismic event frequency-magnitude relationship.

97

special topic

first break volume 30, February 2012

Unconventionals & CCS

Figure 6 Composite focal mechanisms showing observed S/P amplitude ratios and modelled radiation patterns near Well-A (left) and near wells B and C (right).

Gutenberg-Richter relationship similar to earthquakes: log


N = a bM, where the slope or b-value is typically in the
range 12. A maximum likelihood estimate for the b-value
is given by: b = 1/2.3(Mav Mmin), where Mav is the average
magnitude about a cut-off level of Mmin. Average magnitudes
were computed for regions with at least 20 events, shown in
Figure 5. b-values can be used to distinguish fault activation
from microseismic response associated with a hydraulic fracture. Fault activation results in a b-value of approximately 1,
but higher values are observed for hydraulic fractures (Maxwell
et al., 2009). In this case, low b-values (near 1) are found
along wells B and C, which is consistent with the deformation density, indicating that the microseismic response in this
region is related to fault activation. The other major feature
is near Well-A, where the b-value is higher, consistent with a
conventional hydraulic fracture.
Finally, composite focal mechanisms were computed for the
two clusters defined by the different b-values. Composite focal
mechanism plots fit the observed seismic radiation patterns
from groups of events to theoretical mechanisms to estimate
the average failure mechanism (e.g., Rutledge et al., 2004).
Often the S- to P-wave amplitude ratio can be modelled with
the theoretical radiation pattern of a single shear deformation
in the direction of lineations in the microseismic locations (e.g.,
Maxwell, 2011). Interaction of a tensile hydraulic fracture with
pre-existing fractures subparallel to the maximum principal
stress direction would account for the shear mode of microseismic deformation. Figure 6 shows focal mechanisms for the
S/P amplitude ratios of the two clusters, along with modelled
radiation patterns. The cluster to the SW of Well-A is modelled
by a NE-SW strike-slip mechanism in the expected orientation
orthogonal to the regional minimum stress direction. While
some data points fall off the curve, there is a reasonable fit
to the majority of the events. The cluster along wells B and
C is modelled by a NW-SE dip-slip mechanism. The general
consistency of this fracture mechanism with the regional faults
depicted by the ant-tracking algorithm further supports the
notion of fault activation in this area.

98

In summary, microseismic magnitudes, seismic deformation


density, b-values, and composite focal mechanisms indicate
that the large events occurring along wells B and C are consistent with activation of a pre-existing fault or fracture network
striking NW-SE. Any comparable microseismic events to the
NE would have been detected. Apparent asymmetry towards
the SW is not related to detectability. The events near Well-A
clustered in a NE-SW lineation are consistent with a simple,
planar hydraulic fracture oriented orthogonal to the pervasive
minimum stress direction, based on magnitudes, seismic deformation density, b-values, and composite focal mechanisms.

Treatment data
All fracturing treatments were placed at 10 m3/min
0.6 m3/min, with the exception of stage 3 in Well-C when a
high near wellbore pressure loss resulted in a lower average
injection rate of 7 m3/min. Due to the close proximity of the
wellbores, similar treatment responses would be expected in
wells A, B, and C. However, surface-treating pressure data
indicated that not only did the wells respond differently
from wellbore to wellbore, but that there were also notably
different responses for different stages in the same wellbore.
Stage 1 of Well-B and stage 3 of Well-C were particularly anomalous. The highest breakdown pressures of the
campaign were observed in Well-C, stages 3 and 5. Well-C
is located in the area with higher PR, which would lead to
higher overall stresses and near-wellbore stress concentrations,
and hence would be expected to exhibit higher breakdown
pressures. Closure pressures were measured from mini-frac
injection/fall-off tests for the first stage in each well and were
27.4 kPa/m, 22.5 kPa/m, and 23.5 kPa/m for wells A, B, and
C respectively. Pressure decline data was not available for
subsequent stages due to operational constraints.
Instantaneous shut-in (ISIP) gradients at the end of each
propped fracture treatment plotted at the perforation locations
(Figure 7) illustrate the likely areal variation in stress regime.
The relatively high ISIPs coincide approximately with the
higher PR regions, possibly related to high stress. Although the

www.firstbreak.org 2012 EAGE

first break volume 30, February 2012

special topic

Unconventionals & CCS

Figure 8 Daily production rates for the three wells. Note that production on
Well-A drops when production on Well-B starts (circled time period).

Figure 7 Colour-coded instantaneous shut-in (ISIP) gradients (kPA/m) at the


fracturing locations superimposed on the microseismic and Poissons Ratio data.

ISIP variations are modest, the fracture treatments in Well-C


exhibit ISIPs that are about 0.5 MPa higher than wells A and
B. The location of the higher ISIPs correlates with the high concentration of larger magnitude microseismic events, supporting
the interpretation of fault activation during the Well-C fracture
treatments. Detailed analysis of ISIP plots for the three stages
closest to the monitoring well indicated that the behaviour of
Well-C is different than wells A and B: exhibiting a higher ISIP,
less pronounced water hammer pressure oscillations, and a
steeper initial pressure decline. The differences in fracture treatment behaviour and ISIP signatures point toward a different
stress regime in the vicinity of Well-C.

Production data
Daily production data for the three wells (Figure 8) show
that when production on Well-B starts, production on Well-A
drops, suggesting interference between the two wells. The
locations of microseismic events recorded while treating wells
A and B overlap, which is consistent with the possibility of well
interference.

Conclusions
Why are the microseismic events large along Well-C?
The seismic deformation, b-values, and focal mechanisms
all point toward interaction of the hydraulic fractures with
a NW-SE striking fault system. While the fault interaction
is predominantly associated with fracturing wells B and C,
a small indication is found for the NE extent of the Well-A

2012 EAGE www.firstbreak.org

microseismic response. Ant-tracking does not provide a strong


indication of a pre-existing fault in this region, although
other faults are apparent elsewhere with a similar strike. It
should be noted that ant-tracking searches for faults with an
apparent throw, implicitly assuming strain release associated
with earlier tectonic movements. Faults with small amounts
of throw and potentially subseismic resolution are more likely
to be tectonically loaded and prone to releasing seismic energy
due to effective stress changes and lubrication associated with
hydraulic fracturing. Therefore, the microseismically activated
fault systems may be difficult to resolve with seismic reflection
data. Ant-tracking results just below the reservoir provide
evidence of a deeper fault network (Figures 9 and 10), and
the observed microseismic deformation seems to be associated
with the activation of the vertical extension of this fault.
Why are there no microseismic events towards the NE?
If microseismic responses with magnitudes -2.2 or greater had
occurred further to the NE, it would have been detected. While
mapping of PR suggests significant reservoir heterogeneity, it
is important to note that the entire reservoir has relatively low
PR. It is therefore unlikely that the NE region with higher PR
would have a significantly different fracturing mode than the
rest of the reservoir. Deformation could have occurred without
a large seismic event, but even very small magnitude events
would have been detected, especially for the closest stages. It
can be postulated that the hydraulic fracture interacted with
the fault system described above, which acted as a barrier to
the hydraulic fracture extending further to the NE. As long as
the permeable open fault allows fluid penetration within the
fault network, it is easy to imagine that it would be unlikely
to create a new hydraulic fracture. Well-C, however, does have
an indication of microseismic response to the SW of the well,
occurring later in each stage, which suggests that eventually the
hydraulic fracture does start to grow out of the fault network.
As described in the next section, the microseismic activity

99

special topic

first break volume 30, February 2012

Unconventionals & CCS


hydraulic fracture in Well-B propagated into the network created during stimulation of Well-A. The SW progression from
each well is towards the direction of lower PR. As discussed
in Norton et al. (2010) the PR can be inversely correlated to
quartz-clay content and could be related to a change in the
mode of fracturing. Material property changes will also result
in associated elastic stress changes through the reservoir (see,
for example, Cipolla et al, 2010). Furthermore, the faulting
shown by the ant- tracking also implies elastic and inelastic
stress variations, which in turn will also change the material
properties. In addition, a constrained, horizontal stress effect
is associated with transverse strain resulting from the vertical
lithostatic load, which is given by mH = mh = (i/(1-i)) * mzz +
mtectonic and would result in lower stresses in lower PR (i)
material. The stresses in the SW direction would be lower due
to this effect, resulting in more hydraulic fracture growth in
that direction. The observed ISIP gradients are also consistent
with lower stresses in the lower PR regions. While the specific
impact of these different factors is beyond the scope of this
article, there is a clear correlation between the asymmetric SW
growth and the low PR region.

Figure 9 Microseismic events and ant-tracking image.

then moves in a SW direction similar to the rest of the wells.


Therefore, the lack of microseismic event towards the NE is
attributed to fracture asymmetry associated with the fault
system near wells B and C.
Why do the microseismic events tend to go towards the SW?
The microseismic response from each well grows preferentially
in a SW direction. For example, some events to the NE of Well-A
reach as far as Well-B, although the majority of microseismic
activity is in the SW direction. The furthest NE microseismic
events could be related to stress-induced deformation of the
fault. As previously discussed, Well-B microseismic response
appears to overlap with the Well-A activity, suggesting that the

Why is the geometry of the microseismic events


different near Well-A?
The microseismic events associated with stimulation of Well-A
are, as expected, distributed in the maximum stress direction.
Therefore, Well-A appears to be anomalous simply due to the
lack of fault activation seen around wells B and C.

Summary
Hydraulic fracturing of the three wells in this study resulted in
significant differences in the microseismic response. Interaction
of the hydraulic fractures with a pre-existing fault resulted in
relatively large magnitude microseismic response along the NE
edge. In some cases, relatively simple, planar hydraulic fractures were created in the expected NE-SW direction, although
the fractures tend to preferentially grow towards the SW in the
direction of lower PR where material properties and stresses
are different. Understanding the hydraulic fracture response in

Figure 10 Ant-tracking volume and corresponding


seismic section showing two interpreted faults (red
dashed lines). Right side is zoomed in region showing relationship with microseismic.

100

www.firstbreak.org 2012 EAGE

special topic

first break volume 30, February 2012

Unconventionals & CCS


this region has enabled better well placement for more optimal
hydraulic fracture geometry and improved production rates, in
addition to better completion designs with closer perforation
clusters to increase reservoir contact in future wells.

Hydraulic Fracture Complexity. SPE Annual Technical Conference and


Exhibition, San Antonio, Texas. SPE 102801.
Maxwell, S.C., Jones, M., Parker, R., Miong, S. Leaney, S., Dorval, D.
DAmico, D., Logel, J., Anderson, E. and Hammermaster, K. [2009]
Fault Activation during Hydraulic Fracturing. 79th SEG Annual

Acknowledgements

Meeting. Expanded Abstracts, 28, 1552.

This article is based largely on SPE 140449 Enhanced Reservoir


Characterization Using Hydraulic Fracture Microseismicity
presented at the SPE Hydraulic Fracturing Technology
Conference and Exhibition, The Woodlands, Texas, USA,
2426 January 2011.

Maxwell, S.C. 2011. What Microseismic does and does not indicate about
hydraulic fractures. 73rd EAGE Conference & Exhibition, Extended
Abstracts.
Norton, M., Hovdebo, W., Cho, D., Jones, M. and Maxwell, S. [2010]
Surface Seismic to Microseismic: An Integrated Case Study from
Exploration to Completion in the Montney Shale, NE British Columbia,

References

Canada. 80th SEG Annual Meeting. Expanded Abstracts.

Cipolla, C.L., Williams, M.J., Weng, X, Mack, M. and Maxwell, S. [2010]

Rich, J.P. and Ammerman, M. [2010] Unconventional Geophysics for

Hydraulic Fracture Monitoring to Reservoir Simulation: Maximizing

Unconventional Plays. Unconventional Gas Conference, Pittsburgh,


Pennsylvania, USA. SPE 131779.

Value. SPE Annual Technical Conference and Exhibition, Florence,

Rutledge, J.T., Phillips, W.S and Mayerhofer, M.J. [2004] Faulting Induced

Italy. SPE 133877.


Downie, R.C., Kronenberger, E., and Maxwell, S.C. [2010] Using

by Forced Fluid Injection and Fluid Flow Forced by Faulting: An

Microseismic Source Parameters to Evaluate the Influence of Faults

Interpretation of Hydraulic-Fracture Microseismicity, Carthage Cotton

on Fracture Treatment A Geophysical Approach to Interpretation.

Valley Gas Field, Texas. Bulletin of Seismological Society of America,


945, 1817.

SPE Annual Technical Conference and Exhibition, Florence, Italy. SPE

Sayers, C. and Le Calvez, J. [2010] Characterization of Microseismic Data

134772.
Maxwell, S.C., Waltman, C.K., Warpinski, N.R., Mayerhofer, M.J. and
Boroumand, N. [2006] Imaging Seismic Deformation Associated with

in Gas Shales Using the Radius of Gyration Tensor. 80th SEG Annual
Meeting. Expanded Abstracts, 20802084.

EDUCATION DAYS LONDON 2012


19-23 March 2012 - Hilton London Gatwick, UK

Sign up NOW for the

Programme

following 1 & 2 day courses

19 March

Advanced Marine Seismic


Acquisition Techniques
Mr Gordon Brown

19 March

Well Test Analysis


Dr Shiyi Zheng

on current geoscience topics,


taught by acknowledged
experts (both industry
professionals and academics).

20 & 21 March Pragmatic Sequence


Stratigraphy
Dr Gary Hampson
21 March

Seismic Surveillance for


Reservoir Delivery (EET 6)
Mr Olav Inge Barkved

22 & 23 March Seismic Imaging: A Review of


the Techniques, their Principles,
Merits and Limitations
Mr Etienne Robein

22 & 23 March Understanding Subsurface


Pressure and Pressure
Prediction
Dr Phill Clegg

Venue
Hilton London Gatwick
South Terminal
Crawley RH6 0LL
United Kingdom

Daily time schedule


Registration
Courses

08:15-08:45 hrs
09:00-17:00 hrs

Visit the EAGE website for detailed course descriptions and


registration information.
EAGE and PESGB members register for the discounted course fee. Course
materials / book (EET 6) are included in the fee, as well as coffee breaks and
lunch. Registration for non-members includes the EAGE membership for
2012. Special fee available for attending multiple courses.

2012 EAGE www.firstbreak.org

www.learninggeoscience.org

101

You might also like