You are on page 1of 24

SPE 152165

Engineering Guide to the Application of Microseismic Interpretations


C. Cipolla, S. Maxwell, and M. Mack, Schlumberger

Copyright 2012, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Hydraulic Fracturing Technology Conference held in The Woodlands, Texas, USA, 6–8 February 2012.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Microseismic monitoring (MSM) of hydraulic fracture treatments is routine in North America and has added significantly to
our understanding of fracture growth. The interpretation of microseismic images is advancing steadily, extracting more
information from event patterns, temporal evolution, and acoustic waveforms. The increasing amount of information from
MSM provides significant opportunities to improve stimulation designs, completion strategies, and field development.
However, the applications of microseismic interpretations are many times ill-defined, overlooked, or not applied properly.
Numerous applications of microseismic measurements have been documented in technical publications, typically in the form
of case histories focused on specific applications. The industry has lacked a compilation and comprehensive discussion of
microseismic applications. This paper presents a practical guide for the engineering application of microseismic
interpretations, documenting reliable application workflows while highlighting the consequences of misapplication of
microseismic interpretations.
The application of MSM starts with a reliable interpretation of fracture geometry and complexity, but the real value is in
the application of the interpretation. This paper divides microseismic applications into three categories, real-time, completion
strategies & stimulation design, and field development. The MSM interpretation requirements for each category are
documented and a comprehensive guide to properly applying these interpretations is presented. Applications issues such as
determining the “effective” fracture surface area, the relationship between microseismic behavior and well performance, and
fracture model calibration are addressed.
There is a growing interest in advanced processing such as moment tensor inversion (MTI) and b-values to determine
focal mechanisms, source parameters, and failure mechanisms associated with the microseismic events. However, the
engineering application of these interpretations is not well understood. This paper includes a discussion of the applications of
advanced processing results, emphasizing how the limitations and uncertainties of the processing affect the subsequent
applications.

Introduction
Microseismic mapping of hydraulic fracture treatments is now commonplace in North America and has played an important
role in the advancement of stimulation and completion techniques, especially for unconventional reservoirs (King et al.,
2008; King, 2010; Waters et al., 2009a; Cipolla et al., 2010, 2011a). The initial focus for microseismic technology was
improving the acquisition of the seismic waveforms, more accurate event locations (processing), and visualizing the images
with the relevant geological and geophysical context (Fig. 1). However, current efforts focus on more value-added
applications of microseismic data, including real-time treatment control, integration with geology, geophysics, geomechanics,
and fracture modeling, and utilizing microseismic data in reservoir engineering workflows (Fig. 1). These value-added
applications of microseismic data are the primary focus of this paper, as they have the potential to add significant insights
into hydraulic fracture “effectiveness” and completion “efficiency” that could result in considerable improvements in well
performance.
The application of microseismic measurements requires a reliable and detailed interpretation of the geophysical images
and data. Cipolla et al. [2011b] present a guide for interpreting microseismic measurements, which serves as a starting point
for this paper. Since the interpretation and application of microseismic measurements are closely linked, the interpretation
guidelines from this previous work are reviewed to provide the necessary background for the subsequent applications.
2 SPE 152165

Fig. 1 – Initial focus of microseismic technology was improving acquisition, more reliable event locations, visualization of
images with geological and geophysical context, and interpretation of the images. Current efforts now include more focus on
value-added applications, including real-time placement, integration with geology, geophysics, geomechanics, and fracture
modeling, and using microseismic images in reservoir engineering workflows.

Interpretation Guidelines (adapted from SPE 144067, Cipolla et al., 2011b)


The typical interpretation of microseismic event patterns consists of six primary geometrical observations and the estimation
of stimulated volume (SV):

1. Fracture Length
2. Fracture Height
3. Fracture Azimuth
4. Fracture Complexity (i.e. – network or planar fractures)
5. Fracture Location with respect to the perforations, frac port, or other exit point from the wellbore.
6. Anomalous behavior (i.e. – fault activation, asymmetry, etc.)
7. Stimulated Volume - SV (SRV, ESV, etc.)

Basic Microseismic Location Interpretation


Reliably interpreting microseismic locations requires an understanding of the inherent uncertainty in the measurements,
the mechanisms that generate microseisms, the reliability of input data, and the accuracy of the processing algorithms. The
interpretation of microseismic data begins with the fundamental source characteristics including event locations and times,
magnitude and P/S amplitude ratios along with QC attributes consisting of at least signal-to-noise-ratio (SNR) and location
error ellipsoids. It is essential to account for observation well distance and location bias because microseismic images are
often misinterpreted without these corrections, potentially leading to erroneous applications. The final interpretation requires
integration with the geological setting, visualization of the events in space and time, and simple fracture mechanics that link
the microseismic image to the treatment data. The basic interpretation workflow and guidelines are summarized below.

1. Uncertainty. The first step in the interpretation of microseismic measurements is to evaluate the uncertainty in the
event locations. The location uncertainty for most events will differ in different directions (depth, distance, and
azimuth). SNR can be used as an indicator of event waveform quality, with higher SNR typically associated with
more accurate event locations.
a. Filtering. The microseismic events should be filtered using various SNR cutoffs and the resultant event
patterns and error ellipsoids compared. Significant changes in event patterns and/or increases in location
uncertainty could indicate that events with higher location uncertainty are introducing errors into the
interpretation. Individual data sets include distributions of SNR and uncertainty: ranging from events with
low to high confidence. The data should be filtered to minimize events with high location uncertainty prior
to the interpretation, while including sufficient events to characterize the geometry without imposing
location biases. The appropriate SNR and error cutoff will depend on the overall quality and number of
events.
b. Velocity Model. Uncertainty in the velocity model is often overlooked when interpreting microseismic event
patterns. In some cases, errors in the velocity model may result in significant interpretation errors. The
accuracy of the velocity model can be verified by the accuracy of perforation shots if a clear s-wave is
observed. The geophysical work-product may include velocity model uncertainty in the calculation of event
location uncertainty (e.g. error ellipsoids). Quality control during the geophysical processing is a critical
SPE 152165 3

step prior to interpretation to ensure that the microseismic locations are correct and that the uncertainty and
confidence in the results are communicated to the interpreter to avoid over interpretation of the
microseismic data.
2. Observation Well Bias. The location of the observation well with respect to the treatment well and the location of
microseismic activity can introduce errors into the interpretation. Small events can only be detected if they are close
to the sensor array and at some distance even the largest events cannot be detected. The strength of the microseismic
events can vary considerably, resulting in an apparently greater event density close to the sensor array. Increased
background noise levels can also impact detectability. Magnitude-distance plots provide QC of the observation well
distance bias. Directional biases can also occur in cases where the events have similar failure mechanisms, such that
the source radiation results in nodal planes where either the p- or s-waves are not observed. Microseismic detection
algorithms that rely on indentifying both wave types will then result in fewer events in these nodal directions. P/S
amplitude ratio plots provide QC of potential directional biases. Understanding the detection limit is important when
interpreting fracture geometry and comparing stimulation treatments. In cases where the interpretation indicates that
the entire fracture geometry was not detected due to biases, the geometry must be assumed to be symmetrical.
3. Geologic Environment. The interpretation of microseismic event patterns requires a basic knowledge of the
geologic environment in which the hydraulic fracture propagated. Event patterns can vary significantly depending on
the reservoir fluids, stress regime, existence of natural fractures, matrix permeability, and rock properties. An
understanding of the geologic environment and the basic mechanisms that generate microseisms will help constrain
the interpretation. Note that the interpretation should identify anomalous microseismic events that are likely not
associated with fracture propagation, such as stress-induced fault activation.
4. Visualization and Integration. The bulk of routine microseismic interpretation focuses on integration of the
microseismic data with geology, seismic, and treatment data and visualization of the event patterns. The
visualizations can be two dimensional or three dimensional and include both geologic and seismic information and
include Temporal and Spatial Evaluation, Stimulated Volume (SV), Spatial and Temporal Event Histograms and
Anomaly Identification. These interpretation tools are described in detail by Cipolla et al. [2011b] and will be
illustrated later in the text using examples and case histories.
5. Mass Balance and Simple Fracture Mechanics. Integrating the volume of fluid, proppant pumped and the net
pressure at the end of the treatment with the microseismic event pattern can help constrain the interpretation. Simple
fracture mechanics equations can be used to calculate the average fracture width, total fracture area, and total fracture
length from the microseismic fracture height, estimated net pressure and fluid efficiency. In the case of complex
fracture networks, total fracture length can be combined with stimulated volume calculations such as ESV or SRV to
estimate network fracture spacing.

With a reliable interpretation of the microseismic image providing fracture geometry, azimuth, location, and complexity
along with the identification of anomalous behavior (if present), the microseismic application workflows can be executed.
The mechanism of the microseismic data in relation to the hydraulic fracture will be discussed later, but basic interpretation
makes the inherent assumption that the hydraulic fracture(s) is proximal to the microseismic events. Although there are very
limited direct observations, extensive work was performed at the M-Site (MWX project) to validate the application of
microseismic mapping to image hydraulic fracture geometry (Warpinski et al., 1998), while cored hydraulic fractures at the
Mounds drill cuttings injection experiments confirmed microseismic fracture geometry (Moschovidis et al., 2000). In
addition, observations of offset wells being contacted (i.e. – “killed”) by fracturing fluid during an offset well treatment also
support microseismic interpretations of fracture geometry (Fisher et al., 2002). Note that the interpretation should identify
anomalous microseismic events that are probably not associated with fracture propagation (Cipolla et al, 2011b).

Advanced Processing, Stimulated Volume, and Fracture Area


Beyond the microseismic locations, additional parameters can be used to characterize the deformation associated with the
microseismic locations. It is important to understand the value and limitation of advanced seismic processing and
interpretations to properly apply the results. The following sections discuss the issues associated with applying advanced
microseismic processing and stimulated volume interpretations.

Advanced Processing and Fracture Area


In addition to event location and magnitude, there are ongoing efforts to extract additional information from the seismic
waveforms. Advanced processing may provide important insights into the details of the failure mechanisms that can help
define natural fracture characteristics and constrain hydraulic fracture models. However, these advanced processing
techniques are relatively new and untested, and the limitations are not well understood by engineers when applying the
results. It is important to understand the relationship between basic and advanced processing results, actual fracture
geometry, and well productivity.
Microseismic activity rates. An obvious metric of a microseismic data set is the number of events. However, the number
of events should not be misinterpreted as the true activity rate. A better measurement of activity is the total source strength
or cumulative seismic moment (moment magnitude is the logarithm of this value), as discussed by Maxwell et al., (2006). In
4 SPE 152165

terms of the geomechanical deformation associated with each microseism, a relatively large magnitude event can represent
the same total rock strain as numerous smaller magnitude events. Small variations in detection distance and minimum SNR
can result in big differences in the number of events. Cumulative moment is less sensitive to these variations, because most
of the moment is associated with the bigger events. Applications comparing microseismic data often use activity rates as one
of the comparisons; the ‘Comparative Interpretation’ section of this paper reviews the proper workflow to enable such a
comparison.
Frequency-magnitude relationships. Analysis of the relationship between frequency (or number: N) and magnitude (M)
generally follows a Richter-Gutenberg power law relationship of the form: log N = a - b M. The so called “b-value” or slope
of the relationship is a good indicator of fault activation. For most fracs the b-value is approximately 2 (meaning there are
100 times more events for a decrease of one magnitude unit) while faults result in a b-value of approximately 1 (Maxwell et
al., 2009). Geo-hazard assessment using this observation is described later in the paper. Normalizing the frequency-
magnitude for uniform detection is also an important aspect of applications requiring a comparison of activity rates as
described later.
Microseismic source mechanisms. The radiation pattern or directionality of p- and s-waves amplitudes can be used to
define the failure mechanisms: the orientation of the failure plane and type of failure (e.g. shear or tensile). The simplest form
is to assume shear failure and use the corresponding radiation characteristics to define the orientation of the failure plane.
This can be graphically represented by a so-called “beach-ball diagram” akin to earthquake seismology. Alternatively, a
technique called “moment tensor inversion” can be used to estimate the fracture plane and the type of failure simultaneously.
To determine the mechanism for a single event, the radiation pattern must be determined in different directions either through
surface or multiple downhole monitoring. Accuracy or confidence in the resulting moment tensor is critical, with recent
research highlighting potential significant errors in both the failure plane orientation and the type of failure (e.g., Kim, 2011).
Moment tensor can be plotted using standard seismology plots of radiation patterns or source plots, although new
visualization is becoming available for engineering applications with more intuitive depiction of fracture planes and
displacement vectors (Leaney and Chapman, 2011). Moment tensor is potentially of interest, enabling shear failure to be
distinguished from tensile failure as well as identifying fracture opening or closing. However, the geomechanical
interpretation of moment tensor data remains uncertain at this point in time: since the seismic energy associated with either
fracture opening or closing will be relatively small in comparison to shearing because most rocks tend to be relatively weak
in tension. Maxwell and Cipolla (2011c) argue that the majority of microseismic deformation tends to be in shear, and that a
geomechanical based fracture model is needed to reconcile the paradox between shear microseismic deformation and tensile
fracture opening and closing. This will be further discussed in the “Misinterpretation” section. A final category of mechanism
analysis is more appropriate for single downhole monitoring, where groups of events are used to define a common failure
mechanism. One form of this is plotting p- to s- amplitude ratios as a function of azimuth and fitting a theoretical radiation
pattern (e.g. Rutledge et al.). The main applications of mechanisms are to identify fault activation or examine potential
fracture complexity from interaction with pre-existing fractures.
Microseismic source attributes. Source parameters define the geomechanical strain occurring at the microseismic source.
The most common is the seismic moment (defined as shear modulus*area*displacement) which is used to compute the
moment magnitude (Mw = 2/3 log Mo -6) discussed above. The moment is potentially of interest since the product of area
and displacement defines a fracture volume associated with the microseismic event. To explore this microseismic fracture
volume, we assume that the displacement is a tensile fracture opening and hence the microseismic volume represents the
fracture storage capacity. (If the displacement is a shear slip rather than a tensile opening, then the volume available for
storage will be lower, including only the volume opened by misalignment of the irregular surfaces after the shear event.) The
fracture surface area is also of interest. The dominant frequency can be used to estimate a slip area, assuming some dynamic
failure model. Additional parameters related to stress release and seismic energy can also be considered. In terms of the
microseismic fracture volume and area, it is insightful to compare the estimated values with independent estimates of
hydraulic fracture volume and area.
Table 1 – Comparison of Microseismic Area, Volume, and Energy with Hydraulic Table 1 compares the pumped
Fracture Treatment Energy and Geometry volume with the microseismic
Stage 1 Stage 2 Stage 3 Stage 4 volume computed from the total
MS Volume (bbl) 0.82 0.15 0.075 0.30 seismic moment for the example
Frac Volume (bbl) 25300 25300 25300 25300 shown in Fig. 18. Note that the
% 0.0033% 0.00059% 0.00030% 0.0012% microseismic volume is several
MS Area (ft2) 19375 4018 3533 9040 orders of magnitude smaller than the
Frac Area (ft2) 7000000 - 6200000 -
% 0.28% - 0.057% -
injected volume. For two of the
MS Energy (J) 62500 11300 5700 22950
stages, the table also compares the
Frac Energy (J) 149040000000 149040000000 149040000000 149040000000 calculated total microseismic slip
% 0.0042% 0.00076% 0.00038% 0.0015% area (summed for all microseismic
events) with the area estimated from
the unconventional fracture model described later. The hydraulic fracture area for stages 1 and 3 was estimated by calibrating
a complex fracture model to the microseismic image [Cipolla et al., 2011a; Weng et al., 2011]. The microseismic area is also
much smaller than the modeled fracture area.
SPE 152165 5

As noted above, the volume associated with the microseismic events is an extremely small fraction of the injected
volume. Mass conservation, along with the comparably small energy balance percentages, suggests that the majority of the
hydraulic fracture dilation is aseismic (i.e., not detected seismically). Maxwell and Cipolla [2011c] describe the aseismic
deformation as low frequency, outside the bandwidth of seismic instrumentation. Furthermore the microseismic deformation
is likely shear so that in reality only a small portion of the total microseismic deformation is fracture dilation. It should also
be noted that the microseismic volume is an extremely small percentage (roughly 0.0000001) of the stimulated volume of
about 10,000,000 m3. If this deformation is assumed to represent the stimulated volume fracture porosity, the extremely low
value of this ratio casts doubt on what is sometimes described as the frac “rubblizing” the formation. The microseismic
fracture area is a more significant proportion (but still less than 1%) of the hydraulic fracture area, which supports the fact
that the microseismic activity describes the fracture geometry but not the entire fracture deformation. Regardless, the relative
microseismic deformation is useful for applications involving calibrating a fracture/geomechanical model.

Stimulated Volume and Well Productivity


Microseismic images can be used to estimate the portion of the reservoir that was affected by the hydraulic fracture
treatment, often referred to as Estimated Stimulated Volume (ESV) or Stimulated Reservoir Volume (SRV), but denoted
stimulated volume (SV) in this text. SV is calculated based on the spatial distribution and density of microseismic events.
These calculations can be performed using various 3D algorithms that vary in their approach to interpreting the microseismic
image (Daniels et al, 2007; Mayerhofer et al., 2008) and it may
not be possible to directly compare stimulated volume
calculations from different algorithms. In some cases well
productivity can be correlated to SV (Fig. 2), but these
correlations are typically limited to specific regions within a field
and similar stimulation designs (e.g. – water-fracs with similar
volume and size of proppant). SV can also be correlated to
production profiles in horizontal wells (Fig. 3). However, SV
18.0

16.0

14.0

12.0
Production, %

R² = 0.8013
10.0

Fig. 2 – Correlation between SV and well productivity in 8.0


the Barnett Shale (From SPE 90051; Fisher et al., 2004).
6.0

cannot provide insights into the actual fracture structure 4.0

and is often insufficient when treatment designs and/or 2.0

geologic environment vary significantly. 0.0

Stimulated volume calculations may provide an 400 500 600 700 800 900 100

ESV (Mft3)
estimate of the maximum extent of the hydraulic fracture. Fig. 3 – Example of correlating ESV to stage productivity in the
However, numerous corrections are required before the Eagle Ford (From SPE 136873; Inamdar et al., 2010).
most likely location of the hydraulic fracture can be
determined. The application of stimulated volume calculations is similar to peeling an onion, it requires peeling away a
number of layers before the useful portion is discovered. The corrections include location uncertainty and eliminating stress
induced events (Cipolla et al., 2011b). Once the most likely location of the hydraulic fracture is determined, the next step is
to evaluate fracture complexity. In some cases where fracture growth is primarily planar, stimulated volume is misapplied, as
SV is only an appropriate measure for complex hydraulic fractures. Cipolla et al. [2008a] introduced the fracture complexity
index (FCI), which provides a method to estimate fracture complexity using the ratio of the microseismic image width and
length. Note that the microseismic width should be corrected for location uncertainty to avoid misinterpreting location
uncertainty as an indication of fracture complexity. When FCI is less than 0.25, planar fracture growth should be considered
for subsequent applications.
When the microseismic interpretation indicates that hydraulic fracture growth is complex, the most important parameters
required for subsequent applications are hydraulic fracture geometry and the distribution of conductivity. Unfortunately,
stimulated volume calculation, even after the above corrections, cannot provide insights into the underlying hydraulic fracture
structure (i.e. – geometry) or the location of proppant (i.e. – distribution of fracture conductivity). In addition, although SV
may correlate to hydrocarbon production in some cases, geomechanical effects such as closing of propped and partially
propped fractures and/or decreasing propped fracture conductivity due to increasing closure stress may render the correlations
invalid as drawdown increases with time in the fracture system.
Stimulated volume calculations provide an important quantitative interpretation, but the applications are more qualitative.
In the past, SV was the primary interpretation for comparing microseismic images, but with the introduction of complex
hydraulic fracture models and “microseismic to reservoir simulation” workflows there are now more rigorous methods for
applying microseismic measurements. Stimulated volume may continue to be an important interpretation. However, future
6 SPE 152165

applications of SV will probably focus on fracture model calibration, evolving into an interpretation of “large scale” fracture
location rather than a region of enhanced permeability.

Comparing Microseismic Images


Many applications of microseismic interpretations require comparing microseismic images to identify the impact of changes
in either fracture design or completion methods, or to assess the impact of changes in geology on the microseismic response.
This section provides some guidelines to apply whenever the relative changes or differences between two microseismic
results are significant. The workflow associated with comparing the relative microseismic responses between fracs (stages or
wells) is somewhat simpler than the interpretation of the absolute fracture geometry discussed later. Comparison of the
relative microseismic responses typically involves assessing if the fracture geometry or induced microseismic deformation
differs. Relative interpretation of the fracture geometry is easier than the complete absolute interpretation of geometry, for
example when two fracture images are close together and tend to have similar accuracies (i.e. - potential systematic location
errors associated with the velocity model are the same in both cases) and only the precision associated with input arrival
times and directions need to be considered. Comparison of the microseismic deformation can be assessed by the number of
microseismic events, but more robustly by the total strength of the microseisms (cumulative moment), as discussed
previously. Diligent filtering of the data for consistent geophysical sensitivity and confidence between fracs is required to
ensure appropriate comparison. Both aspects of
comparative interpretation are highlighted here
using data examples.
Fig. 4 shows a map view of two stages in a
horizontal treatment well in the Barnett Shale. One
stage is approximately 2000’ from the monitoring
well and the other about 2400’. Microseismic
locations along with an interpreted geometry
encompassing 95% of the data are also shown. The
event locations have been normalized using a
minimum SNR cut-off of 5 for both stages. Note
that the more distant stage is slightly larger in
length (transverse to the well), and that there is an
overlap of almost 300’ between the two stages
along the well. The images of these two stages are
now compared to interpret the differences in
geometry imaged with the microseismicity.
First, the two stages are overlain with
uncertainties in order to examine whether the
slightly greater transverse length is significant. In
this geometry, the location error in the transverse
component of the frac is dominated by the
geometric directional error, determined to be 4° in
this case. This results in a ‘cone’ shape increasing
with distance away from the monitoring well, as
shown in the middle panel. The interpreted lengths
both fall within these average uncertainty lengths,
indicating that the interpreted extremes are
consistent within location uncertainty, and therefore
Fig. 4 - Map view of two stages in a horizontal treatment well. Top no significant differences in lengths are found
panel perf locations and the interpreted extent of the microseismicity. between the two stages. More statistically
Middle panel showing microseismic events and the location sophisticated comparisons can be made, including
uncertainty associated with the direction to the microseismicity.
Bottom panel shows the location uncertainty associated with the statistical tests between the interpreted geometries.
horizontal offset direction. The final panel of Fig. 4 shows the overlap in the
stages with the average location uncertainty in the
offset direction, as a function of radial orientation from the monitoring well. This offset uncertainty is controlled by the
relative timing accuracy between the s- and p-waves, and is determined to be about 50’. Therefore the observed overlap
between stages is significant, and can be used to assess the completion strategy, such as making trade-offs between favorable
fracture network connection versus the consequences of closely spaced stages resulting in uneconomic overstimulation or the
closure of fractures stimulated in earlier stages.
SPE 152165 7

The second aspect of a comparative interpretation


involves comparison of the number of events detected
-0. 8
stage A
between two fracs. Fig. 5 shows a plot of magnitude versus
-1. 0
stage B distance for two stages. More events are detected during the
-1. 2 closer stage B than the more distant stage A, although the
-1. 4 smallest magnitude event that is detected is lower. The
-1. 6 Complete impact of both aspects is needed to determine if one frac is
inherently more microseismically active or if the
-1. 8
Uniform interpretation is affected by observation well bias. Table 2
-2. 0
shows the number of events recorded for both stages and the
-2. 2 corresponding cumulative seismic moment. As discussed
-2. 4 previously, cumulative moment is the more robust metric for
-2. 6 microseismicity rate or deformation comparison, since one
-2. 8
Detection Limit relatively large magnitude microseismic event can be the
2000 2500 3000 3500 4000 4500 deformation equivalent of numerous small magnitude events.
Di sta nce (fee t) In this example, a higher minimum SNR of 3 was used for
Fig. 5 - Magnitude versus distance comparison of two event detection for the further stage, compared to the 2.5 used
stages. The red line represents the detection limit of the for the closer stage. In order to compare these stages, a
smallest event detected at various distances. The horizontal
blue line represents the minimum magnitude that is
consistent detection threshold must be used. The first step is
“uniformly” detected at all distances (-1.8). The green line thus to filter both stages and consider only events with a SNR
represents the “magnitude of completeness”: the value of 3 or greater. The corresponding number and cumulative
above which all events have been recorded (-1.7). moment are also shown in Table 2. In this case the results do
not change significantly and stage B still has more events and
a lower total strength. To fully quantify the
microseismic activity in each stage, magnitude- Table 2 - Number of events and cumulative seismic moment
frequency analysis is required. Fig. 6 shows the
# events/cumulative moment Stage A Stage B
comparative magnitude and cumulative frequency
data for this example. Note that at low magnitudes,
Raw data 397/1148 MN-m 1191/993 MN-m
the curve becomes flatter corresponding with the
inability to detect small magnitude events. Above
SNR > 3 397/1148 MN-m 1129/981 MN-m
magnitude -1.7, all events in both stages are detected.
Note that the frequency-magnitude plot shows a
Mag > -1.7 98/ 683MN-m 34/263MN-m
linear power law relationship above this level as
expected. This “magnitude of completeness” (defined
as the magnitude above which all events have been
detected) is slightly higher than the minimum detectable
magnitude at the largest offset, corresponding to a value of 10000
Uniform Complete
about -1.8. The total strength of events above this cutoff is stage A
also included in Table 2 as well as shown graphically in stage B

Fig. 6. It is interesting to note that when the data are 1000


Number of Events

normalized, stage A has significantly more events and


much greater total strength than stage B. Applications
involving comparison of the total microseismic 100

deformation with either number of events of total


microseismic strength require proper normalization for
detection levels and magnitude of completeness. 10

Finally, consider the design of a monitoring project to


compare the fracture response of two or more different
stimulation or completion designs. Prior to deciding the 1
-3.0 -2.5 -2.0 -1. 5 -1. 0 -0. 5
stages that are to be compared, it is critical to perform a M agnitude
pre-survey design study. Such design studies involve Fig. 6 Cumulative number of events versus magnitude for the
determining the minimum detectable magnitude and two stages. The vertical green line is the magnitude of
expected location accuracy at different positions (Maxwell completeness, above which a linear power law relationship
et al. 2003a), along with potential assessment of the exists. For reference the blue line is the uniformly detected
magnitude, and the vertical orange and red arrow indicates the
expected number of detected events (Maxwell, 2011b). smallest magnitude recorded in each stage from Fig. 5.
The output of this design study is an expected range within
which accurate microseismic results can be expected with a specific sensitivity. This range can then be used to identify the
planned stages of the treatment well(s) that can be effectively monitored and compared. Fig. 7 shows two conceptual
completion and stimulation designs that could be undertaken for such a multi-well project. The best scenario is to alternate
designs within each well, and then apply the aforementioned comparative analysis. Not only are neighboring stages most
8 SPE 152165

favorable for comparison, but such a design allows comparison between wells which could otherwise be complicated with
geologic variations under a scenario where different designs were executed in different wells. An unfavorable fracture
comparison is also depicted where the stages in the toe and heel of the well are varied, in which case comparison of the more
distant fracs would not be expected to result in high quality microseismic data. Comparative interpretation would then be
compromised.

Microseismic Applications
Microseismic interpretations can be applied in both
Frac 2 Frac 1 reservoir characterization and development. The
reservoir development applications can be divided
into three categories based on the timeframe of the
Comparison Valid application.

1. Real-time Fracture Control


a. Diversion and re-fracturing
Monitoring Well
b. Identification of geo-hazards
c. Stage modification
X Comparison Invalid 2. Completion Effectiveness
a. Completion efficiency
b. Fracture treatment design
Pre-Survey Design Limits c. Completion strategy

3. Field Development
Fig. 7 - Conceptual plot of an ideal treatment comparison (stages in a. Well placement
blue box for top two wells) and non-ideal comparison (stages in red box b. Well spacing
for bottom two wells) for two different comparisons overlain with c. Drainage patterns and recovery
design limits from a pre-survey design study.

Real-time applications of microseismic interpretations require fast and reliable event detection and location algorithms
along with associated QC attributes and skilled engineers to interpret the images and make immediate recommendations for
changes in the treatment execution. Stage modification might be better labeled “almost” real-time, as subsequent stages are
modified based on the results of previous stages. Within a short time after the microseismic interpretation is complete, the
interpretation can be used to evaluate the current completion strategy and modify future completions. However, fracture
treatment design and field development applications of microseismic interpretations typically require a minimum of 6-12
months of production data. Fracture treatment design, staging and perforation strategy, and field development applications
require integration with geological, geomechanical and geophysical data and more comprehensive engineering workflows.
The basic microseismic reservoir development application workflow is illustrated in Fig. 8, separating the applications
that are based primarily on visualization of the microseismic interpretation and the applications that require modeling.
Currently, Real-Time applications are based on visualization of the microseismic event patterns, ESV, event histograms, etc.
Completion Effectiveness applications such as improving staging and perforating, and evaluating completion efficiency are
also visualization-based applications. However, hydraulic fracture modeling and reservoir simulation are required for
fracture treatment design and completion strategy applications (optimum number of perforation clusters and stages).
The remainder of the text provides more detail on both the reservoir characterization and development applications of
microseismic interpretation, followed by a review of some of the most common misapplications of microseismic data to
provide background and perspective to the application workflows.

Reservoir Characterization Using Microseismic Measurements


Integrating microseismic with geological earth models and seismic reservoir characterization to understand the impact of
geology on the hydraulic fractures is a growing application. Changes in faulting, fractures, material properties and stress
conditions can all lead to variations in the hydraulic fractures. Interpretation relies on integration of the microseismic
interpretation with various other data including geologic models, seismic reservoir characterization, wireline borehole
images, geomechanical earth models and hydraulic fracture injection data. The integrated microseismic interpretation can
rely on either absolute or relative interpretation between stages and/or wells. Ultimately these studies could result in
predictive capabilities to assess what fracture geometry will be created and identify ‘sweet spots’ for strategic well
placement. The geological integration also aids understanding of the mechanism of anomalous microseismic patterns,
including height growth, asymmetric or preferential growth in certain directions.
SPE 152165 9

Geology,  Diversion and Re‐fracturing 
Geophysics &  Real‐Time  Identification of geo‐hazards 
Geomechanics Applications Stage modification 
Visualization 
based  Staging, perforating, and 
Microseismic  Applications Completion  completion efficiency
Interpretation Effectiveness
Modeling  Fracture treatment design
based  Completion Strategy
Applications
Reservoir  Field  Well Placement & Spacing
Characterization Development Drainage Patterns & Recovery

Fig. 8 – Basic microseismic application workflow illustrating visualization based applications (brown) and modeling based
applications (blue).

Fault activation can have a significant impact on the stimulation and can be investigated through microseismic attributes
as well as geologic confirmation of the fault. Microseismic magnitude and b-values will change as a fault is activated during
a hydraulic fracture. Furthermore, focal mechanisms can be used to distinguish microseismic deformation occurring along
orientations consistent with known geologic faults. Geologic evidence of suspected faulting indicated by the microseismic
interpretation can be detected using seismic ‘edge’ detection algorithms to find discontinuous reflectors in the seismic
reflection data and help improve the resolution and detect subtle faults. Alternatively, geologic borehole evidence such as
cores or wireline imaging can be used independently or supplementary to the seismic data. Obviously confidence in the
microseismic fault activation interpretation is increased with independent geologic verification of the fault (Fig. 9). Ideally
the fault would be identified beforehand, and the well placement and completion strategy adjusted to mitigate any risk of
fault activation. Real-time geo-hazard control can be realized through populating the microseismic data as they occur into the
geologic model, to validate the mitigation strategy.
Identification of fracture
 
complexity is clearly topical,
particularly in stimulation of
shales. Fracture complexity (with
fractures growing in multiple
directions) requires pre-existing
fractures in various directions and
relatively small differences in the
principal horizontal stresses.
Interpretation of pre-existing
fractures is similar to fault
activation, although the utility of
microseismic b-value and
magnitude changes is not as clear.
However, microseismic source
Fig. 9 - Left side shows results of an edge detection algorithm applied to a potential fault
mechanisms are useful to confirm
activation indentified through magnitude, b-values and mechanisms. Edge detection deformation in directions
indicates a lineation associated with a distinct fault located below the reservoir (right consistent with the fractures.
side). In this example, the fault acts as a barrier and limits the hydraulic fracture growth Additional seismic attributes such
to the NE (after Maxwell et al., 2011b).
as anisotropy determined from
either the microseismic or
reflection data can further help detect the rock fabric. Evidence of isotropic stresses can further confirm geomechanical
conditions favorable to generation of fracture complexity, as discussed next. Potential complexity could be identified prior to
the stimulation treatment and the fracture design and completion strategy could then be implemented based on expected
fracture geometries. Real-time monitoring can then be used to make refinements on a stage-by-stage basis. Careful filtering
for the most accurate microseismic locations is a critical step, to avoid apparent complexity in the microseismic data
associated with low confidence locations (Cipolla et al., 2011b).
10 SPE 152165

The geomechanical assessment of stress anisotropy and heterogeneity uses microseismic workflows similar to fracture
complexity. On the reservoir characterization side, however, additional integration of mechanical earth models based on
logging results and hydraulic fracture pressure data can be used to quantify the stress state. Stress heterogeneity can be
assessed through injection pressure, sonic logging and seismic reservoir characterization. For example, Daniels et al. [2007]
and Rich and Ammerman [2010] describe assessment of stress anisotropy and heterogeneity for a Barnett frac where the
microseismic data showed significant variability along the treatment well. Cipolla et al. [2010] describe a geomechanical
reconstruction of the stress state that confirms the observed fracture geometry changes related to changes in the stress
anisotropy (Fig. 10). In another case study, Maxwell et al. [2011a] describe identifying preferential fracture growth into
regions of the reservoir identified as having low
Poisson’s Ratio (PR) by amplitude-versus-offset
analysis of seismic reflection data (Fig. 11). In this
example, the variation in PR was interpreted to be
related to changes in the stress state and confirmed by
integrating with ISIP’s.

23.9
23.7
23.4
23.8
23.5 24.1
24.3
24.1
24.1
23.5

Fig. 11 - Map view of PR contours overlain with


microseismic data recorded during stimulation of
three wells in the Montney Shale (after Maxwell et al.,
2011). Yellow and red is low PR (~0.1) and blue and
purple is high PR (~0.3). Also plots are ISIP gradients
(MPa/m) for various stages. Note that the
microseismic shows asymmetric fractures growing
preferentially into the lower PR regions. ISIPs confirm
these are regions with lower stresses.

Fig. 10 - Map view of microseismic recorded during a 4 stage


treatment of a Barnett horizontal well (top) overlain on seismic
curvature (contours) and seismic fabric direction (arrows) (after Real-Time Applications
Rich and Ammerman, 2010). Bottom shows contours of stress The application of microseismic measurements in real-
anisotropy from a geomechanical reconstruction (after Cipolla et
al., 2010). The first two stages are in a region of higher stress time is focused primarily on the location of the
anisotropy and where the fabric is aligned with the maximum microseismic events, providing a measurement of the
stress direction resulting in creation of a relatively planar fracture top and/or bottom and the fracture location. The
fracture. The last two stages are in a region with more isotropic real-time applications are:
stresses, which along with the fabric more parallel to the well
results in more fracture complexity.
SPE 152165 11

Seismic Real‐Time Microseismic Interpretation - Diversion or re-fracturing


- Identifying and/or avoiding geo-hazards
Event locations , Event 
Structure
count,  Magnitudes, SV, b‐ - Height growth control
Natural Fractures, Faults
values,  Event histograms  - Stage modification

Visualization Software
The basic workflow for real-time microseismic
Earth Models
applications is shown in Fig. 12. The data requirements
Location of fracture and SV
Magnitude distribution (b‐Value)  for real-time microseismic applications are usually much
Geological Model
Magnitude and event location less than for other applications, but a basic geologic
Sh/P location graph
model is normally required for most real-time
applications. The input parameters required for real-time
Logs Treatment Modification interpretation consist of event locations, number of
Location of Natural 
Diversion events, magnitude, stimulated volume, b-value, and event
Fractures, Faults, Karsts, 
Rate histograms. The applications are primarily based on
Proppant schedule
etc.
Abort visual observations of microseismic behavior to identify
the location of the hydraulic fracture. Based on the real-
Fig. 12 – Basic workflow for real-time microseismic mapping time visualization of microseismic behavior, the treatment
applications.
can be modified to achieve the desired geometry or avoid
geo-hazards.

Diversion and Re-fracturing


It is often desirable to control where the fracture propagates, but very difficult in practice to successfully apply diversion
technology to effectively control fracture growth. Pressure measurements alone are typically not sufficient to reliably
evaluate diversion effectiveness and fracture location. Thus microseismic mapping has become the primary tool to evaluate
real-time treatment control. It is important to note that real-time applications such as diversion and re-fracturing rely only on

Initial Fracture Treatment ESV
Under‐stimulated 
Un‐stimulated 
reservoir region?
portion of lateral

Stage 1
ESV ESV

Stage 2 Stage 3
Fig. 13 - Case history #3, Real-Time diversion using MSM. Initial fracture treatment was did not effectively stimulate the lateral
(upper left graph). Re-fracture treatment stage 1 (yellow dots) propagated in the same portion of the lateral as the initial treatment
(upper right graph); as did the first diversion attempt, stage 2 (blue dots, bottom left graph). More aggressive diversion allowed
stage 3 (red dots) to direct fracture propagation to un-stimulation portion of the lateral and more effectively stimulate the under-
stimulated reservoir regions (lower right graph). Histogram charts show the location of events along the lateral for each re-
fracture stage (upper right), while estimated stimulated volume or ESV for each re-fracture stage is shown in the upper left of
each graph.
12 SPE 152165

the relative position of the microseismic events at various stages in the treatments and, as described in the comparative
interpretation section, are less sensitive to uncertainty in the event locations.
Fig. 13 shows how real-time MSM can be used to control the location of the hydraulic fracture using fiber diversion
technology to improve stimulation effectiveness (Daniels et. al. 2007). The initial stimulation on this well was monitored
using MSM, but real-time control technologies were not yet being routinely implemented and the fracture network coverage
of the lateral was inefficient, with about 30% of the lateral
un-stimulated (Fig. 13). The figure shows the interpretation
tools used for real-time applications; histograms of event
locations to locate the primary depth of fracture
propagation in the lateral (upper right corner) and
800 MCFD stimulated volume (ESV in this case, upper left corner) for
each re-fracture stage to identify the volumetric location of
fracture growth. Fig. 13 shows the microseismic event
distribution for the initial fracture treatments and the re-
500 MCFD
fracture treatment, illustrating the un-stimulated and under-
stimulated regions after the initial fracture treatment. Stage
1 of the re-fracture treatment did not contain any diverting
material and propagated in the same region as the initial
Fig. 14 – Production date for case history #3 showing a 60% fracture treatment (Fig. 13, upper right graph). Stage 2
production improvement after re-fracture treatment. contained diverting material in a modest concentration and
failed to change the location of fracture propagation (Fig.
13, bottom left graph). Based on the real-time application
of the microseismic interpretation, more aggressive concentrations of diverting material were pumped in stage 3, resulting in
significant improvement in the stimulation coverage (Fig. 13, lower right graph). The production improvement after the re-
stimulation is shown in Fig. 14. Initial well productivity more than doubled and longer term production improved by 60%.

Identifying and Avoiding Geo-Hazards, Height Growth Control


Fracturing into geo-hazards such as water-productive zones, faults, and karsts can significantly reduce stimulation
effectiveness. The real-time application of microseismic can be used to identify and avoid geo-hazards. When identifying
geo-hazards, the event magnitude and the temporal magnitude distribution are used in conjunction with event locations. Fig.
15 is an example of downward fracture height growth
into a water-productive zone due to fault activation.
Reservoir Top
The visualization of the spatial distribution of
microseismic events clearly shows an anomalous
pattern. Maxwell et al. [2008], Warpinski [2009], and
Well
Maxwell et al. [2011a] show that fault activation
during a hydraulic fracture treatment can be identified
Reservoir Bottom by the magnitude and location of microseismic events.
Top of Water
Fig. 16, from Maxwell et al. [2009], shows how event
magnitudes can be used to identify fault activation.
Fault activation indicated by high 
concentration of large magnitude events
Combining the event magnitude and the event
locations provides a reliable real-time tool to identify
Fig. 15 - Example of fault activation. Microseismic mapping shows a
fault activation. Maxwell at al. [2009] and Downie et
high concentration of large magnitude events with a spatial al. [2010] show how the distribution of event
distribution indicating the hydraulic fracture contacted a fault. magnitudes or b-values can provide an additional
diagnostic to identify fault activation. Fig. 17, from
Maxwell et al. [2009], illustrates how real-time application of b-value calculations can be used as an indicator of potential
fault activation. Microseismic event distributions during a typical hydraulic fracture treatment exhibit a b-value of
approximately two, while the b-value during episodes dominated by fault activation is around one. Real-time calculations of
b-values using a moving average time window may provide a tool to confirm fault activation much earlier in the treatment. In
this illustration, the deviation of the b-value from two around 125 minutes would be a warning sign of fault activation.
In many cases geo-hazards are already identified based on petrophysical or seismic measurements and the application of
real-time microseismic measurements simply requires the integration of the geologic model with the microseismic locations.
Most real-time visualization software can provide this integration. The real-time visualization of event locations and
magnitudes (events symbols are sized proportional to the event magnitude) can then be used to recognize that the hydraulic
fracture is propagating into a geo-hazard. Once propagation into a geo-hazard is identified there are a number of options to
modify the treatment including aborting the treatment, reducing the injection rate, and the application of diversion
technologies (fiber, proppant slugs, viscous pills, etc.). Waters et al. [2009b] provide an excellent example of using real-time
microseismic interpretations to adjust injection rate and apply diversion technology to avoid a fault. King et al. [2008]
discuss techniques to control downward growth into a water-productive zone that incorporated real-time microseismic
SPE 152165 13

measurements. In many cases the most prudent option is to abort the current stimulation treatment and proceed to the next
stage to minimize the risk of water production or economic waste (i.e. – fracturing into a fault or karsts). Unless the option to
use diversion technology was planned in advance of the treatment, the real-time application of microseismic mapping is
usually limited to adjusting the injection rate or aborting the treatment.

-2.2 4.0

3.5
-2.4

3.0
Magnitude
-2.6 Col 20 vs Col 25
2.5
Col 35
magnitude

Decrease in b‐value 

b-value
Col 42
-2.8 2.0
2D Graph 10 indicates fault 
-3.0 Large magnitude 
1.5 activation
Easting vs Northing

events indicate fault  1.0
-3.2
activation 0.5

-3.4 0.0
10:00:00 11:00:00 12:00:00 13:00:00 14:00:00 10:00:00 11:00:00 12:00:00 13:00:00 14:00:00
Time Time
Fig. 16 – Example of fault activation identified using event Fig. 17 – Illustration of real-time b-value calculations to
magnitude (from Maxwell et al., 2009). identify fault activation (from Maxwell et al., 2009).

Stage Modification
Stage modification might be considered a “near” real-time application of microseismic mapping, as the treatment design
and perforating strategy are changed in subsequent stages based on the microseismic interpretation of previous stages. The
primary interpretations required are fracture location and geometry and stimulated volume. Based on the microseismic
interpretations, the perforating strategy and/or treatment design may be changed on subsequent stages to improve vertical or
lateral wellbore stimulation and completion efficiency. Ejofodomi et al. [2010] illustrate how near real-time microseismic
mapping can be used to modify staging and perforating in vertical wells in West Texas, while Fisher et al. [2004] illustrate
this application in horizontal Barnett shale wells. The details of real-time stage modification are a subset of applications that
target improving completion effectiveness (discussed next).

Completion Effectiveness Applications


Microseismic mapping can be applied to improve completion effectiveness, resulting in more efficient and productive
completions. The application of microseismic interpretations to improve completion effectiveness consists of both
visualization-based applications and model-based applications (Fig. 8).
Completion effectiveness is controlled by three factors, the completion strategy, the stimulation design, and the
completion efficiency, with completion efficiency and completion strategy closely linked. Most unconventional resources
are developed using multiple fracture treatments with multiple perforation “clusters” in each fracture treatment stage. The
stage and cluster geometry is “designed” to evenly distribute stimulation fluid and proppant to each cluster, in an attempt to
create separate hydraulic fractures at each perforation cluster in each stage. There is a tradeoff between the number of stages,
number of perforation clusters per stage, spacing between perforation clusters, well productivity, and completion cost.
Completion efficiency is controlled by the connection between the reservoir and the wellbore and the stimulation
coverage of the target interval for a given completion strategy and stimulation design. The completion strategy consists of
parameters such as the distance between stages, number of perforation clusters, and/or distance between perforation clusters.
Even a good completion strategy can be ineffective if the completion efficiency is low due to such issues as poor cement.
For horizontal well completions, increasing the number of fracture treatment stages for a given lateral length will increase
productivity, but completion costs will also increase. There is typically a compromise between the spacing between clusters,
number of stages, and cost. For a given spacing and treatment volume per perforation cluster, increasing the number of
perforation clusters per stage will reduce the number of stages required, reducing completion time and cost. However,
production logs show that 25% or more of the perforation clusters in horizontal shale-gas completions are not producing
(Miller et al., 2011). In addition, this data shows that the percentage of non-productive perforation clusters increases as the
number of perforation clusters is increased.
14 SPE 152165

Completion Efficiency
The visualization-based applications are primarily
focused on improving the completion efficiency,
which include evaluating stage isolation techniques
such as cemented versus un-cemented completions,
identifying cement quality issues and assessing
perforation cluster spacing and location and stage
overlap. The workflow for these visualization-based
applications is essentially the same as that for real-
time applications (Fig. 12), except that the final step
is now “Completion Modification” instead of
“Treatment Modification”. Fig. 18 illustrates how
stimulated volume calculations can be visualized to
Fig. 18 – Example of stimulated volume showing less than optimum evaluate completion efficiency in a horizontal shale-
lateral coverage and opportunities to improve completion efficiency. gas completion (Daniels et al., 2007). In this example,
(after Daniels et al., 2007).
the four stage stimulation strategy did not stimulate
the entire lateral, as evidenced by the portions of the lateral that do not exhibit microseismic activity. In most shale-gas
reservoirs, the drainage area is limited to the proximity of the hydraulic fracture, thus large areas with no microseismic
activity can be assumed to be non-productive. Increasing the number of stages on subsequent wells will improve completion
efficiency. Fig. 19 is an example
of applying microseismic Stage Overlap: Poor Cement Multiple Fractures: Un‐cemented
interpretations to evaluate
completion efficiency in a tight
gas reservoir (Baihly et al., 2009).
This example shows stage overlap
due to poor cementing in a plug-
and-perf completion and multiple
fractures propagating from
isolation packers in an un-
cemented “ball-drop” completion.
Fig. 19 – Example of the application of microseismic mapping to evaluate completion
Fig. 20 shows an example of efficiency in a tight gas well (after Baihly et al., 2009). Left graph shows stage overlap in
stage overlap in a vertical tight gas a case and cemented completion (plug & perf). Right graph shows two fractures
well. In this example stages 1, 2, initiating for near isolation packers in an un-cemented completion (external casing
and 3 stimulate essential the same packers with frac ports actuated using balls).
vertical section. Subsequent
production logging measurements indicated that very little gas was being produced from stages 2 and 3. In this example,
eliminating stages 2 and 3 and pumping a larger treatment with perforations in the lower portion of the interval would reduce
completion costs and provide similar production.
g These examples illustrate how simple
visualization-based MS applications can
e6 6 8% be used to evaluate completion efficiency
and identify areas where future completion
e5 5 23% strategies and operational procedures can
be improved.
e4 4 42%

Completion Strategy and Stimulation


e3 3
4% Design
Microseismic interpretations can be
used to improve completion strategies and
0%
e2 2 stimulation designs. These applications are
e1 1 23% much more modeling intensive and can
require additional data and measurements.
The workflow for completion strategy and
fracture design applications of
Fig. 20 – Illustration of stage overlap in a vertical well (after Peterman et al., microseismic mapping is shown in Fig. 21.
2005). In this example, the bottom three stages overlap, suggesting they can be
combined into a single stage with a larger fracture treatment covering the entire
This workflow requires fit-for-purpose
interval. A production log shows that most of the gas in coming from stage 1, hydraulic fracture and reservoir simulation
indicating the location of the perforation clusters should be in the lower portion models and the ability to efficiently
if the interval. Fracture half-length, fracture conductivity, and permeability are integrate multi-discipline data and
shown in the table for reference.
interpretations. The workflows used when
SPE 152165 15

improving completion strategies and stimulation designs are very similar to those used for field development applications of
microseismic interpretations (discussed later). The primary difference is that the field development applications are focused
on calibrating the reservoir simulation model and determining drainage architecture, while completion effectiveness
applications are focused on calibrating the hydraulic fracture model and improving the completion strategy. Therefore, the
integration of production data and production history
matching are not included in the completion Seismic Microseismic Mapping
effectiveness workflow. Structure
Natural Fractures, Faults
Microseismic Image 
Applying microseismic interpretations to improve Rock Properties Interpretation & Advanced 
Processing
Stress Variations
stimulation designs and completion strategies starts
with a detailed Earth Model, which includes
Earth Models
integration of both seismic and wellbore Hydraulic Fracture Models
Completion Strategy
1D or 3D MEM
measurements. As discussed, microseismic Reservoir Model
Fracture model Calibration 
Fracture Geometry and  Perforation Locations 
measurements can also be an important input into Geological Model
DFN
Conductivity Distribution Number of Stages

seismic workflows, providing insights into the


relationships between hydraulic fracture propagation, Well Performance
Logs, Core & Petrophysics Reservoir Simulation Models
natural fractures and faults, structure, and rock Natural Fractures Production Logs
properties. The primary application of microseismic Rock Properties Production Profile & 
Stress Profile/ Anisotropy Hydrocarbon Recovery
interpretations in this workflow is for fracture model Reservoir Properties
calibration (reference Fig. 21). The appropriate Fig. 21 – Workflow for completion strategy and stimulation design
hydraulic fracture model will depend on the level of applications of microseismic mapping.
fracture complexity, with planar fracture models
being appropriate when fracture growth is relatively simple. However, complex fracture models are required for many shale
completions. The evaluation of fracture complexity and the calibration of complex fracture models are discussed by Cipolla
et al. [2011a]. Determining the degree of hydraulic fracture complexity is an important application of microseismic
measurements and a critical component of stimulation design (Cipolla et al., 2008a, 2008b).
The workflow shown in Fig. 21 assumes that the completion strategy will be determined independently, using simple
geometric spacing of perforation clusters and stages or more sophisticated algorithms that account for variations in reservoir
quality, rock properties and stress, location of nature fractures and faults, and wellbore configuration (Cipolla et al., 2011c) .
Production logs, if available, can provide important measurements to evaluate completion efficiency and calibrate completion
strategy algorithms. Comparing production log results to fracture modeling predictions of near-wellbore conductivity may
improve the fracture model calibration.
The primary microseismic interpretations required for fracture model calibration and improving completion strategy are
fracture location and either geometry (e.g. height and length) or stimulated volume, depending on the degree of fracture
complexity.

Example Application. A horizontal Barnett shale completion


is used to illustrate the application of microseismic
measurements for fracture model calibration and stimulation
design. This example draws from the work of Daniels et al.
[2007], Rich and Ammerman [2010] and Cipolla et al. [2010],
This well has a rich dataset, including microseismic monitoring
for all stages of the stimulation, advanced sonic logs that
provided estimates of minimum and maximum horizontal stress
and 3D seismic interpretations of curvature and natural fracture
orientations (Rich and Ammerman 2010). The 3200-ft lateral
was drilled in the direction of minimum horizontal stress (σh),
encouraging transverse hydraulic fractures, and four slickwater
fracture treatments were pumped. Each treatment consisted of
25,000 bbls of water and 440,000 lbs of proppant. The fracture
Fig. 22 – Microseismic event patterns for two fracture treatments were monitored using an array of geophones in an
treatment stages in a Barnett Shale completion. Based on offsetting wellbore. In this example the microseismic image has
the results presented by Daniels et al. [2007]. been patterned after the results presented by Daniels et al.
[2007], with the example focused on two of the four stages for
brevity. The microseismic images patterned after the stage 1 and 3 fracture treatments are shown in Fig. 22, illustrating
relatively linear event patterns in stage 1 and a mixture of complex and linear event patterns in stage 3. The difference in
event patterns was explained by variations in stress regime and natural fracture development (Daniels et al., 2007; Rich and
Ammerman, 2010; Cipolla et al., 2010).
16 SPE 152165

The combination of the 3D seismic interpretation,


advanced sonic log, and microseismic measurements
provide a much more reliable understanding of differences
in hydraulic fracture growth along the lateral and constrain
the DFN and Earth Model. A detailed geologic description
of the natural fractures was not available and the DFN was
generated using a stochastic model guided by the advanced
seismic interpretation and microseismic measurements
(Fig. 23). Minimum horizontal stress in the Barnett layer
varied from 0.70 psi/ft in the toe-section of the lateral to
0.62 psi/ft in the heel section, while maximum horizontal
stress ranged from 0.77 psi/ft to 0.65 psi/ft in the toe and
heel, respectively. The difference in the two horizontal
stresses varied from 0.10 psi/ft in the toe section of the
Example  lateral to 0.03 psi/ft in the heel.
horizontal  The complex hydraulic fracture propagation model
well (Weng et al., 2011) was calibrated to the microseismic
image (Fig. 22) for the two stages. The calibration
consisted of minor adjustments in the maximum horizontal
Fig. 23 – Discrete Fracture Network (DFN) representing the stress and the evaluation of numerous realizations of the
distribution of natural fractures in the vicinity of the example
horizontal well. Note that the dominant natural fractures DFN. The complex fracture geometry is compared to the
trending NE-SW in the toe section of the lateral are absent in microseismic image in Fig. 24, showing good agreement.
the heel section. As noted, the fracture model calibration assumes that the
hydraulic fracture geometry is defined by the microseismic
volume. Although not shown, the fracture was mostly
contained in the reservoir interval. Fig. 25 shows a three
dimensional view of the stage 1 complex fracture
geometry, illustrating that a large portion of the fracture
network is not propped. A large portion of the stage 3
fracture geometry was also not propped.
The complex fracture modeling results indicated that
un-propped fracture conductivity will play a critical role in
well productivity. It is often assumed that propped fracture
conductivity in shale plays is essentially infinite with
respect to the very tight matrix permeability; however the
relationships between propped fracture conductivity, un-
propped fracture conductivity, fracture complexity, and
well productivity are complex and not adequately
described using classical fracture conductivity theory
developed for planar hydraulic fractures (Cipolla et al.,
2008).
Fig. 24 – Comparison of complex hydraulic fracture geometry
and microseismic image (plan view). Evaluating the impact of variations in fracture
treatment designs, propped and un-propped fracture
conductivity, and well productivity requires detailed
Propped regions  reservoir simulation. Although reservoir simulation is
included in the completion effectiveness workflow, this
portion of the workflow will be discussed in the following
section for expediency. However, even without reservoir
simulation and production forecasting, the complex
fracture modeling results indicate that treatment design
changes that increase propped fracture area are warranted.
Once the fracture model is calibrated and the
stimulation design has been “optimized”, the final step in
the fracture design and staging strategy workflow is to
evaluate the impact of stage spacing and number of
perforation clusters per stage on well productivity and
Un‐propped regions  completion cost. As with fracture design, this step requires
Fig. 25 – Stage 1 fracture geometry showing proppant and un- an efficient integration of fracture modeling and reservoir
propped regions of the fracture network. simulation results and will be discussed in the context of
the field development workflow.
SPE 152165 17

Field Development Applications


The field development applications of microseismic mapping are focused on evaluating well productivity to determine
drainage architecture and hydrocarbon recovery. The basic workflow for field development applications (Fig. 26) is very
similar to the completion effectiveness workflow, but focuses on calibrating the reservoir simulation model by history
matching well performance. However, fracture model calibration remains in the field development workflow, as reservoir
simulation history matching results may indicate changes in proppant distribution, requiring additional fracture model
calibration.
The example presented in the previous section will be evaluated in further detail using reservoir simulation to determine
the effective fracture area or effective stimulated volume. Prior to the introduction of complex hydraulic fracture models, it
was difficult to estimate how much of the stimulated volume calculated from the microseismic image was actually productive
(Fig. 27). In addition, there was no direct “link” between the fracture treatment design, stimulated volume, and production
evaluation. With the introduction of complex hydraulic fracture models, there is now an opportunity to perform much more
rigorous evaluation. However, two obstacles must
be overcome; (1) efficient integration of Seismic Microseismic Mapping
Structure
microseismic data, complex fracture models, and Natural Fractures, Faults
Microseismic Image 
Interpretation & Advanced 
reservoir simulation models along with geological, Rock Properties
Stress Variations Processing
petrophysical, and geophysical interpretations on a
common software platform and (2) automated Earth Models Hydraulic Fracture Models
gridding routines to generate reservoir simulation 1D or 3D MEM Fracture Model Calibration
grids that exactly represent the fracture geometry Reservoir Model Fracture Geometry and 
Geological Model
and distribution of fracture conductivity predicted DFN
Conductivity Distribution

by the complex fracture model. These two obstacles


have been addressed with the recent introduction of Logs, Core & Petrophysics Reservoir Simulation Models
Well Performance
seismic-to-simulation software for unconventional Natural Fractures
Reservoir Modeling Calibration Production Data
reservoirs (Cipolla et al., 2011d). Rock Properties
Stress Profile/ Anisotropy
Hydrocarbon Recovery &  Production Logs
Drainage Architecture
The complex hydraulic fracture geometry for Reservoir Properties

stages 1 and 3 (Fig. 24) was automatically gridded Fig. 26 – Workflow for field development applications of microseismic
and input into the reservoir simulation model. mapping.
Previous history matching work using a much less
sophisticated complex fracture model was presented by Cipolla et al. [2010] and
indicated a propped fracture conductivity of 15 md-ft and an un-propped fracture Table 3 – Reservoir properties
conductivity of 0.03 md-ft. As a reference, un-propped and partially propped fracture top depth 6900 ft
conductivity values are presented in Fig. 28, with the range of closure stress for this reservoir thickness 400 ft
example highlighted. The history match value of 0.03 md-ft falls on the low end of reservoir pressure 3200 psi
the range. These values will be used as a starting point to illustrate the impact of un-
porosity 3%
propped conductivity on productivity, drainage area, and gas recovery. Table 3
shows the basic reservoir properties used in the reservoir simulations. permeability 100 nd
The pressure distribution after 20 years of production is shown in Fig. 29. The water saturation 30%
simulations were performed using a constant flowing bottomhole pressure of 1000 gas gravity 0.6
psi. The drainage architecture is dominated by the propped regions of the fracture
network (blue regions), with limited drainage outside the
propped regions (yellow/orange regions). The projected gas
recovery is 3.5 BCF in 20 years. There is very little pressure

1000
Fracture Conductivity (md‐ft)

Un‐propped with 
shear offset or 
100 partially propped

10

0.1 Un‐propped
No shear offset

0.01
0 1000 2000 3000 4000 5000 6000 7000 8000
Closure Stress (psi)
Fig. 27 – Stimulated volume calculated from the
microseismic image (shaded area). But how much is Fig. 28 – Un-propped and partially propped fracture conductivity
actually productive? (adapted from Fredd et al., 2001).
18 SPE 152165

depletion at the un-propped extremities of the fracture network, even after 20 years. Fig. 30 compares the microseismic
stimulated volume (Fig. 27) and the 20-year pressure depletion, illustrating that the effective stimulated volume is much less
that the microseismic stimulated volume due to the un-propped portions of the fracture network. This illustrates the
importance of integrating complex fracture modeling and reservoir simulation, as the distribution of propped and un-propped
conductivity will dictate productivity and drainage patterns. In this case, un-propped conductivity appears very low, limiting
drainage in the un-propped regions of the network.
The impact of un-propped fracture conductivity on
well productivity and gas recovery for un-propped
conductivities ranging from 0.0003 md-ft to 0.3 md-ft is
shown in Fig. 31. There is a dramatic difference in the
effective stimulated area and gas drainage depending on
the un-propped conductivity. Simulations were also
performed to evaluate the impact of increasing propped
fracture conductivity and the results indicated that 15 md-
ft was essentially infinite conductivity. Fig. 32 compares
the cumulative gas production for un-propped fracture
conductivities of 0.0003, 0.03, and 1 md-ft, showing that

Fig. 29 – Pressure distribution after 20-yrs. Propped FC = 15


md-ft, Un-propped FC = 0.03 md-ft. Cumulative gas production
= 3.5 BCF.

even an un-propped conductivity of 0.03 md-ft contributes


significantly to well productivity and gas recovery. This is
illustrated by comparing the gas recovery when un-propped
conductivity is 0.0003 md-ft to that of the history match value
of 0.03 md-ft; without the contribution of the un-propped
regions of the fracture network there is a 40% decrease in gas
production. Conversely, modifying fracture treatment designs to
prop more of the fracture network could result in significant Fig. 30 – Comparison of microseismic stimulated volume
and 20 year pressure depletion. The effective stimulated
improvements in well productivity and gas recovery, as volume is much less that the microseismic stimulated
evidenced by the 90% increase in gas recovery when fracture volume due to un-propped portions of the fracture
conductivity is increased to 1 md-ft in the un-propped regions of network.
the fracture network. When un-propped fracture conductivity is
“relatively” high, the effective stimulated volume approaches the microseismic volume.
The un-propped fracture conductivity presented in Fig. 28 is based on laboratory experiments using relatively high
modulus core. Un-propped fracture conductivity can be significantly less for lower modulus rock (Cipolla et al., 2008). In
addition, exposure to, and ineffective cleanup of, stimulation fluids may further reduce un-propped fracture conductivity.
Therefore the importance of un-propped fracture conductivity on well productivity and drainage area could differ

Un‐Propped FC = 1 md‐ft Un‐Propped FC = 0.0003 md‐ft
Propped FC = 15 md‐ft Propped FC = 15 md‐ft
Fig. 31 – Comparison of 20-year pressure depletion for un-propped conduct ivies of 1 md-ft (left) and 0.0003 md-ft (right).
SPE 152165 19

dramatically depending on the geological and


geomechanical environment.
This example illustrates the application of
microseismic mapping to determine drainage
architecture and hydrocarbon recovery. The
results show that the effective stimulated
volume can be significantly less than the
microseismic volume. In addition, un-propped
fracture area can contribute significantly to well
productivity and hydrocarbon recovery. In this
example, well spacing is dictated primarily by
the propped regions of the fracture network.
Increasing the percentage of the fracture that is
propped or partially propped could significantly
improve well productivity and increase
Fig. 32 – Comparison of 20-year cumulative production for un-propped
fracture conductivities of 0.0003 md-ft, 0.03 md-ft and 1 md-ft.
drainage area.

Well Spacing and Placement


The results of the previous example are used to illustrate the application of microseismic mapping for well spacing and
placement. This is probably one of the most important issues that impacts recovery factor and economics for unconventional
reservoir development. A detailed well spacing and placement evaluation is beyond the scope of this text, but the application
will be illustrated using drainage patterns based on the previous reservoir simulations.
Fig. 33 (left graph) shows the
approximate drainage patterns
for the base case reservoir
simulation results presented in
Fig. 29. The drainage patterns
are shown as rectangles for
simplicity. The center graph in
Fig. 33 illustrates the well
spacing required to develop the
area outlined by the rectangle
given the drainage patterns
predicted by the base-case
reservoir simulations. Four wells
are required and large areas are
not effectively drained. It is
Fig. 33 – Applying microseismic mapping for well placement and spacing
unlikely that the exact drainage
patterns would result in all wells,
but this is assumed in this case for illustrative purposes only. The right graph in Fig. 33 illustrates the potential impact of
increasing the propped area and pumping more stages, reducing the number of wells from four to three and improving
hydrocarbon recovery. Although this is only a conceptual example, the number of stages has increased steadily in recent
years and smaller proppants (e.g. – 100 mesh) have been used to increase the propped area in the Barnett.

Misapplication of Microseismic Interpretations


It is widely assumed that the observed microseismic events directly represent the hydraulic fracture geometry. However,
most of the microseismic events are only indirectly associated with the actual hydraulic fracture and are thus a proxy for the
hydraulic fracture. This section highlights the implications of this proxy effect and provides some guidelines to convert the
microseismic map into an accurate representation of the hydraulic fracture behavior. Hydraulic fracture operations typically
involve the pumping of large volumes of fluid (typically 100,000 gal to 1,000,000 gal or more), while the total microseismic
event volume (calculated from the cumulated magnitude of the events) is less than 1 barrel, as shown in Table 1. Clearly, the
pumped volume cannot be stored only in the volume created by microseismic events. In addition, the energy released by the
microseismic events is only a tiny fraction (e.g. - 1 ten thousandth) of the energy delivered by the pumps (Maxwell et al.,
2008). This non-seismic energy is consumed as friction losses and tensile opening of the hydraulic fracture planes. Finally,
the majority of observed microseismic events represent shear slip, which acts as a proxy for the tensile hydraulic fracture
opening (Maxwell and Cipolla, 2011). Sources of shear movement include shear stresses in the fracture process zone around
the tip, dog-legs in the generally planar fracture surface, and lubrication of pre-existing discontinuities by fluid leakoff.
Table 1 shows the calculated volume, area and energy of the microseismic events for the case study presented previously
in the text. The microseismic volume and energy are a tiny fraction of the total volume and energy input of the hydraulic
20 SPE 152165

fracture treatment, and the area is a slightly larger fraction (but still less than 1%) of the calculated hydraulic fracture surface
area. This highlights the concept of the microseismic events as a proxy for the actual hydraulic fracture network. Some
common misapplications of microseismic images are now highlighted.
Moment Tensor Inversion: As discussed previously, Moment Tensor Inversion (MTI) can be used to infer the mechanism
of the microseismic events. In some monitoring configurations (e.g. - single downhole monitor well), this inversion is non-
unique. Even when it is unique (e.g. - from multiple downhole, surface or near-surface monitoring configurations) it is
critical to understand that the source mechanism (opening, closing or shear) represents the deformation associated with the
event, not the hydraulic fracture. For example Prince et al. [2011] suggest that if the cumulative volume strain from events
during a treatment is negative, this net closing will result in reduced productivity in the reservoir. However, this statement
would only be valid if the events in fact represented the hydraulic fracture network after both treatment and flowback.
Furthermore, the example indicates predominant fracture closure while fluid is still being injected; suggesting that if the MTI
is correct, aseismic fracture opening must be the dominant factor for mass conservation. Any deformation during pumping
could very well be reversed during flowback and subsequent production. Perhaps more importantly, the only conductivity of
interest relies on the residual deformation (e.g. - Moos et al., 2011). It should also be noted that shear deformation of real
natural fractures results in some opening and residual conductivity (Fredd et al., 2001). Maxwell and Cipolla [2011c] show
schematically how the shear deformation could be associated directly with the HF, and also how subsequent events could be
associated with repeated deformation on the same area (Maxwell and Cipolla, 2011c, Fig. 10 and 11). For a very simple case
of a pressurized crack, Nagel and Sanchez (2011) showed how shear strain (and hence microseismic events) can often be
associated with deformation away from the fracture itself. It is thus essential to use a fracture model which accounts for the
interaction of hydraulic fractures and natural fractures to simulate the development of the fracture network during stimulation
(e.g.- Weng et al., 2011), and to calibrate this model with the microseismic interpretation (e.g.- Cipolla et al., 2011a), rather
than to assume the microseismic interpretation itself represents the fracture network.
Stimulated Volume: The concept of Stimulated Volume (SV) has evolved from simply drawing a rectilinear volume
around the events (Mayerhofer et al, 2008), through approaches which exclude small or isolated events (Daniels et al., 2007),
to algorithms which attempt to weight different parts of the stimulated volume differently, depending on the cumulative
magnitude of the microseismic events. The last approach is based on work originally published by Maxwell et al., (2003b),
and assumes that seismic deformation enhances permeability, which in turn leads to improved production. It is important to
recognize that while SV is a valid qualitative measure of disturbed rock volume, it is a purely visual concept. These more
advanced calculations of SV implicitly assume that more microseismic events imply better production. While this
assumption may be valid in very specific circumstances, it is not generally true. It should be obvious from Figs Fig. 27 to
Fig. 31 that production is very dependent on the fracture surface area, connectivity and conductivity. An individual
microseismic event can only contribute to production if it results in area, connectivity and conductivity, none of which are
captured in the calculation of SV. Furthermore, it is easy to overestimate the stimulated volume by ignoring event location
uncertainty. The tempting assumption that “more events are better” encourages the use of lower magnitude events and/or
lower SNR events. Typically these have higher location uncertainty, further increasing the volume over-estimation. The
workflow to obtain a more realistic SV is outlined by Cipolla et al (2011b).
Underestimating hydraulic fracture dimensions: Assuming the SV is accurately determined by appropriately filtering of
the event set and accounting for location error, it is commonly assumed that the fracture network lies entirely within the SV.
However, it is important to account for distance bias. It is possible that some parts of the stimulation occurred too far from
the observation well to be detected. Cipolla et al (2011b) explain how the magnitude-distance plot can be used to determine
whether all events have been observed, and also how to eliminate the effects of distance bias. Unless there is a good
explanation for the observed lack of symmetry, such as geological variations or interaction with a depleted zone, or the
magnitude-distance plot clearly shows that all events would have been observed, any asymmetry should be assumed to be an
artifact. The geophysical QC process also enables the identification of nodal planes which could explain unobserved events.
Furthermore, since the majority of the fracture tensile opening is slow (and hence aseismic), it cannot be assumed that
lack of microseismic events implies there is no fracture. Although the events do not represent the actual fracture, they
generally outline the location of the fracture, as discussed previously. There are some exceptions, such as activation of a
fault, but these exceptions can generally be identified via other characteristics of the microseismic events, such as the b-value
and changes in magnitude and source orientation (discussed previously).
Misinterpreting fracture complexity: As shown by Cipolla et al (2011b), event location uncertainty may result in
apparent complexity, i.e. the cloud of events being assumed to represent complexity when in fact the events could equally
well lie in a plane. Even if the events do not lie exactly in a plane, they could be the result of slip on natural features near the
hydraulic fracture. There are several obvious mechanisms to trigger slip close to, but not exactly on, the fracture plane, such
as leakoff, which may lubricate a natural feature, reducing the friction coefficient, or stress changes causing additional shear
stress on these nearby features.
Ignoring mass balance and fracture mechanics: As noted previously, the microseismic volume is a small fraction of the
total pumped volume. As such, any representation of the hydraulic fracture network must account for aseismic tensile
opening and storage of the massive volumes of fluid and proppant pumped. In addition, fractures propagate perpendicular to
the minimum stress, and are affected by natural fractures and other rock fabric. Hence the simulation of hydraulic fracture
propagation, accounting for geology, rock properties, and stress, is essential to explain any microseismic interpretation.
SPE 152165 21

Interpretation of overlapping microseismic volumes between different stages or wells: If the concept of SV is used
without consideration of fracture conductivity, it is possible to assume that overlapping microseismic volumes imply
inefficient stimulation. For example, it may be stated that some region is over-stimulated or that no more SV is being
created. However, if the fracture networks are not fully conductive, such as the extremities of those in Fig. 31, then overlap
is necessary to generate conductivity and access the resource.
The misapplications discussed in this section highlight three critical areas that must be considered for the proper
application of microseismic interpretations:
1. Geomechanics, specifically the increase in shear stress on natural fractures during and after stimulation,
2. Complex Hydraulic Fracture simulation, accounting for the creation of connected surface area, proppant
distribution, and other conductivity, and
3. Explicit representation of hydraulic fracture planes in production simulation.

Summary
The keys to applying microseismic measurements are high quality geophysical processing, reliable and consistent
interpretations and the integration of G&G data, fit-for-purpose fracture models, and reservoir simulation. In addition to
engineering applications, microseismic measurements have reservoir characterization applications, providing important
insights into natural fracture characteristics and stress regime. The complex workflows for Completion Effectiveness and
Field Development applications of microseismic interpretations are now practical with the recent introduction of complex
hydraulic fracture models, automated reservoir simulation gridding routines, and common software platforms where diverse
datasets, interpretations, and models can be shared by engineers and geo-scientists.
Advanced processing of the microseismic waveforms holds great promise to provide significant insights into natural
fracture characterization and to constrain hydraulic fracture models. However, microseismic deformation is a very small
fraction of the total deformation and current interpretations cannot provide reliable insights into hydraulic fracture area,
which is one of the key components that dictate well productivity. Nevertheless, changes in the microseismic response during
hydraulic fracturing can be used to indentify fault activation, and help constrain the relative deformation predicted by
geomechanical models. Attempting to define hydraulic fracture “opening” or “closing” using advanced processing such as
moment tensor inversion (MTI) can be fraught with error and lead to misapplications. Therefore, current applications of
advanced microseismic processing are limited primarily to reservoir and geomechanical characterizations (e.g. – natural
fractures, faults, stress regime).
A number of important issues associated with advanced processing and stimulated volume interpretations impact the
application of the results. Microseismic mapping may help define the location/distribution/orientation of the natural fractures,
but cannot be used to characterize hydraulic fracture area or volume (not to be confused with MS stimulated volume). There
is a significant potential for misapplication of advanced processing and SV interpretations if application workflows omit the
following components or fail to recognize the limitations of the results.
• Mass balance and fracture mechanics are required to estimate the fracture geometry.
• Fracture models are required to estimate the location and concentration of proppant.
• Geomechanics (e.g. – frac conductivity versus closure stress) is required to estimate the conductivity of the propped
and un-propped regions.
• Discretely modeling of the hydraulic fracture in a reservoir simulation model is required to understand the
relationship between MS, fracture geometry, fracture conductivity, and production. This important step provides the
“link” between MS behavior and well performance.
Most misapplications of microseismic data are associated with the wrong assumption that the microseismic events can be
directly correlated to hydraulic fracture geometry and well productivity. Microseismic images are excellent indicators of
hydraulic fracture location, but current processing and interpretations cannot provide reliable estimates of complex fracture
geometry or the distribution of conductivity within the fracture, except when integrated with other data, interpretations and
simulations.
The integration of microseismic mapping, fracture modeling and reservoir simulation is required to estimate the effective
stimulated volume. Although microseismic stimulated volume may be correlated to well productivity in limited areas of some
unconventional plays, variations in fracture treatment design, rock properties, and stress regime can result in significant
differences in the effective stimulated volume for the same microseismic volume. The complexity of the relationship between
microseismic stimulated volume and effective stimulated volume (SV) precludes simple correlations in most cases. One of
the most common misapplications of microseismic interpretations is the assumption that larger SV will automatically result
in increased well productivity. In areas where un-propped fracture conductivity is relatively large with respect to the matrix
permeability, the effective stimulated volume may approach the stimulated volume. Characterizing un-propped fracture
conductivity will be a critical factor when evaluating well performance and estimating drainage area and hydrocarbon
recovery.
22 SPE 152165

Acknowledgements
The authors thank Schlumberger for supporting this work and publication of this paper.

Nomenclature
AVO = Amplitude Versus Offset
bbl = barrels, L3
BCFGE = billion standard cubic feet of gas equivalent, L3
bpm = barrels per minute, L3/t
DFN = Discrete Fracture Network
ESV = Estimated Stimulated Volume, L3
FC = Fracture Conductivity (md-ft)
FCI = Fracture complexity index
G&G = geological and geophysical
Gal = gallons, L3
ISIP = Instantaneous shut-in pressure, F/ L2
k = permeability
lb = pounds, M
md = 10-3 Darcy, L2
Mcf/D = 1000 standard cubic feet per day, L3/t
MMCF, MMscf = million standard cubic feet, L3
Mscf/d, MCFD = 1000 standard cubic feet per day, L3/t
MS, MSM = microseismic, microseismic mapping, microseismic monitoring
MTI = Moment Tensor Inversion
nd = 10-9 Darcy, L2
p = pressure, F/ L2
PR = Poisson’s ratio
QC = Quality Control
SNR = Signal to Noise Ratio
SV = Stimulated Volume, L3
SRV = Stimulated Reservoir Volume, L3
UFM = Unconventional Fracture Model
σh = minimum horizontal stress, F/L2
σH = maximum horizontal stress, F/L2

SI Metric Conversion Factors


bbl x 1.589 874 e-01 = m3
ft x 3.048 e-01 = m
psi x 6.894 757 e+00 = kPa

References
Baihly, J., Grant, D., Fan. L., and Bodwadkar, S. 2009. Horizontal Wells in Tight Gas Sands – A Method for Risk Management To
Maximize Success. SPE Production & Operations Journal, May 2009, p277-288.
Cipolla, C.L., Warpinski, N.R., Mayerhofer, M.J., and Lolon, E. P. 2008a. The Relationship between Fracture Complexity, Reservoir
Properties, and Fracture Treatment Design. Paper SPE 115769 presented at the 2008 SPE Annual Technical Conference and
Exhibition, Denver, Colorado, USA, 21-24 September.
Cipolla, C.R., Warpinski, N.R., Mayerhofer, M.J. 2008b. Hydraulic Fracture Complexity: Diagnosis, Remediation, and Exploitation. Paper
SPE 115771 presented at the 2008 Asia Pacific Oil & Gas Conference, Perth, Australia, October 20-22
Cipolla, C.L., Williams, M.J., Weng, X., Mack, M., and Maxwell, S. 2010. Hydraulic Fracture Monitoring to Reservoir Simulation:
Maximizing Value. Paper SPE 133877 presented at the SPE Technical Conference and Exhibition held in Florence, Italy, September
19-22.
Cipolla, C., Weng, X., Mack, M., Ganguly, U., Gu, H., Kresse, O., Cohen, C., and Wu, R. 2011a. Integrating Microseismic Mapping and
Complex Fracture Modeling to Characterize Fracture Complexity. Paper 140185 presented at the SPE Hydraulic Fracturing
Technology Conference and Exhibition held in The Woodlands, Texas, USA, 24–26 January.
Cipolla, C., Maxwell, S., Mack, M, and Downie, R. 2011b. A Practical Guide to Interpretation Microseismic Measurements. SPE 144067
presented at the SPE North American Unconventional Gas Conference and Exhibition, The Woodlands, Texas, USA, 14-16 June
2011.
Cipolla, C., Weng, X., Onda, H., Nadaraja, T., Ganguly, U., and Malpani, R. 2011c. New Algorithms and Integrated Workflow for Tight
Gas and Shale Completions. SPE 146872 presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, 30
October–2 November 2011.
Cipolla, C., Fitzpatrick, T., Williams, M., and Ganguly, U. 2011d. Seismic to Simulation for Unconventional Reservoirs. SPE 146876
presented at the SPE Reservoir Characterization and Simulation Conference and Exhibition held in Abu Dhabi, UAE, 9–11 October
2011.
SPE 152165 23

Daniels, J., Waters, G., LeCalvez, J., Lassek, J., and Bentley, D. 2007. Contacting More of the Barnett Shale Through an Integration of
Real-Time Microseismic Monitoring, Petrophysics, and Hydraulic Fracture Design. Paper SPE 110562 presented at the 2007 SPE
Annual Technical Conference and Exhibition, Anaheim, California, USA, October 12-14.
Downie, R. C., Kronenberger, E. and Maxwell, S.C. 2010. Using Microseismic Source Parameters to Evaluate the Influence of Faults on
Fracture Treatments - A Geophysical Approach to Interpretation. Paper SPE 134772, presented at the SPE Annual Technical
Conference and Exhibition, Florence, Italy, 19–22 September.
Ejofodomi, E., Bizzel, K., Long, T., and Mills, D., Yates, M., Downie, R., Itibrout, T., and Catoi, A. 2010. Improving Well Completion via
Real-Time Microseismic Monitoring: A West Texas Case Study. Presented at the SPE Tight Gas Completions Conference held in San
Antonio, Texas, USA, 2–3 November 2010.
Fisher, M.K., Davidson, B.M., Goodwin, A.K., Fielder, E.O., Buckler, W.S., and Steinberger, N.P. 2002. Integrating Fracture Mapping
Technologies to Optimize Stimulations in the Barnett Shale. Paper SPE 77441 presented at the 2002 SPE Annual Technical
Conference and Exhibition, San Antonio, Texas, USA, September 29-October 2.
Fisher, M.K., Heinze, J.R., Harris, C.D., Davidson, B.M., Wright, C.A., and Dunn, K.P. 2004. Optimizing Horizontal Completion
Techniques in the Barnett Shale Using Microseismic Fracture Mapping, paper SPE 90051 presented at the SPE Annual Technical
Conference and Exhibition, Houston, Texas, USA, 26-29 September.
Fredd, C.N., McConnell, S.B., Boney, C.L., and England, K.W. 2001. Experimental Study of Fracture Conductivity for Water-Fracturing
and Conventional Fracturing Applications. SPE Journal Vol. 6 No. 3 pp 288-298 (September 2001).
Inamdar, A., Malpani, R., Atwood, K., Brook, K., Erwemi, A., Ogunda, T., and Purcell, D. 2010. Evaluation of Stimulation Techniques
Using Microseismic Mapping in the Eagle Ford Shale. SPE 136873 presented at the SPE Tight Gas Completions Conference, San
Antonio, Texas, USA, 2-3 November 2010.
Kim, A., 2011, Uncertainties In Full Waveform Moment Tensor Inversion Due To Limited Microseismic Monitoring Array Geometry,
SEG Expanded Abstracts 30, 1509 (2011), DOI:10.1190/1.3627488
King, G.E., Haile, L., Shuss, J., and Dobkins, T.A. 2008. Increasing Fracture Path Complexity and Controlling Downward Fracture
Growth in the Barnett Shale. Paper SPE 119896 presented at the 2008 SPE Shale Gas Production Conference, Fort Worth, Texas,
USA, 16–18 November.
King, G.E. 2010. Thirty Years of Gas Shale Fracturing: What Have We Learned? Paper SPE 133456 presented at the SPE Annual
Technical Conference and Exhibition, Florence, Italy, 19-22 September.
Leaney, S., and Chapman, C., 2010. Microseismic Sources in Anisotropic Media. Presented at the EAGE Conference and Exhibition,
Barcelona, Spain, 14-17 June 2010.
Maxwell, S.C., Urbancic, T., and McLellan, P., 2003a, Assessing the Feasibility of Reservoir Monitoring Using Induced Seismicity,
presented at EAGE Conference and Exhibition, Stavenger, Norway, 2-5 June 2003.
Maxwell, S.C., Urbancic, T., Prince, M., and Demerling, C., 2003b, Passive Imaging of Seismic Deformation Associated With Steam
Injection in Western Canada, Paper SPE 84572,presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado,
5-8 October 2003.
Maxwell, S.C., Waltman, C.K., Warpinski, N.R., Mayerhofer, M.J., and Boroumand, N. 2006. Imaging Seismic Deformation Associated
with Hydraulic Fracture Complexity, SPE 102801 Annual Technical Conference and Exhibition, San Antonio, Texas, U.S.A., 24–27
September.
Maxwell, S.C., Shemeta, J., Campbell, E., and Quirk. D. 2008. Microseismic Deformation Rate Monitoring. Paper SPE 116596, presented
at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, USA, 21–24 September.
Maxwell, S.C., Jones, M., Parker, R., Miong, S., Leaney, S., Dorval, D., D’Amico, D., Logel, J., Anderson, E., Hammermaster, K. 2009.
Fault Activation During Hydraulic Fracturing. SEG Expanded Abstracts 28, 1552 (2009), DOI:10.1190/1.3255145
Maxwell, S.C., Pope, T., Cipolla, C., Mack, M., Trimbitasu, L., Norton, M., and Leonard, J. 2011a. Understanding Hydraulic Fracture
Variability Through Integrating Microseismicity and Seismic Reservoir Characterization. SPE 144207 presented at the SPE North
American Unconventional Gas Conference and Exhibition, The Woodlands, Texas, USA, 14-16 June 2011.
Maxwell, S.C., 2011b, Microseismic Network Design: Estimating the Number of Detected Microseismic Events During Hydraulic
Fracturing, SEG Expanded Abstracts 30, 4404 (2011), DOI:10.1190/1.3658769
Maxwell and Cipolla. 2011c. What Does Microseismicity Tell Us About Hydraulic Fracturing? SPE 146932 presented at the SPE Annual
Technical Conference and Exhibition, Denver, Colorado, USA, 30 October–2 November 2011.
Mayerhofer, M.J., Lolon, E.P., Warpinski, N.R., Cipolla, C.L., Walser, D.,Rightmire, C.M. 2008. What is Stimulated Reservoir Volume
(SRV)? SPE 119890 presented at the 2008 SPE Shale Gas Production Conference, Fort Worth, Texas, U.S.A., 16–18 November 2008.
Miller, C., Waters, G., and Rylander, E. 2011. Evaluation of Production Log Data from Horizontal Wells Drilled in Organic Shales. SPE
144326 presented at the SPE Americas Unconventional Gas Conference, The Woodlands, Texas, USA, 14-16 June 2011.
Moschovidis, Z. et al. 2000. The Mounds Drill-Cuttings Injection Field Experiment: Final Results and Conclusions. SPE 59115 presented
at the 2000 IADC/SPE Drilling Conference, New Orleans, Louisiana, 23–25 February 2000.
Moos, D., Vassilellis, G., Cade, R., Franquet, J., Lacazettel, A., Bourtembourg, E., and Daniel, G. 2011. SPE 145849 presented at SPE
Annual Technical Conference and Exhibition, 30 October-2 November 2011, Denver, Colorado, USA.
Nagel, N. and Sanchez-Nagel, M. 2011. Stress Shadowing and Microseismic Events: A Numerical Evaluation. SPE 147363 presented at
SPE Annual Technical Conference and Exhibition, 30 October-2 November 2011, Denver, Colorado, USA.
Peterman, F., McCarley, D.L., Tanner, K.V., Calvez, J.H., Grant, W.D., Hals, C.F., Bennett, L., and Palacio, J.C. 2005. Hydraulic Fracture
Monitoring as a Tool to Improve Reservoir Management. Paper SPE 94048 presented at the SPE Production and Operations
Symposium, Oklahoma City, Oklahoma, 17-19 April.
Prince, M., Baig, A., Urbancic, T., Buckingham, K., and Guest, A. 2011. Using Moment Tensors To Determine the Effectiveness of
Hydraulic Fracture Stage Spacing. SPE 144190 presented at the SPE North American Unconventional Gas Conference and
Exhibition, The Woodlands, Texas, USA, 14-16 June 2011.
Rich, J.P. and Ammerman, M. 2010. Unconventional Geophysics for Unconventional Plays. Paper SPE 131779 presented at the
Unconventional Gas Conference, Pittsburgh, Pennsylvania, USA, February 23-25.
24 SPE 152165

Rutledge, J.T., Phillips, W.S, and Mayerhofer, M.J., 2004, Faulting Induced by Forced Fluid Injection and Fluid Flow Forced by Faulting:
An Interpretation of Hydraulic-Fracture Microseismicity, Carthage Cotton Valley Gas Field, Texas, Bulletin of Seismological Society
of America, 94-5: 1817.
Waters, G., Dean, B., Downie, R., Kerrihard, K., Austbo, L., and McPherson, B. 2009a. Simultaneous Hydraulic Fracturing of Adjacent
Horizontal Wells in the Woodford Shale. Paper SPE119635 presented at the SPE Hydraulic Fracturing Tech. Conference., The
Woodlands, Texas, USA, 19–21 January.
Waters, G., Ramakrishnan, H., Daniels, J., Bentley, D., Belhadi, J., and Ammerman, M. 2009b. Utilization of Real Time Microseismic
Monitoring and Hydraulic Fracture Diversion Technology in the Completion of Barnett Shale Horizontal Wells. Paper OTC 20268
presented at the Offshore Technology Conference held in Houston, Texas, USA, 4–7 May 2009.
Warpinski, N.R., Branagan, P.T., Peterson, R.E., Wolhart, S.L. and Uhl, J.E. 1998. Mapping Hydraulic Fracture Growth and Geometry
Using Microseismic Events Detected by a Wireline Retrievable Accelerometer Array. SPE 40014, 1998 Gas Technology Symposium,
Calgary, Alberta, Canada, March 15-18.
Warpinski, N.R. 2009. Integrating Microseismic Monitoring With Well Completions, Reservoir Behavior, and Rock Mechanics. Paper SPE
125239 presented at the Tight Gas Completions Conference, San Antonio, Texas, June 15-17.
Weng, X., Kresse, O., Cohen, C., Wu, R. and Gu, H. 2011. Modeling of Hydraulic Fracture Network Propagation in a Naturally Fractured
Formation. Paper SPE 140253 presented at the SPE Hydraulic Fracturing Technology Conference and Exhibition held in The
Woodlands, Texas, USA, 24–26 January 2011.

You might also like