You are on page 1of 10

ISSN 1062-7391, Journal of Mining Science, 2015, Vol. 51, No. 3, pp. 477–486. © Pleiades Publishing, Ltd., 2015.

Original Russian Text © S.V. Yaskevich, V.Yu. Grechka, A.A. Duchkov, 2014, published in Fiziko-Tekhnicheskie Problemy Razrabotki Poleznykh Iskopaemykh,
2014, No. 6, pp. 41–52.

_________________________________ GEOMECHANICS _______________________________


____________________________________________________________________________________________________________________________________ ___________________________________________________________________________________________________________________________

Processing Microseismic Monitoring Data, Considering


Seismic Anisotropy of Rocks
S. V. Yaskevicha, V. Yu. Grechkab, and A. A. Duchkova
a
Chinakal Institute of Mining, Siberian Branch, Russian Academy of Sciences,
Krasnyi pr. 54, Novosibirsk, 630091 Russia
e-mail: yaskevichsv@gmail.com
b
Marathon Oil,
5555 San Felipe St, Houston, TX 77056, USA
Received October 21, 2014

Abstract—Using the model information and in situ data on hydrofracturing in an oil and gas reservoir of
the Bakken Formation (USA), potential of locating hypocenters of microseismic events concurrently with
determining parameters of velocity anisotropy of seismic waves in rock mass is analyzed. It is shown that
inclusion of anisotropy in the analysis improves accuracy of spatial location of microseismic event
hypocenters and increases validity of estimation of the fracturing direction.
Keywords: Microseismic monitoring, rock mass, velocity model, anisotropy.
DOI: 10.1134/S1062739115030084

INTRODUCTION
Microseismic monitoring is one of the most efficient modern techniques of observation of induced
geodynamic events due to underground mineral mining. This method requires installation of sensor
array at an object and recording of the object’s microseismic activity. Spatial distribution (cloud) of
hypocenter of microseismic events allows location of geodynamic processes and their energy and
kinematic characteristics. It is possible to use these data to assess efficiency of various operations (for
instance, hydraulic fracturing [1]) and to optimize the mining method in use.
One of the most important areas of microseismic monitoring application is hydrofracturing of
reservoirs (HFR). Data on shape, size and orientation of hydrofractures are of importance for
optimization of hydrocarbon recovery technologies, in particular, gas recovery from shale and coal.
The current microseismic monitoring uses borehole [2–4] and surface [5, 6] systems of observation.
The reliable location of dynamic event hypocenters in microseismic monitoring data processing needs
the velocity model of the medium. The significance of the accurate velocity model for microseismic
monitoring is commonly acknowledged [7–11], but the effect of the inaccuracy of the model on the
overall error of the event hypocenter location is scarcely studied [12]. The velocity modeling is
complicated by that many unconventional hydrocarbons are represented by shales which possess
heavily anisotropic properties [13, 14]. Moreover, natural anisotropy may combine with anisotropy of
jointing induced by hydrofractuirng [15], which should also be included in the velocity modeling.
Usually the velocity model for the microseismic monitoring data processing is based on the
interpretation of the acoustic logging data, which are then calibrated based on the data on perforation
shots the source locations of which are known exactly. The anisotropic velocity modeling in this case
is a difficult problem. In [16] the authors determined anisotropy parameters in a transversely isotropic
medium with a vertical axis of symmetry (VTI) by the data on the arrival of waves due to the
perforation shots. This approach to determination of compound anisotropy is complicated, and
anisotropy parameters are recovered partly, due to limited aperture of observation system [17–20].
In practical microseismic monitoring, it is only possible to enlarge the observation system aperture
by means of using the data on microseismic event in the velocity modeling [21, 22]. In this case, the

477
478 YASKEVICH et al.

medium anisotropy parameters can be found later on, simultaneously with location of hypocenter of
microseisimic events [23, 24].
This article analyzes the inclusion of the seismic anisotropy in the data of borehole microseismic
monitoring data processing, and presents the mathematical formulation and solution for an inverse
problem aimed at concurrent determination of hypocenters of microseismic events and anisotropic
velocity modeling. The authors perform a model experiment to simulate microseismic monitoring in
the anisotropic model. It is shown that inversion of such data under assumed isotropy of the model
may result in false image of a fracture, its location and azimuth.
The developed approach has applied to processing of actual data on microseismic monitoring of
hudrofracturing at Bakken deposit in USA. Polarization analysis of the data has shown stable splitting
of the transversal waves, which is a reliable indication of seismic anisotropy. As a result of the data
inversion, the location of the hypocenters of the events has been refined, and the stratified model
anisotropy parameters have been determined.
1. PROBLEM FORMULATION
Under consideration is the inverse kinematic problem on selection of the arrival times for
longitudinal (P) and two transversal (S1 and S2) waves for anisotropy models. The data vector is
compose of the observe arrival times of the direct waves {t Qobs
еr
} , where Q—type of a wave (P, S1 or
S2), arrived in an r-th sensor for an e-th microseismic event. The desired vector of the model
parameters (position of horizontal interfaces of the strata is assumed known and fixed) has a form of:
m ≡ [ cl , x е ,τ е ] , (1)
where cl —set of anisotropic parameters for an l-th stratum ( l = 1, ..., N l ), N l —number of strata;
( x е ,τ е )—hypocenter coordinates ad origination time, respectively, for the e-th microseismic event
( e = 1, ..., N е ), N е —number of the events.
The direct problem consists 9in calculating the wave arrival times {t Qsyn ( m )} for a preset model m
еr
(see Fig. 1), using the algorithm of ray tracing for stratified–homogenous anisotropic media [25].
The inverse problem on the vector of the model parameters, m, is formulated as minimization of
the deficiency functionals between wave arrival times in observations and calculations:
L ( m ) = Q , e , r (t Qobs
er
− t Qsyner ( m )) 2 → min . (2)
Minimization of the functional L(m) uses the gradient method. The proper gradient (Fréchet
derivatives of the model parameters) in the given formulation can be found analytically [26]:
∂t Qsynеr
F≡ . (3)
∂mk

Fig. 1. The model parametrization: — seismic sensors; —microseismic events; —perforations; the dashed
lines —interfaces of the strata.

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015


PROCESSING MICROSEISMIC MONITORING DATA, CONSIDERING SEISMIC ANISOTROPY 479

Fig. 2. Resultant location of hypocenters of microseismic events under assumption of (a) anisotropic and (b)
isotropic model: —seismic sensors; —actual location of hypocenters; —calculated location of hypocenters from
the inverse problem solution.

2. EXAMPLE OF SYNTHETIC DATA


For the purpose of modeling, the arrival times are calculated for the P- and S-waves of a few
microseismic events line up as shown by the black points in Fig. 2 (simulation of hydrofracture). The
observation system is composed of 12 three-component sensor installed in two vertical holes (spacing
of the sesnors is 50 m). The homogenous transversally isotropic model under analysis has a horizontal
axis of symmetry and the Thomsen parameters: VP 0 = 3.0 km/s, VS 0 = 2.0 km/s, ε = 0.2 , δ = 0.2 ,
γ = 0.2 , symmetry axis azimuth α = 30° .
For the chosen parameters and locations of hypocenters of mixcroseismic events and sesnors, the
times of arraivals of waves are calculated. After addition of ranodm noise (standard deviation 0.5 ms),
the calculated times are assumed as being observed.
Then, the inverse kinematic problem is solved, i.e. the minimization of the closure error functional
(2). First, the vector of the model parameters, m (refer to Fig. 1), is analyzed, with the correct set of the
anisotropic parameters, c1 = [VP 0 , VS 0 , ε , δ , γ , α ] . Then, the consideration is given to the situation
when the correct type of the anisotropy is known and its parameters are not. The unshaded circles in
Fig. 2a show the hypocenters of microseismic events as a result of the inverse problem solving. It is
apparent the calculated (recovered) and the actual locations of the hypocenters agree well. The
anisotropic parameters are recovered from the solution of the inverse problem in the same manner.
Later on, with the same set of the synthetic data, the analysis involves m with c1 = [VP , VS ] , i.e.,
this is the situation when the medium is falsely assumed isotropic. The unshaded circles in Fig. 2b
show the microseismic event hypocenters calculated after solving this formulation inverse problem.
As seen, the recovered hypocenters line up again, but the position and azimuth of this line differ from
the actual data; e.g. the azimuth deviation is 15°.
So, it is concluded that the microseismic data processing under assumption of isotropy (for an
anisotropic actual model) yields error results that can also be interpreted as the image of a
hydrofracture. Though, the position and azimuth of this hydrofracture growth are erroneous
either. This inversion was implemented for many models and the results were similar.
The same result was observed in a series of synthetic tests for varied values of anisotropic
parameters, orientations of symmetry axis of the transversally isotropic medium and other types of
anisotropy.

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015


480 YASKEVICH et al.

Fig. 3. The velocities of P- (1) and S- (2) waves by the acoustic logging data; the dashed lines show the stratified
velocity model for Bakken deposit [24].

3. ACTUAL DATA PROCESSING


Let us consider the processing of the real data obtained in microseismic monitoring of
hydrofracturing at the Bakken Formation; its geology is described in [27]. In the examined area, the
Bakken Formation is divided into three parts: the top portion is clayey shale; the middle is siltstone,
the bottom portion is clayey shale. Total thickness of the strata is 36 m. Oil production is carried out
from the middle Bakken horizon.
Figure 3 shows the acoustic logging data obtained in a well located closely to the hydrofracturing
place. The dashed line shows the interpretation of the actual data as a 13-strata isotropic velocity model
later on used by a service company to process the microseismic monitoring data. The model contains
five clear layers (see Fig. 3): layer 1—coal measures; layers 2, 3 and 4—top, middle and bottom
Bakken; layer 5—Devonian measures. According to the core data, layer 4 rocks possess pronounced
anisotropy of properties.
The multistage HFR was carried out in a few isolated intervals uniformly spaced along the entire
length of a horizontal well.
The monitoring involved three-component sensors installed in two subvertical holes: 12 sensors
were placed in one borehole, spaced at 12.7 m, at the total array length of 160 (hereinafter, this
borehole is hole 1); 17 sensors were placed in the other borehole, spaced at 12.7 m, with the total
array length of 200 m (hereinafter, hole 2). The recording digitalization was 0.25 s. The both arrays,
at their bottom, cover the fluid injection depth. HFR was implemented in an open hole, thus, there
were no perforation shots (which gave no opportunity for the pre-calibration of the velocity model).
All in all, 900 microseismic events were recorded, out of 104 good quality seismograms were
obtained. An example of a seismogram is given in Fig. 4. In each seismograms, the first 14 traces are
for hole 1, the last 17 traces are for hole 2. Various components of the recording and the wave arrival
times are shown. The errors of the selected arrival times are high, which is identified with the mean
square discrepancy of 3.6 ms.
Furthermore, splitting of S-waves is systematically observed in the actual data, which is a reliable
identification of the seismic anisotropy. To confirm the splitting, Fig. 5 shows the travel distance graphs
(trajectories of particle movement in sensors) plotted by the three-component records in the S-wave
tracking window. The graphs posses a clear “cross” shape, which confirms the existence of the two
shear waves (S1 and S2) with the linear polarization orthogonally one the other.
During the research, the microseismic monitoring data were additionally processed: seismograms
for which waves were observed in the both holes were chosen, and the arrival times of P, S1 and S2
were determined. The S-wave splitting was observed for 70 out of 104 seismograms. The total amount
of the taken arrival times of the waves was 4607, out which there were 1172 arrival times of P-wave,
2585 arrival times of the quick shear wave and 850 arrival times of the slow shear wave.

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015


PROCESSING MICROSEISMIC MONITORING DATA, CONSIDERING SEISMIC ANISOTROPY 481

Fig. 4. Microseismic event seismogram, components: (a) X; (b) Y; (c) Z; “+” marks the actual arrival times of
the P- and S-waves; “×” marks the synthetic arrival times calculated for the isotropic model.

Fig. 5. Travel distance graphs for the P-wave—trajectories of the particle train in the wave window [24]: (a) hole 1;
(b) hole 2.

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015


482 YASKEVICH et al.

Fig. 6. Exposure of the five-layer model to the microseismic event rays. The points on the spheres—directions of rays of
all events in each layer for the Bakken Formation [24].

Fig. 7. Resultant location of hypocenters of synchronous seismic events: (a) plan view; (b) north size view and (c)
east side view.

Fig. 8. The observed arrival times of the waves (explanation is in the text).

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015


PROCESSING MICROSEISMIC MONITORING DATA, CONSIDERING SEISMIC ANISOTROPY 483

The preliminary analysis of which type of anisotropy is to be recovered based on the available set f
data involved examination of the geometry of the medium exposure to rays of the microseismic
events. In the isotropic mode, the ray scheme is calculated for the events. Directions of the rays in
each layer are plotted as red color point on the unit spheres (Fig. 6). ). The bulk of events took places
in the top later (layer 1) and this layer is, therefore, best exposed from different angles. Exposure of
layers 2 and 3 is also good, and these layers can be assumed anisotropic (orthorhombic or triclinic
anisotropy). The two bottom show limited exposure along the azimuths, which makes it impossible to
determine the azimuthal anisotropy.
So, in the minimization of the time closure error functional (2) for determining parameters of m,
two lower layers are transversely isotropic with the vertical axis of symmetry (VTI) and three upper
layers have triclinic type anisotropy (TRI).
The inverse problem solution results are shown in Figs. 7 and 8. Figure 7 shows the locations of
microseismic event hypocenters. It is seen that as a result of using the anisotropy velocity model, the
cloud of microseismic event is narrower, i.e., it resembles a hydrofracture. Furthermore, the cloud is
re-located (approximately by 40 m towards the observation holes), which conforms with the data of
the synthetic tests.
The data processing, including anisotropy, yielded 6 times smaller mean square closure error
between the actual and calculated arrival times of the waves: 0.54 ms for the anisotropic model and
3.6 ms for the isotropic model. This reduction is an evidence of the higher quality of the data
processing. The selected wave arrival times for the anisotropy velocity model are exemplified in
Fig. 8, where the unshaded and shaded markers show the observed and calculated arrival times,
respectively.
The inversion using the anisotropic stratified medium recovered the following density normalized
stiffness tensors, km2/s2:
 26.57 12.14 10.00 0.92 − 0.27 − 0.49 
 
 27.80 10.66 0.89 − 0.86 − 0,53 
 22.11 0.94 − 0.54 − 0.10 
c[1] =  ,
 7.32 − 0.14 − 0.12 
 symmetrical 6.95 0.20 
 
 6.49 
 
 15.78 1.15 1.82 0.09 − 0.04 − 0.08 
 
 15.52 1.98 0.12 − 0.05 0.08 
 9.42 − 0.04 0.00 − 0.14 
c[ 2] = ,
 3.88 0.35 − 0.21 
 symmetrical 4.06 0.03 
 
 6.83 
 
 19.81 8.62 9.00 − 2.37 − 1.44 0 
 
 25.79 9.09 0.57 − 0.99 − 0.89 
 20.68 2.10 0.43 − 0.08 
c[3] = 
 3.88 0.35 − 0.21 
 symmetrical 4.06 0.03 
 
 6.49 
 
and the parameters of the two bottom layers with VTI anisotropy:

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015


484 YASKEVICH et al.

c[ 4 ] : VP 0 = 2.81 km/s, VS 0 = 1.97 km/s, ε = 0.27 , δ = 0.19 , γ = 0.35 ,


c[ 5 ] : VP 0 = 4.17 km/s, VS 0 = 2.38 km/s, ε = 0.09 , δ = −0.16 , γ = 0.13 .
Then, the inverse problem was solved for the case with the simplified symmetry of the three top
layers, and they were assumed as orthorhombic (ORT) or transversely isotropic with the vertical axis
of symmetry (VTI). The recovered parameters of the layers are given in the table. Based on these
results, it is possible to come to the conclusion that layers 2, 3 and 4 of the Bakken Formation are
heavily anisotropic. This agrees with the core data from layer 4.
It is worth noticing that the choice of the anisotropy type for the top three layers has no significant
influence on the location of the cloud of microseismic event hypocenters in Fig. 7.
Estimate of the solution stability of the inverse problem used the condition number of the matrix of
the Fréchet derivatives F of the travel times based on the model parameters (see Fig. 3). The
condition number logarithm is 3.08, which, based on experience of numerical experiments, enables
stable solving of the inverse problem on location of microseismic events and determining the type of
the velocity model anisotropy.
The problem solution stabuiluty is also confirmed by the result of analysis of the problem
regulaization necessity. Let us take, for example, Tikhonov’s regularization that suggest solving the
inverse problem on determination of the parameters of the model m λ by minimizing the modified
closure error functional:
Fm λ − d 22 + λ2 m λ 22 ,
where λ —regularization parameter; ⋅ 22 —Euclide norm of the vector, i.e. v = k ϑk2 .
2
2
The L-curve method [28] enables choosing the optimal regularization parameter by plotting a
curve with the coordinates ( log10 Fm λ − d , log10 m λ ) for different λ consistent with m λ . In the
problems that need regularization, this curve will have a characteristic L-shape incurved towards the
coordinate origin. The curve L ( λ ) plotted for the problem under analysis is shown in Fig. 9. The
curve is barreled, i.e., the problem is solved stably without the regularization.
4. DISCUSSION OF THE RESULTS
The obtained results show that the data on travel times of the direct waves of microseismic
events allow sufficient information for location of hypocenter of the events and parameters of the
anisotropic velocity model. This method of concurrent location of microseismic events and
refinement of velocity model differs from the standard borehole monitoring data processing when
the velocity model calibration only uses records of perforation events with the known coordinates
of the source. The proposed approach is particularly important in hydrofracturing in an open hole
when there are no data on perforation shots or other events with the known coordinates.
Recovered parameters of the anisotropic models
ORT/VTI models (closure error 0.64 ms)
VP0 VS0 ε1 ε2 δ1 δ2 δ3 γ1 γ2 α
4.69 2.66 0.13 0.11 0.06 0.11 – 0.01 0.3 0.04 149°
3.11 2.12 0.34 0.31 – 0.04 0.07 0.01 0.36 0.48 145°
4.54 2.69 – 0.05 0.04 0.25 0.20 0.11 – 0.01 0.01 141°
VTI/VTI models (closure error 0.94 ms)
VP0 VS0 ε δ γ
4.56 2.72 0.10 0.07 0.02
3.16 2.01 0.37 – 0.01 0.33
4.63 2.83 0.01 0.17 – 0.12

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015


PROCESSING MICROSEISMIC MONITORING DATA, CONSIDERING SEISMIC ANISOTROPY 485

Fig. 9. L-curve for choosing Tikhonov’s regularization parameter.

Even in case perforations are present, the data on microseismic events allow better seismic sounding
coverage, which improves reliability of a velocity model, especially in determining parameters of
azimuthally anisotropic media. It is shown in the article that with the better seismic sounding coverage,
elastic moduli of anisotropic layers can be recovered up to the triclinic symmetry. The authors have
found out the determination of the parameters of such media requires a network of the sufficiently long
observation holes to intersect depthwise the interval where seismic events originate.
The made examples illustrate better location of hypocenters of microseismic events in their
processing with account for anisotropy. This is connected with both the more accurate velocity model
and the use of new kinematic information of the arrival times of two shear waves. In particular, the
hypocenters have been located for microseismic events for which P-wave was unavailable but the
arrival times of two shear waves were well traced.
Owing to the anisotropy velocity model, the closure error between the observed and calculated
travel times of the waves had been greatly lessened as compared with the isotropic velocity model.
CONCLUSIONS
The authors have offered the microseismic monitoring data processing method for the concurrent
location of hypocenters of microseismic events and determination of parameters of the azimuthally
anisotropic velocity model of a rock mass.
The method enhances accuracy of location of microseismic event hypocenters in an anisotropic
medium owing to the use of arrival tomes of two shear waves. It has experimentally been shown that
neglect of anisotropy in monitoring of hydraulic fracturing results in shifted location of the fracture as
against its true position and in erroneous estimate of the fracture radius.
The authors confirm the proposed method efficiency in the data of hydrofracturing monitoring at
the Bakken Formation, USA. The estimates anisotropic parameters agree with the core data from the
lower lying productive stratum.
ACKNOWLEDGMENTS
This work was supported by the RF Ministry of Education and Sciences, project
no. RFMEF160414X0047.
REFERENCES
1. Dobroskok, A.A. and Linkov, A.M., Modeling of Fluid Flow, Stress State and Seismicity Induced in Rock
by an Instant Pressure Drop in a Hydrofracture, J. Min. Sci., 2011, vol. 47, no. 1, pp. 10–19.
2. Rutledge, J. and Soma, N., Using Reflected Phases to Improve Depth Resolution of Microseismic Source
Locations from Single-Well Observations, Proc. Unconventional Resources Technology Conf., 2013.
3. Hayles, K., Horine, R.L., Checkles, S., and Blangy, J.P., Comparison of Microseismic Results from the
Bakken Formation Processed by Three Different Companies, SEG Technical Program Expanded
Abstracts, 2011.
4. Aleksandrov, S.I., Mishin, V.A. and Burov, D.I., Surface Microseismic Monitoring of Hydrofracturing:
Quality Control and Prospects, Ekspoz. Neft. Gaz., 2014, no. 2.

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015


486 YASKEVICH et al.

5. Chambers, K., Kendall, M., Brandsberg-Dahl, S., and Rueda, J., Testing the Ability of Surface Arrays to
Monitor Microseismic Activity, Geophysical Prospecting, 2010, vol. 58.
6. Shmakov, F.D., Procedure for Processing and Interpreting Data of Microseismic Activity Monitoring in
Hydrocarbon Reservoirs, Tekhnol. Seismorazv., 2012, no. 3.
7. Eisne, L., Duncan, P., Heig, W.M., and Keller, W.R., Uncertainties in Passive Seismic Monitoring, The
Leading Edge, 2009, vol. 28.
8. Zhang, H., Sarkar, S., Toksoz, M.N., Kuleli, H.S., and Al-Kindy, F., Passive Seismic Tomography Using
Induced Seismicity at a Petroleum Field in Oman, Geophysics, 2009, vol. 74, no. 6.
9. Zimmer, U., Bland, H., Du, J., Warpinski, N., Sen, V., and Wolfe, J., Accuracy of Microseismic Event
Locations Recorded with Single and Distributed Downhole Sensor Arrays, SEG Technical Program
Expanded Abstracts, 2009.
10. Jansky, J., Plicka, V., and Eisner, L., Feasibility of Joint 1D Velocity Model and Event Location Inversion by
the Neighborhood Algorithm, Geophysical Prospecting, 2010, vol. 58.
11. Abel, J. S., Coffin, S., Hur, Y., and Taylor, S., An Analytic Model for Microseismic Event Location
Estimate Accuracy, First Break, 2011, vol. 29.
12. Usher, P.J., Angus, D.A., and Verdon, J.P. Influence of a Velocity Model and Source Frequency on
Microseismic Waveforms, Some Bakken Microseismic Implications for Microseismic Locations,
Geophysical Prospecting, 2013, vol. 61.
13. Vernik, L. and Nur, A., Ultrasonic Velocity and Anisotropy of Hydrocarbon Source Rocks, Geophysics,
1992, vol. 57.
14. Vernik, L., and Liu, X., Velocity Anisotropy in Shales: A Petrophysical Study, Geophysics, 1997, vol. 62.
15. Tsvankin, I. and Grechka, V., Seismology of Azimuthally Anisotropic Media and Seismic Fracture
Characterization, SEG, Geophysical References, 2011, Series no. 17.
16. Maxwell, S., Shemeta, J., and House, N., Integrated Anisotropic Velocity Modeling Using Perforation
Shots, Passive Seismic and VSP Data, CSPG-CSEG-CWLS Convention, 2006.
17. Verdon, J. P., Kendall, J.–M., and Wustefeld, A., Imaging Fractures and Sedimentary Fabrics Using Shear
Wave Splitting Measurements Made on Passive Seismic Data, Geophysical J. Int., 2009, vol. 179.
18. Verdon, J.P. and Kendall, J.–M., Detection of Multiple Fracture Sets Using Observations of Shear-Wave
Splitting in Microseismic Data, Geophysical Prospecting, 2011, vol. 59.
19. Grechka, V. and Duchkov, A., Narrow-Angle Representations of the Phase and Group Velocities and
Their Applications in Anisotropic Velocity Model Building for Microseismic Monitoring, Geophysics, 2011,
vol. 76, no. 6.
20. Grechka, V., Singh, P., and Das, I., Estimation of Effective Anisotropy Simultaneously with Locations of
Microseismic Events, Geophysics, 2011, vol. 76, no. 6.
21. Li, J., Rodi, W., Toksoz, M.N., and Zhang, H., Microseismicity Location and Simultaneous Anisotropic
Tomography with Differential Traveltimes and Differential Back Azimuths, SEG Technical Program
Expanded Abstracts, 2012, pp. 1–5.
22. Li, J., Toksoz, N., Li, C., Morton, S., Dohmen, T., and Katahara, K., Locating Bakken Microseismic
Events with Simultaneous Anisotropic Tomography and Extended Double-Difference Method, SEG
Technical Program Expanded Abstracts, 2013.
23. Grechka, V. and Yaskevich, S., Inversion of Microseismic Data for Triclinic Velocity Models,
Geophysical Prospecting, 2013, vol. 61.
24. Grechka, V. and Yaskevich, S., Azimuthal Anisotropy in Microseismic Monitoring: A Bakken Case
Study, Geophysics, 2014, vol. 79, no. 1.
25. Obolentseva, I.R. and Grechka, V.Yu., Luchevoi metod v anizotropnoi srede (algoritmy,
programmy) (Ray-Path Method in an Anisotropic Medium (Algorithms, Programs), Novosibirsk: IGiG SO
AN SSSR, 1989.
26. Сerveny, V., Seismic Ray Theory: Cambridge University Press, 2001.
27. Meissner, F.F., Petroleum Geology of the Bakken Formation Williston Basin, North Dakota, and Montana,
Proc. Montana Geological Society 24th Annual Conference, 1991.
28. Hansen, P.C., Pereyra, V., and Scherer, G., Least Squares Data Fitting with Applications, Johns Hopkins
University Press, 2012.

JOURNAL OF MINING SCIENCE Vol. 51 No. 3 2015

You might also like