Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Geothermics 71 (2018) 225–233

Contents lists available at ScienceDirect

Geothermics
journal homepage: www.elsevier.com/locate/geothermics

Seismic velocity structure and characteristics of induced seismicity at the MARK


Geysers Geothermal Field, eastern California

Guoqing Lina, , Bateer Wub
a
Department of Marine Geosciences, Rosenstiel School of Marine and Atmospheric Science, University of Miami, Miami, FL 33149, USA
b
Institute of Disaster Prevention, Xueyuan Road, Village Yanjiao, Sanhe, Hebei Province 065201, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: We present a newly developed three-dimensional seismic velocity model, high-precision earthquake relocations
The Geysers and focal mechanisms near the Geysers Geothermal Field in eastern California using seismic data recorded by the
Induced seismicity Northern California Earthquake Data Center from 1984 to 2015. The velocity model generally agrees with those
Crustal structure in previous studies with the Vp model mainly corresponding to rock composition and the Vp/Vs model more
correlated with fluid content. The dominating low Vp/Vs anomalies observed from 0 to 4 km depth below sea
level is a reflection of the geothermal reservoir. The waveform cross-correlation relocated seismicity shows both
spatial and temporal correlations with the geothermal operations in the study area, indicating that they are
induced seismicity. Although the focal solutions are dominated by normal and strike-slip regimes throughout our
entire study time period, there has been an increase in reverse faulting since 2008, which may be caused by the
thermal contraction associated with the Northwest Geysers Enhanced Geothermal System project between 2008
and 2012. Our study provides a groundwork for future seismological studies in The Geysers.

1. Introduction the surface geology and the geologic and hydrologic features of the
reservoir. Although some studies found no consistent correlation be-
The Geysers Geothermal Field (GGF) in eastern California is situated tween injection and seismicity, it has been widely accepted that there is
approximately 110 km north of San Francisco and is the largest geo- a correlation between the two parameters (Majer and Mcevilly, 1979;
thermal field in the world. Although intensively exploited, it still pro- Denlinger and Bufe, 1980; Bufe and Shearer, 1981; Allis, 1982;
vides about 6% of California's electrical power. The Geysers is also one Eberhart-Phillips and Oppenheimer, 1984; O’Connell and Johnson,
of the most seismically active regions in California and is dominated by 1991; Stark, 1992; Romero et al., 1995; Alwyn, 1996; Trugman et al.,
shallow induced seismicity. It is located on the eastern side of the 2016). The stress state study by Boyle and Zoback (2014) revealed a
Maacama Fault Zone and the Healdsburg Fault, bounded by the right- normal/strike-slip faulting regime within and below the reservoir and
lateral strike-slip Mercuryville Fault and Collayomi Fault, and sur- also suggested that the geothermal operations have no significant effect
rounded by several Mountains, such as the Geyser Peak, the Cobb Mtn., on the local stress field.
the Boggs Mtn., the Seigler Mtn., and the Mt. Hannah (Fig. 1). Since the In this paper, we aim to investigate the structural heterogeneities
geothermal operations commenced in the 1960s, numerous studies and examine the spatial and temporal variations and characteristics of
have been conducted to investigate their effects on hydrologic and induced seismicity in The Geysers Geothermal Field by using a newly
tectonic features in the area. The tectonic setting and structure in The developed high-resolution three-dimensional (3-D) Vp and Vp/Vs model,
Geysers have been reviewed by previous researchers (e.g., McLaughlin, precise earthquake relocations, and focal mechanisms. This study is also
1981; Thompson, 1989, 1992). The spatial/temporal variations in motivated by a recent work on remotely triggered earthquakes by
seismic velocities have been intensively investigated (e.g., Majer and Zhang et al. (2017). They found that although the adjacent fault areas
Mcevilly, 1979; O’Connell and Johnson, 1991; Zucca et al., 1994; are responsive, the two hot water systems themselves, the Coso and the
Romero et al., 1995; Alwyn, 1996; Julian et al., 1996; Kirkpatrick et al., Salton Sea Geothermal Fields, are insusceptible to large distant earth-
1997; Boitnott and Kirkpatrick, 1997; Foulger et al., 1997; Gunasekera quakes. The Geysers Geothermal Field is an under-pressurized, vapor-
et al., 2003; Gritto et al., 2013; Gritto and Jarpe, 2014). The majority of dominated field, which may have different response to remote earth-
these studies show that the shallow velocity structure correlates with quake triggering. Our results in this study will provide essential


Corresponding author.
E-mail address: glin@rsmas.miami.edu (G. Lin).

http://dx.doi.org/10.1016/j.geothermics.2017.10.003
Received 23 March 2017; Received in revised form 29 September 2017; Accepted 8 October 2017
Available online 17 October 2017
0375-6505/ © 2017 Elsevier Ltd. All rights reserved.
G. Lin, B. Wu Geothermics 71 (2018) 225–233

Fig. 2. Events (dots) and inversion grid nodes (green squares) used in the tomographic
inversion. Gray dots stand for earthquakes that are only recorded by the EGS stations, red
Fig. 1. Major geological structures in the study area, including the surface fault lines ones for those recorded by the NCSN stations, and two big blue ones for the three ex-
(black lines) and several mountain ranges (pink squares). Gray dots represent ∼246,000 plosion events (with two at the same location). The Cartesian coordinate is rotated 45°
earthquakes between 1984 and 2015 recorded by the NCSN (black triangles). Blue and counterclockwise with respect to latitude and longitude, with the X axis pointing to the
white triangles stand for the EGS and the temporary stations, respectively. The back- southwest and Y axis to the northwest direction. The origin is located at (38.8°, −122.8°)
ground is the topography base map from the U. S. Geological Survey. The red box in the (shown by the yellow star). Dark blue squares denote the same mountains in Fig. 1. The
inset map shows the location of our study area in California. (For interpretation of the two yellow straight lines are the profiles for the cross-sectional views in the following
references to color in this figure legend, the reader is referred to the web version of the figures. The inset figure shows the 1-D initial Vp model for the tomographic inversion,
article.) which is used for the NCSN daily operation in the Geysers area. (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of the
article.)
groundwork for the future fine-scale remote triggering analyses and
other seismological studies in The Geysers.
temporary seismic stations (white triangles in Fig. 1) and collected by
previous investigators (Meador et al., 1985), were also included in the
2. Data tomographic inversion to constrain the shallow crustal structure and
absolute hypocenter locations.
We obtained all the seismic data used in this study from the
Northern California Earthquake Data Center (NCEDC, 2014). These 3. Results
data are mainly recorded by the Northern California Seismic Network
(NCSN) stations (black triangles in Fig. 1), consisting of ∼246,000 local 3.1. 3-D seismic velocity model
earthquakes (gray dots in Fig. 1) from 1984 to 2015 with 3.7 million
compressional (P) and 0.18 million shear (S) wave first-arrivals and all The simul2000 algorithm is one of the most widely applied tomo-
their waveform data. The majority of these events are confined within graphic programs, which solves for 3-D Vp, Vp/Vs models and earth-
the steam-production zone of the Geysers Geothermal Field, bounded quake locations simultaneously (Thurber, 1983, 1993; Eberhart-
by the Mercuryville Fault to the southwestern side and the Collayomi Phillips, 1990; Thurber and Eberhart-Phillips, 1999). It combines
Fault to the northeastern side. parameter separation and damped least squares for model perturba-
For the tomographic inversion, due to the small number of S-picks tions. The original simul2000 algorithm allows a constant minimum
available in the NCSN catalog, we also take advantage of the seismic earthquake depth, i.e., none of the earthquake locations can be shal-
record from the Enhanced Geothermal Systems (EGS) stations (blue lower than the given minimum depth. In this study, we apply a mod-
triangles in Fig. 1), operated by the US Department of Energy Geo- ified version of the simul2000 that replaces the constant minimum
thermal Technologies Program. The EGS data include ∼444,000 depth with the specific topographic information to avoid artificial ef-
earthquakes since April 2003 with over 5 million P- and 2 million S- fects (Lin, 2015).
picks. Note that the EGS catalog events are only used in the velocity The horizontal grid spacing in our 3-D velocity model is 2 km (ro-
tomography, not in the relocation or focal mechanism inversions. tated squares in Fig. 2). The vertical node intervals range between 1 and
For the selection of the master events used in the tomographic in- 4 km from −1 to 20 km depth. Zero depth corresponds to mean sea
version, we apply different requirements for the NCSN and EGS net- level. Note that in this paper all depths are relative to mean sea level,
work catalogs. For the EGS catalog events, we require more than 8 P unless otherwise stated. A one-dimensional (1-D) Vp model (inset in
picks and 4 S-P times per event. The resulting 1937 events (gray dots in Fig. 2), which is used for the NCSN daily operation in the Geysers area,
Fig. 2) consist of 26,975 P and 12,243 S picks. Due to the very few is applied as the initial model for the 3-D tomographic inversion. The
number of S-picks in the NCSN catalog, we require at least 8 P picks and starting value for the Vp/Vs inversion is a constant ratio of 1.732.
no constraints on the number of S picks for each event. The 2038 events In order to investigate the model recovery, we performed a check-
(red dots in Fig. 2) comprise 34,617 P and 333 S picks. The inclusion of erboard resolution test similar to previous studies (e.g., Thurber et al.,
the events outside of the GGF are helpful to improve deep velocity 2009; Lin et al., 2010, 2014; Lin, 2013, 2015). We created a synthetic
structure resolution. These total 3975 events and their corresponding velocity model based on the 1-D starting model with ± 5% Vp and ∓5%
picks are used in our final velocity model inversion. P-wave arrival Vp/Vs perturbations (i.e., opposite signs) that alternate at different
times for 3 active-source data (big blue dots in Fig. 2), recorded by depths and across two grid nodes. Synthetic times were computed

226
G. Lin, B. Wu Geothermics 71 (2018) 225–233

Fig. 3. Checkerboard resolution test for the Vp model. The synthetic times are computed from the 1-D starting velocity model with ± 5% velocity anomalies across two grid nodes.
(a1–a4) Map views of the synthetic model. (b1–b4) Map views of the inverted model. (c1–c2) Cross sections of the synthetic model along profiles 1-1′ and A-A′ in (a1) and Fig. 2. (d1–d2)
Cross sections of the inverted model along the same profiles. The green contours enclose the well-resolved area with the diagonal element of the resolution matrix greater than 0.3. Dotted
curves at top of cross sections illustrate the local topography. MF is short for the Mercuryville Fault and CF for the Collayomi Fault.

through this model for the same event hypocenters, station locations, The tomographic inversion is convergent after 7 iterations. For the
and ray geometry as those in the real data. Fig. 3 shows map view and master events, the root-mean-square (RMS) of the arrival-time residuals
cross-sectional comparisons between the synthetic and inverted Vp is reduced from 0.78 to 0.29 s for P arrivals and from 0.47 to 0.37 s for S
models in the test. The green contours enclose the well-resolved area arrivals after the inversion. For the explosion data with only P arrivals,
with the diagonal element of the resolution matrix greater than 0.3, the RMS drops from 0.34 to 0.19 s, indicating that the shallow crustal
where 1.0 represents the best resolution and 0.0 not resolved at all. At structure near the geothermal field, where the explosion data are
−1 km depth (i.e., 1 km above sea level), the Vp model is poorly re- available, has been greatly improved.
solved, with only the central part of the steam-production zone re- Fig. 5 shows the map views and cross-sections of our final P-wave
covered, due to the lack of seismicity at this shallow depth. At sea level, velocity model. The white contours enclose the well-resolved regions
the model resolution is significantly improved within the geothermal where the diagonal element of the resolution matrix is greater than 0.3.
field. The next two layers are well recovered both within the geo- At −1 and 0 km depths, the geothermal field is dominated by relatively
thermal field and its surrounding area, owing to the inclusion of both high velocity anomalies than the surrounding area. In contrast, the two
the NCSN and EGS catalog data. The Vp model along the two cross- deeper layers (2 and 4 km depths) show low velocity structures within
section profiles is almost fully-recovered with slight smearing near the the geothermal field. The final Vp/Vs model is shown in Fig. 6. The first
edge. The Vp/Vs model (Fig. 4) is not resolved as well as the Vp model, layer has limited resolution due to the lack of seismicity and displays
with only the structure within the geothermal field recovered because high Vp/Vs values up to 1.82. The GGF in the next two layers shows low
the S arrivals are mostly from the EGS stations, which are located Vp/Vs ratios with relatively high values near the boundary. The model is
within the geothermal field. only resolved in the central area of the steam-production zone at 4 km

227
G. Lin, B. Wu Geothermics 71 (2018) 225–233

Fig. 4. Checkerboard resolution test for the Vp/Vs model. The synthetic times are computed from the constant Vp/Vs with ± 5% anomalies across two grid nodes. (a1–a4) Map views of the
synthetic model. (b1–b4) Map views of the inverted model. (c1–c2) Cross sections of the synthetic model along the profiles shown in (a1). (d1–d2) Cross sections of the inverted model
along the same profiles. Symbols are the same as those in Fig. 3.

depth. location uncertainties estimated by a bootstrap approach (Efron and


Gong, 1983; Efron and Tibshirani, 1991) have a median of 7 m for
3.2. Earthquake relocation horizontal and 10 m for vertical.
Fig. 7 shows the location comparison between the initial NCSN
The final velocity model is used to relocate all the ∼246,000 NCSN catalog (a–c) and the waveform cross-correlation relocation of this
recorded earthquakes in the study area. The simul2000 algorithm es- study (d–f). Note that the NCSN catalog is based on 1-D velocity models,
timates the absolute location uncertainties in the horizontal and ver- therefore zero depth refers to the average station elevation used in the
tical directions. For all the 3-D relocations, the median of the location location process (about 800 m above sea level) instead of mean sea
uncertainties is 98 m for horizontal and 137 m for vertical. level in this study. The difference in absolute locations between the two
Relative location accuracy of the 3-D relocated earthquakes is im- catalogs displays the importance of 3-D velocity structure in earthquake
proved by applying similar-event cluster analysis and differential-time relocation studies. The use of waveform cross-correlation data in dif-
location methods based on waveform cross-correlation data, as pre- ferential time relocation process enables the dramatic sharpening in
sented in Lin et al. (2007). The cluster analysis approach is to identify seismicity distribution.
groups of earthquakes that are correlated with one other. Given our
criteria for the waveform cross-correlation processing, which are the 3.3. Earthquake focal mechanism
same as those used in a previous study (Lin, 2015), over 64% of all the
3-D relocated earthquakes fall within similar-event clusters. The Based on the waveform cross-correlation relocations, we compute
147,663 earthquakes in 1588 similar-event clusters were then relocated focal mechanisms from P-wave first motion polarity observations by
using the waveform cross-correlation times. The relative earthquake applying the HASH program (Hardebeck and Shearer, 2002), assuming

228
G. Lin, B. Wu Geothermics 71 (2018) 225–233

Fig. 5. Map views and cross sections of the final P-wave velocity model. (a–d) Map views of the P-wave velocity model at each layer depth where the inversion nodes are located. The
white contours enclose the well-resolved area with the diagonal element of the resolution matrix greater than 0.3. Black lines denote the surface traces of mapped faults. Red straight lines
in (d) are the profiles for the cross sectional views. Inset in (a) shows the Vp versus Vs (derived from dividing Vp by Vp/Vs) values at the well-resolved inversion nodes. (e–f) Cross sections
of the Vp model along the two profiles in (d), including the waveform cross-correlation relocated seismicity (gray dots) within ± 0.5 km distance of the profile lines. Dotted curves at top
of cross sections illustrate the local topography. Abbreviations are MF for the Mercuryville Fault and CF for the Collayomi Fault. (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of the article.)

double-couple fault plane solutions. Over 2.2 million first-motion po- normal, [−0.33, 0.33] for strike-slip, and (0.33, 1] for reverse), for
larity records are obtained from the vertical component of each seismic each solution from the rakes of the two nodal planes (Shearer et al.,
station and were used in the computation. We assign four types of 2006; Lin and Shearer, 2009; Lin and Okubo, 2016). The mechanisms in
qualities (A–D) to the solved 98,453 focal mechanisms based on the our study area are dominated by normal faulting (51%) and strike-slip
criteria in the HASH algorithm and consider the quality A and B me- (38%) and much fewer reverse (11%) focal types. Fig. 8 shows map
chanisms as good-quality solutions, which are ∼15% of the total me- view and depth distribution of good quality focal mechanisms plotted at
chanisms. We then compute the scalar faulting type ([−1, −0.33) for their cross-correlation relocations.

Fig. 6. Map views and cross sections of the final Vp/Vs model. (a–d) Map views of the Vp/Vs model at each layer depth where the inversion nodes are located. The white contours enclose
the well-resolved area with the diagonal element of the resolution matrix greater than 0.3. Black lines denote the surface traces of mapped faults. Red straight lines in (d) are the profiles
for the cross sectional views. (e–f) Cross sections of the Vp/Vs model along the two profiles in (d), including the waveform cross-correlation relocated seismicity (gray dots) within ±
0.5 km distance of the profile lines. Dotted curves at top of cross sections illustrate the local topography. Abbreviations are MF for the Mercuryville Fault and CF for the Collayomi Fault.
(For interpretation of the references to color in this figure legend, the reader is referred to the web version of the article.)

229
G. Lin, B. Wu Geothermics 71 (2018) 225–233

Fig. 7. Earthquake location comparison between the initial NCSN catalog location (a–c) and the waveform cross-correlation relocation in this study (d–f). Events in (a) and (d) are colored
by depth. Red triangles in (d) are the locations of the operational wells obtained from the California Oil, Gas, and Geothermal Resources. The depth distributions show seismicity
within ± 0.5 km distance of the two profile lines (straight pink lines in a and d). Orange dotted curves at top of the cross-sections illustrate the local topography. MF is short for the
Mercuryville Fault and CF for the Collayomi Fault. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of the article.)

4. Discussion at 8 km and deeper. Some teleseismic and local studies reveal low P-
wave velocity anomalies under the steam-production zone from the
In this study, we perform a series of seismic analyses in order to crust to the upper mantle (e.g., Iyer et al., 1981; Oppenheimer and
investigate the effect of the geothermal operational activities on the Herkenhoff, 1981; Zucca et al., 1994). However, some studies found
geological and seismological features in the Geysers area. We first de- that P-wave velocity in The Geysers production zone is faster than the
velop a fine-scale 3-D seismic velocity model using both P and S wave surrounding area (e.g., Eberhart-Phillips, 1986; Majer and Mcevilly,
arrivals obtained from the NCEDC. The major contribution of this work 1979). The Vp model in our study is well resolved from the surface to
is a high-precision earthquake catalog with over 147,000 events re- 4 km depth due to the dominating shallow earthquakes used in the
located based on the 3-D velocity model and waveform cross-correla- tomographic inversion. According to the laboratory measurements for
tion data and 67% of them have focal mechanisms derived from P wave The Geysers core material, Vp and Vs vary in opposite ways with degree
first motion polarity observations. In addition to the enormous volume of saturation (Kirkpatrick et al., 1997). However, the well-resolved Vp
of seismic data (both inside and outside of the GGF) recorded by the and Vs values in our study show a positive correlation (inset in Fig. 5a),
NCSN stations, the EGS data and the controlled sources used in the indicating that they do not significantly depend on the saturation. In-
tomographic inversion greatly helped to improve the resolution of the stead, our Vp model greatly agrees with the geological structure in the
Vp/Vs model and the shallow Vp structure, therefore the overall ro- area (McLaughlin, 1981), suggesting that lithology plays a more im-
bustness of our results. portant role in determining Vp value in our study area. Our velocity
The seismic velocity structure in The Geysers has been investigated model is most consistent with the one by Romero et al. (1995) for the
by numerous studies. The 1-D initial velocity model used in this study is Northwest Geysers region, whose surface geology reflects the general
similar to the layered model by Eberhart-Phillips and Oppenheimer geology of our study region. We interpret the high velocity anomalies at
(1984) that shows Vp varying from 4.4 km/s at the surface to 5.9 km/s shallow depth (1 km above and at sea level) as a reflection of the

230
G. Lin, B. Wu Geothermics 71 (2018) 225–233

Fig. 8. Map view (a) and cross-sections (b and c) of good quality (i.e., quality A and B) focal mechanisms based on the waveform cross-correlation relocation, colored by their focal type
with red for normal, green for strike-slip, and blue for reverse. The depth profiles for (b) and (c) are shown by the yellow straight lines in Fig. 2, same as those for the velocity model and
earthquake relocation presentations. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of the article.)

Franciscan graywacke and greenstone and the low-velocity anomalies earthquake clusters, in general seismicity seems to occur where the
at 2 km depth as a correlation with Franciscan melange. wells are located. In addition, the seismicity at depth is more correlated
In contrast to the Vp model, the Vp/Vs model is more sensitive to with the resolved Vp/Vs anomalies (Fig. 6e and f) than the absolute Vp
temperature and fluid content than Vp and Vs alone. The common fea- variations (Fig. 5e and f), suggesting that the seismicity is induced. We
ture of the existing Vp/Vs models (e.g., Julian et al., 1996; Foulger et al., also acquired the well extraction and injection data from the DOGGR
1997; Kirkpatrick et al., 1997; Gritto et al., 2013; Gritto and Jarpe, and plot their annual rates as a function of time for our study time
2014) is the dominating low ratio (∼1.58), corresponding to the period (Fig. 9a). Between 1984 and 1995, the net production (i.e., ex-
exploited section of the geothermal reservoir, which is usually ex- traction-injection) rate showed similar pattern as the extraction because
plained by low pore pressure and dry conditions at depth. The low Vp/ the injection remained almost constant. Since 1995, the injection rate
Vs anomalies observed in our model between 0 and 4 km depth below has been fluctuating with a general increasing trend and the extraction
sea level (bsl) is consistent with these previous models. The high Vp/Vs rate has been nearly flat. Following the method of the maximum cur-
ratios at shallow depth may suggest a recharge region or a condensation vature by Wiemer and Baer (2000), we calculated the magnitude of
zone (Romero et al., 1995). The relocated seismicity is well-correlated completeness for the relocated events in our study as 1.4. In Fig. 9a, we
with the inverted low Vp/Vs anomalies (Fig. 6e and f). Bigger event plot the annual seismicity rate for events above magnitude 1.4, which is
clusters seem to occur at the lower boundary of the anomalous body correlated with the injection rate extremely well. The seismicity rate
(∼4 km depth), similar to the seismicity distribution observed by increases in 1997 and 2003 correspond to the pipeline constructions
O’Connell and Johnson (1991). Gritto et al. (2013) also observed the during those time periods, which was also observed in Trugman et al.
correlation of seismicity with higher Vp/Vs estimates. However, our (2016). The event spatio-temporal correlation with the well operations
resolved area is dominated by low Vp/Vs anomalies and does not show and their shallow depths (≤4 km bsl) further support that the majority
this pattern. of the seismicity is induced seismicity.
With more and more seismic stations and records available in the Our study area is dominated by normal (51%) and strike-slip (38%)
past few decades, the correlation between injection and seismicity has focal types, consistent with the extensional and strike-slip tectonic
been commonly observed in The Geysers (e.g., Majer and Mcevilly, setting in the area surrounding The Geysers. For events above 1 km
1979; Denlinger and Bufe, 1980; Bufe and Shearer, 1981; Allis, 1982; depth, 56% of the well-resolved focal solutions are strike-slip and 29%
Eberhart-Phillips and Oppenheimer, 1984; O’Connell and Johnson, are normal, indicating the shallow events are more affected by tectonic
1991; Stark, 1992; Romero et al., 1995; Alwyn, 1996), although the stress and deeper ones by operational activity. We also apply the focal
seismicity rate and shape vary over the years. The recent work by mechanism stress algorithm SATSI by Hardebeck and Michael (2006) to
Trugman et al. (2016) observed influence of fluid injection on the invert for the stress field in the study area. The N28°E orientation of the
seismicity rate, b value, earthquake magnitude, and event depth within inverted maximum horizontal compressional stress SH max is consistent
The Geysers. In order to investigate the spatial relationship between with previous focal-mechanism-based stress directions in the area (e.g.,
seismicity and injection, we plot the locations of the operational wells Oppenheimer, 1986; Boyle and Zoback, 2014; Heidbach et al., 2016). In
obtained from the California Oil, Gas, and Geothermal Resources order to investigate the temporal variations of the event focal types, if
(DOGGR) together with the relocated events in similar event clusters there are any, we plot the scalar focal solutions for all the relocated
(Fig. 7d). Although not all the wells have their corresponding events above magnitude 1.4 with quality A and B focal mechanisms in

231
G. Lin, B. Wu Geothermics 71 (2018) 225–233

Fig. 9. (a) Background seismicity rate compared with the operational fluid volumes at the
Geyers Geothermal Field. (b) Scalar solutions for quality A and B focal mechanisms
(M ≥ 1.4). Annual average is shown by the black circles. Vertical arrows mark 2008,
when the average focal scalar started to increase.

Fig. 9b. There is a gap in the focal mechanism data in 2007 because the
waveforms are not available from the NCSN for events between 2007/ Fig. 10. (a) Number of earthquakes above magnitude 1.4 per year with quality A and B
04/23 and 2007/07/20. The average focal solution before 2008 is focal mechanisms. (b) Percentage of each focal type per year. Pink arrows mark the time
period of the Northwest Geysers EGS project (2008/09/30–2012/10/31). (For inter-
about −0.37, indicating the dominating normal and strike-slip faulting
pretation of the references to color in this figure legend, the reader is referred to the web
regime. However, it rose to about −0.23 since 2008, which was mainly version of the article.)
due to the increase in the number of reverse faulting after 2008. In
order to examine this temporal change, we plot both the number of
focal solutions and percentage of each focal type as a function of time in are dominated by normal and strike-slip regimes throughout our entire
Fig. 10 and found that both have increased since 2008. In addition, the study time period, there is an increase in reverse faulting since 2008,
percentage of normal faulting shows the corresponding reduction. This which may be due to the thermal contraction associated with the
focal type variation matches with the dates of the Northwest Geysers Northwest Geysers EGS project. Our study provides a groundwork for
EGS project (2008/09/30–2012/10/31) very well, which aimed at en- the future remote triggering study and other seismological studies in
hancing production from a known high temperature reservoir located The Geysers.
under a conventional geothermal steam reservoir. The increased re-
verse faulting after 2008 in our study area is consistent with the thermal Acknowledgements
contraction created by this project.
Seismic data for this study were accessed through the Northern
California Earthquake Data Center (NCEDC), doi:10.7932/NCEDC. We
5. Conclusions thank the anonymous reviewer who provided constructive comments.
Figures were made using the public domain GMT software (Wessel and
In this paper, we present a comprehensive investigation of the Smith, 1991) and matlab. This research work was supported by the
structural variations near the Geysers Geothermal Field, including a National Science Foundation grant EAR-1447105.
new 3-D Vp and Vp/Vs tomographic model, precise earthquake reloca-
tions, and focal mechanisms based on local seismic data. Our Vp model References
mainly reflects the local geological features and the Vp/Vs model is
more sensitive to the fluid saturation in the geothermal reservoir. The Allis, R., 1982. Mechanism of induced seismicity at The Geysers geothermal reservoir,
local seismicity shows both spatial and temporal correlations with the California. Geophys. Res. Lett. 9, 629–632.
geothermal operations in the study area. Although the focal solutions Alwyn, R.C., 1996. The geysers geothermal area, California: tomographic images of the

232
G. Lin, B. Wu Geothermics 71 (2018) 225–233

depleted steam reservoir and non-double-couple earthquakes. Durham theses), Lin, G., Shearer, P.M., Hauksson, E., 2007. Applying a three-dimensional velocity model,
Durham University. Durham E-Theses Online: http://etheses.dur.ac.uk/5349/. waveform cross correlation, and cluster analysis to locate Southern California seis-
Boitnott, G.N., Kirkpatrick, A., 1997. Interpretation of field seismic tomography at The micity from 1981 to 2005. J. Geophys. Res. 112. http://dx.doi.org/10.1029/
Geysers geothermal field, California. In: Proceedings, Twenty-Second Workshop on 2007JB004986.
Geothermal Reservoir Engineering, Stanford University. Stanford, California, January Lin, G., Shearer, P.M., Matoza, R.S., Okubo, P.G., Amelung, F., 2014. Three-dimensional
27–29, 1997, SGP-TR-I 55. seismic velocity structure of Mauna Loa and Kilauea volcanoes in Hawaii from local
Boyle, K., Zoback, M., 2014. The stress state of the northwest geysers, California geo- seismic tomography. J. Geophys. Res. 119, 4377–4392. http://dx.doi.org/10.1002/
thermal field, and implications for fault-controlled fluid flow. Bull. Seismol. Soc. Am. 2013JB010820.
104, 1–10. Lin, G., Thurber, C.H., Zhang, H., Hauksson, E., Shearer, P.M., Waldhauser, F., Brocher,
Bufe, C.G., Shearer, P., 1981. Geothermal induced seismicity-Bane or blessing. T.M., Hardebeck, J., 2010. A California statewide three-dimensional seismic velocity
Proceedings Second DOE-ENEL Workshop for Cooperative Research in Geothermal model from both absolute and differential times. Bull. Seismol. Soc. Am. 100,
Energy 86–97. 225–240. http://dx.doi.org/10.1785/0120090028.
Denlinger, R., Bufe, C., 1980. Seismicity induced by steam production at The Geysers Majer, E.L., Mcevilly, T.V., 1979. Seismological investigations at the Geysers geothermal
steam field in northern California. Eos Trans. AGU 61, 1051. field. Geophysics 44, 246–269.
Eberhart-Phillips, D., 1986. Three-dimensional velocity structure in the northern McLaughlin, R.J., 1981. Tectonic setting of pre-Tertiary rocks and its relation to geo-
California Coast Ranges from inversion of local earthquake arrival times. Bull. thermal resources in The Geysers-Clear Lake area. Research in The Geysers-Clear
Seismol. Soc. Am. 76, 1025–1052. Lake Geothermal Area, Northern California, vol. 1141. pp. 3–23.
Eberhart-Phillips, D., 1990. Three-dimensional P and S velocity structure in the Coalinga Meador, P.J., Hill, D.P., Luetgert, J.H., 1985. Data report for the July–August 1983
region, California. J. Geophys. Res. 95, 15343–15363. seismic-refraction experiment in the Mono Craters-Long Valley region, California.
Eberhart-Phillips, D., Oppenheimer, D.H., 1984. Induced seismicity in The Geysers geo- U.S. Geol. Surv. Open-File Rept. 85-708 70 pp.
thermal area, California. J. Geophys. Res. 89, 1191–1207. NCEDC, 2014. Northern California Earthquake Data Center. UC Berkeley Seismological
Efron, B., Gong, G., 1983. A leisurely look at the bootstrap, the jackknife and cross-va- Laboratory. Dataset. http://dx.doi.org/10.7932/NCEDC.
lidation. Am. Stat. 37, 36–48. O’Connell, D.R.H., Johnson, L.R., 1991. Progressive inversion for hypocenters and P wave
Efron, B., Tibshirani, R., 1991. Statistical data analysis in the computer age. Science 253, and S wave velocity structure: application to the Geysers, California, Geothermal
390–395. Field. J. Geophys. Res. 96, 6223–6236. http://dx.doi.org/10.1029/91JB00154.
Foulger, G.R., Grant, C.C., Ross, A., Julian, B.R., 1997. Industrially induced changes in Oppenheimer, D.H., 1986. Extensional tectonics at The Geysers Geothermal Area,
Earth structure at The Geysers geothermal area, California. Geophys. Res. Lett. 24, California. J. Geophys. Res.: Solid Earth 91, 11463–11476. http://dx.doi.org/10.
135–137. 1029/JB091iB11p11463.
Gritto, R., Jarpe, S.P., 2014. Temporal variations of Vp/Vs-ratio at The Geysers geo- Oppenheimer, D.H., Herkenhoff, K.E., 1981. Velocity–density properties of the litho-
thermal field, USA. Geothermics 52, 112–119. http://dx.doi.org/10.1016/j. sphere from three-dimensional modeling at the Geysers-Clear Lake region, California.
geothermics.2014.01.012. J. Geophys. Res. 86, 6057–6065.
Gritto, R., Yoo, S.H., Jarpe, S.P., 2013. Three-dimensional seismic tomography at the Romero, A.E., McEvilly, T.V., Majer, E.L., Vasco, D., 1995. Characterization of the geo-
Geysers geothermal field, CA, USA. In: Proceedings, Thirty-Eighth Workshop on thermal system beneath the Northwest Geysers steam field, California, from seismi-
Geothermal Reservoir Engineering, Stanford University. Stanford, California, city and velocity patterns. Geothermics 24, 471–487. http://dx.doi.org/10.1016/
February 11-13, 2013, SGP-TR-198. 0375-6505(95)00003-9.
Gunasekera, R., Foulger, G., Julian, B., 2003. Reservoir depletion at The Geysers geo- Shearer, P.M., Prieto, G.A., Hauksson, E., 2006. Comprehensive analysis of earthquake
thermal area, California, shown by four-dimensional seismic tomography. J. source spectra in Southern California. J. Geophys. Res. 111, B06303. http://dx.doi.
Geophys. Res. 108, 2134. http://dx.doi.org/10.1029/2001JB000638. org/10.1029/2005JB003979.
Hardebeck, J., Shearer, P., 2002. A new method for determining first-motion focal me- Stark, M.A., 1992. Microearthquakes – a tool to track injected water in The Geysers re-
chanisms. Bull. Seismol. Soc. Am. 92, 2264–2276. servoir. Monograph on the Geyser geothermal field, Special report 111–117.
Hardebeck, J.L., Michael, A.J., 2006. Damped regional-scale stress inversions: metho- Thompson, R.C., 1989. Structural stratigraphy and intrusive rocks at the geysers geo-
dology and examples for southern California and the Coalinga aftershock sequence. J. thermal field. Geotherm. Resour. Counc. Trans. 13, 481–485.
Geophys. Res. 111, B11310. http://dx.doi.org/10.1029/2005JB004144. Thompson, R.C., 1992. Structural stratigraphy and intrusive rocks at The Geysers geo-
Heidbach, O., Rajabi, M., Reiter, K., Ziegler, M., 2016. The WSM Team: World Stress Map thermal field. Monogr. Geysers Geotherm. Field 17, 59–63.
Database Release 2016, GFZ Data Services. Thurber, C., Eberhart-Phillips, D., 1999. Local earthquake tomography with flexible
Iyer, H., Oppenheimer, D., Hitchcock, T., Roloff, J., Coakley, J., 1981. Large teleseismic p- gridding. Comput. Geosci. 25, 809–818.
wave delays in the geysers-clear lake geothermal area. Research in The Geysers-Clear Thurber, C., Zhang, H., Brocher, T., Langenheim, V.E., 2009. Regional three-dimensional
Lake Geothermal Area, Northern California, vol. 1141. pp. 97–116. seismic velocity model of the crust and uppermost mantle of northern California. J.
Julian, B.R., Ross, A., Foulger, G.R., Evans, J.R., 1996. Three-dimensional seismic image Geophys. Res. 114, B01304. http://dx.doi.org/10.1029/2008JB005766.
of a geothermal reservoir: the Geysers, California. Geophys. Res. Lett. 23, 685–688. Thurber, C.H., 1983. Earthquake locations and three-dimensional crustal structure in the
http://dx.doi.org/10.1029/95GL03321. Coyote Lake area, central California. J. Geophys. Res. 88, 8226–8236.
Kirkpatrick, A., Peterson, J.J., Majer, E., 1997. Three-dimensional compressional- and Thurber, C.H., 1993. Local earthquake tomography: velocities and Vp/Vs-theory. In: Iyer,
shear-wave seismic velocity models for the southeast Geysers geothermal field, H.M., Hirahara, K. (Eds.), Seismic Tomography: Theory and Practice. Chapman and
California. Proceedings 27th Workshop on Geothermal Reservoir Engineering, vol. 22 Hall, London, pp. 563–583.
399–410. Trugman, D.T., Shearer, P.M., Borsa, A.A., Fialko, Y., 2016. A comparison of long-term
Lin, G., 2013. Three-dimensional seismic velocity structure and precise earthquake re- changes in seismicity at The Geysers, Salton Sea, and Coso geothermal fields. J.
locations in the Salton Trough, southern California. Bull. Seismol. Soc. Am. 103, Geophys. Res. 121. http://dx.doi.org/10.1002/2015JB012510.
2694–2708. http://dx.doi.org/10.1785/0120120286. Wessel, P., Smith, W.H.F., 1991. Free software helps map and display data. Eos Trans.
Lin, G., 2015. Seismic velocity structure and earthquake relocation for the magmatic AGU 72, 441.
system beneath Long Valley Caldera, eastern California. J. Volcanol. Geotherm. Res. Wiemer, S., Baer, M., 2000. Mapping and removing quarry blast events from seismicity
296, 19–30. catalogs. Bull. Seismol. Soc. Am. 90, 525–530.
Lin, G., Okubo, P.G., 2016. A large refined catalog of earthquake relocations and focal Zhang, Q., Lin, G., Zhan, Z., Chen, X., Qin, Y., Wdowinski, S., 2017. Absence of remote
mechanisms for the Island of Hawai’i and its seismotectonic implications. J. Geophys. earthquake triggering within the Coso and Salton Sea geothermal production fields.
Res. http://dx.doi.org/10.1002/2016JB013042. Geophys. Res. Lett. http://dx.doi.org/10.1002/2016GL071964.
Lin, G., Shearer, P.M., 2009. Evidence for water-filled cracks in earthquake source re- Zucca, J., Hutchings, L., Kasameyer, P., 1994. Seismic velocity and attenuation structure
gions. Geophys. Res. Lett. 36. of the Geysers geothermal field, California. Geothermics 23, 111–126.

233

You might also like