Wastewater disposal and earthquake swarm activity at the southern end of the Central Valley, California

Similar documents
A comparison of seismicity rates and fluid injection operations in Oklahoma and California: Implications for crustal stresses

3D Finite Element Modeling of fault-slip triggering caused by porepressure

A comparison of seismicity rates and fluid-injection operations in Oklahoma and California: Implications for crustal stresses

Modeling pressure response into a fractured zone of Precambrian basement to understand deep induced-earthquake hypocenters from shallow injection

How will induced seismicity in Oklahoma respond to decreased saltwater injection rates?

An Investigation on the Effects of Different Stress Regimes on the Magnitude Distribution of Induced Seismic Events

FAQs - Earthquakes Induced by Fluid Injection

Limitations of Earthquake Triggering Models*

What We Know (and don t know)

A century of oil-field operations and earthquakes in the greater Los Angeles Basin, southern California

What allows seismic events to grow big?: Insights from fault roughness and b-value analysis in stick-slip experiments

Tectonic Seismogenic Index of Geothermal Reservoirs

SI: Wastewater disposal and earthquake swarm activity at the southern end of the Central Valley, California

Location uncertainty for a microearhquake cluster

SHale gas. in relations to technological activity at The Geysers geothermal field, California

IGF PAS activity in WP4

Kinematic inversion of pre-existing faults by wastewater injection-related induced seismicity: the Val d Agri oil field case study (Italy)

Enhanced remote earthquake triggering at fluid injection sites in the Midwestern US

Chapter 6. Conclusions. 6.1 Conclusions and perspectives

Characterization of Induced Seismicity in a Petroleum Reservoir: A Case Study

Automatic Moment Tensor Analyses, In-Situ Stress Estimation and Temporal Stress Changes at The Geysers EGS Demonstration Project

Recent Earthquakes in Oklahoma. and Potential for Induced Seismicity Austin Holland Oklahoma State Seismologist

Extending the magnitude range of seismic reservoir monitoring by Utilizing Hybrid Surface Downhole Seismic Networks

Key words induced earthquake water injection experiment Nojima fault diffusion process pore water pressure permeability

Microseismic monitoring of borehole fluid injections: Data modeling and inversion for hydraulic properties of rocks

California foreshock sequences suggest aseismic triggering process

Addressing the risks of induced seismicity in sub-surface energy operations

Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source

Modeling Injection-Induced Seismicity with the Physics-Based Earthquake Simulator RSQSim

RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE

of other regional earthquakes (e.g. Zoback and Zoback, 1980). I also want to find out

Finite element modelling of fault stress triggering due to hydraulic fracturing

SUPPLEMENTAL INFORMATION

MEASUREMENT OF HYDRAULICALLY ACTIVATED SUBSURFACE FRACTURE SYSTEM IN GEOTHERMAL RESERVOIR BY USING ACOUSTIC EMISSION MULTIPLET-CLUSTERING ANALYSIS

The Coso Geothermal Area: A Laboratory for Advanced MEQ Studies for Geothermal Monitoring

High-resolution analysis of microseismicity related to hydraulic stimulations in the Berlín Geothermal Field, El Salvador

USGS: USGS: NEIC NEIC

Imaging sharp lateral velocity gradients using scattered waves on dense arrays: faults and basin edges

Table One: Induced seismicity since 2011 in five previously aseismic Ohio counties. Locality County Year Induced Recorded Reference

Figure 1 shows a sketch of loading conditions and sample geometry of the employed Westerly

Di#erences in Earthquake Source and Ground Motion Characteristics between Surface and Buried Crustal Earthquakes

Seismological monitoring of the GRT1 hydraulic stimulation (Rittershoffen, Alsace, France)

Velocity contrast along the Calaveras fault from analysis of fault zone head waves generated by repeating earthquakes


Earthquake Measurements

stress direction are less stable during both drilling and production stages (Zhang et al., 2006). Summary

Identifying fault activation during hydraulic stimulation in the Barnett shale: source mechanisms, b

Geomechanical Analysis of Hydraulic Fracturing Induced Seismicity at Duvernay Field in Western Canadian Sedimentary Basin

Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip

Does Aftershock Duration Scale With Mainshock Size?

Oil and natural gas production from shale formations

SUPPLEMENTARY INFORMATION

Induced seismicity in Enhanced Geothermal Systems (EGS) in Alsace, France. Jean Schmittbuhl 1

Julie Shemeta, MEQ Geo Inc., WCEE Webinar 1/13/2016 1/14/2016

Hydraulic fracturing and seismicity in the Western Canada Sedimentary Basin

Daniel O Connell and Robert Creed. Fugro Consultants, Inc Cole Blvd., Suite 230, Lakewood CO

Risk Evaluation. Todd Shipman PhD, Alberta Geological Survey/Alberta Energy Regulator November 17 th,2017 Induced Seismicity Workshop, Yellowknife NWT

State of Stress in Seismic Gaps Along the SanJacinto Fault

Unraveling the Mysteries of Seismicity in Oklahoma. A Story Map

Spatial clustering and repeating of seismic events observed along the 1976 Tangshan fault, north China

Fracture induced shear wave splitting in a source area of triggered seismicity by the Tohoku-oki earthquake in northeastern Japan.

Rotation of the Principal Stress Directions Due to Earthquake Faulting and Its Seismological Implications

Potential Injection-Induced Seismicity Associated With Oil & Gas Development

Effects of Surface Geology on Seismic Motion

Concerns About the Poten/al for Induced Seismicity Associated with the Mississippian Play: Perceived or Real?

Thomas H. W. Goebel v (001)

Aftershocks are well aligned with the background stress field, contradicting the hypothesis of highly heterogeneous crustal stress

Earth Tides Can Trigger Shallow Thrust Fault Earthquakes

Part 2 - Engineering Characterization of Earthquakes and Seismic Hazard. Earthquake Environment

Estimating energy balance for hydraulic fracture stimulations: Lessons Learned from Basel

Supporting Information for Break of slope in earthquake-size distribution reveals creep rate along the San Andreas fault system

The Non-volcanic tremor observation in Northern Cascadia. Hsieh Hsin Sung 3/22

Application of Phase Matched Filtering on Surface Waves for Regional Moment Tensor Analysis Andrea Chiang a and G. Eli Baker b

External Grant Award Number 04HQGR0058 IMPROVED THREE-DIMENSIONAL VELOCITY MODELS AND EARTHQUAKE LOCATIONS FOR CALIFORNIA

Negative repeating doublets in an aftershock sequence

CHANGES OF COULOMB FAILURE STRESS DUE TO DISLOCATIONS DURING STIMULATION OF GPK2 AT SOULTZ-SOUS-FORÊTS

Calculation of Focal mechanism for Composite Microseismic Events

Passive seismic monitoring in unconventional oil and gas

RISK ANALYSIS AND RISK MANAGEMENT TOOLS FOR INDUCED SEISMICITY

Relocation of aftershocks of the Wenchuan M S 8.0 earthquake and its implication to seismotectonics

Earthquake patterns in the Flinders Ranges - Temporary network , preliminary results

Some aspects of seismic tomography

Hydraulic Fracturing and Seismicity in the Western Canada Sedimentary Basin

to: Interseismic strain accumulation and the earthquake potential on the southern San

Effect of an outer-rise earthquake on seismic cycle of large interplate earthquakes estimated from an instability model based on friction mechanics

Regional tectonic stress near the San Andreas fault in central and southern California

Once you have opened the website with the link provided choose a force: Earthquakes

An Open Air Museum. Success breeds Success. Depth Imaging; Microseismics; Dip analysis. The King of Giant Fields WESTERN NEWFOUNDLAND:

On May 4, 2001, central Arkansas experienced an M=4.4 earthquake followed by a

Abstracts ESG Solutions

Characterizing and Responding to Seismic Risk Associated with Earthquakes Potentially Triggered by Saltwater Disposal and Hydraulic Fracturing

1 of 5 12/10/2018, 2:38 PM

BRIEFING MEMO ON RESERVOIR TRIGGERED SEISMICITY (RTS)

Anomalous early aftershock decay rate of the 2004 Mw6.0 Parkfield, California, earthquake

Ground displacement in a fault zone in the presence of asperities

SOURCE MODELING OF RECENT LARGE INLAND CRUSTAL EARTHQUAKES IN JAPAN AND SOURCE CHARACTERIZATION FOR STRONG MOTION PREDICTION

Short Note Source Mechanism and Rupture Directivity of the 18 May 2009 M W 4.6 Inglewood, California, Earthquake

DETAILED IMAGE OF FRACTURES ACTIVATED BY A FLUID INJECTION IN A PRODUCING INDONESIAN GEOTHERMAL FIELD

Measurements in the Creeping Section of the Central San Andreas Fault

From an earthquake perspective, 2011 was. Managing the Seismic Risk Posed by Wastewater Disposal. 38 EARTH April

Transcription:

GEOPHYSICAL RESEARCH LETTERS, VOL.???, XXXX, DOI:10.1029/, Wastewater disposal and earthquake swarm activity at the southern end of the Central Valley, California T. H. W. Goebel, 1* S. M. Hosseini, 2 F. Cappa, 3,4 E. Hauksson, 1 J. P. Ampuero, 1 F. Aminzadeh, 2 J. B. Saleeby 1 Key Points anomalous swarm activity associated with White-Wolf fault induced seismicity likely caused by localized pressure-increase along a seismically active fault induced seismicity may be masked by natural earthquake activity in California T. H. W. Goebel, University of California, Santa Cruz 1156 High St. Santa Cruz, 95064, California, tgoebel@ucsc.edu 1 Division of Geological and Planetary

X - 2 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 1 2 3 4 5 6 7 8 9 10 11 Fracture and fault zones can channel fluid-flow and transmit injection-induced pore pressure changes over large distances (>km), at which seismicity is rarely suspected to be human-induced. We use seismicity analysis and hydrogeological models to examine the role of seismically active faults in inducing earthquakes. We analyze a potentially injection-induced earthquake swarm with three events above M4 near the White-Wolf fault (WWF). The swarm deviates from classic main-aftershock behavior, exhibiting uncharacteristically low Gutenberg-Richter b of 0.6, and systematic migration patterns. Some smaller events occurred south-east of the WWF in an area of several disposal wells, one of which became active just five month before the main swarm activity. Hydrogeological modeling revealed that wastewater disposal likely contributed Sciences, CA Institute of Technology, 91125 Pasadena, CA, USA. *now at University of California, Santa Cruz 2 Department of Chemical Engineering and Material Sciences, University of Southern CA, 90089 Los Angeles, CA, USA. 3 Geoazur Laboratory, University of Nice Sophia-Antipolis, Nice, France. 4 Institut Universitaire de France, Paris, France

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 3 12 13 14 15 to seismicity via localized pressure-increase along a seismically active fault. Our results suggest that induced seismicity may remain undetected in CA without detailed analysis of local geologic setting, seismicity and fluid diffusion.

X - 4 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 1. Introduction 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 Fluid injection into hydrocarbon and geothermal reservoirs can change the local stress field potentially causing earthquakes at several kilometers distance, both immediately and months to years after peak-injection [e.g. Hsieh and Bredehoeft, 1981; Bourouis and Bernard, 2007; Keranen et al., 2013; Kim, 2013; Martínez-Garzón et al., 2014; Schoenball et al., 2014; Rubinstein et al., 2014]. Injection-induced fault slip is, in addition to poroelastic stress changes, commonly attributed to an increase in pore-pressure, which reduces the frictional resistance on fault surfaces [e.g. Healy et al., 1968; Hsieh and Bredehoeft, 1981; Ellsworth, 2013; Keranen et al., 2013]. In recent years, induced earthquakes have been observed in the proximity of many high-volume wastewater disposal (WD) wells in the central U.S. [e.g. Horton, 2012; Frohlich and Brunt, 2013; Keranen et al., 2013; Kim, 2013; Skoumal et al., 2014]. Many of the identified injection wells responsible for inducing seismicity are suspected to inject at depth close to the upper basement surface, where critically stressed faults slip more easily when exposed to changing pressures and earthquake ruptures can grow to larger sizes than in sedimentary basins [e.g. Das and Scholz, 1983; Horton, 2012; Keranen et al., 2013; Kim, 2013; Ellsworth, 2013]. While many previous studies focused on regions within the central and eastern U.S., where induced event detection is facilitated by low background seismicity rates, detecting induced events in California hydrocarbon basins received less attention, despite extensive injection operations and seismically active faults. Up to now, few injection-induced earthquakes outside of geothermal reservoirs have been observed in CA [Kanamori and Hauksson, 1992; Goebel et al., 2015]. Consequently, the most acute anthropogenically-

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 5 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 controlled seismic hazard stems from fluid-injection activity close to active faults. Such injection activity can lead to earthquakes up to M w 5.7 or larger as observed in OK [Keranen et al., 2013]. In CA, annual fluid-injection volumes exceed those in OK where induced seismicity is suggested to be widespread. The average number of active injection wells between 2010 and 2013 of 9,900 exceeds the 8,600 wells in OK [CA Department of Conservation, 2012]. The wells in CA inject on average at depth of 1.5 km which is about 0.5 km deeper than in OK [Goebel, 2015]. More recently and coincident with continuously increasing oilprices between 2001 and 2014 (except 09 & 10), CA experienced a systematic increase in injection volumes in connection with more extensive well-stimulation operations. This increase in injection activity together with the lack of available injection sites away from active faults require a more detailed assessment of possible seismogenic consequences of fluid injection in CA. This study is structured as follows: we first provide an overview of the geologic setting, as well as injection and seismic activity at the southern end of the Central Valley. This is followed by a detailed assessment of the seismically active faults within the study region. We then investigate a possible correlation between fluid-injection and seismicity, including variations in frequency-magnitude distribution with the onset of injection rate increase. Lastly, we create a detailed hydrogeological model that incorporates local geology such as active fault structures and examine injection-induced pressure changes.

X - 6 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 2. Swarm activity, wastewater disposal operations and geological setting 57 58 59 This study concentrates on a potentially injection-induced earthquake swarm in 2005, which is associated with the White-Wolf fault (WWF) and occurred at the southern end of the Central Valley, CA. The swarm, which is referred to as White Wolf swarm in the 60 following, deviates from standard mainshock-aftershock patterns. It is comprised of a 61 M L 4.5 event on September 22 nd, followed by two M L 4.7 (M w 4.6) and M L 4.3 events the 62 same day as well as some smaller magnitude fore-shocks. The White Wolf swarm is 63 64 65 66 67 68 69 70 71 72 73 suspected to be connected to fluid injection activity based on a statistical assessment of injection and seismicity rate changes [Goebel et al., 2015]. The statistical assessment showed that an abrupt increase in injection rates in 2005 was followed by a large increase in seismicity rates, which exceeded the 95% confidence interval of previous rate variations since 1980. In other words, the area did not experience a comparable rate increase within a 10 km radius of the well prior to the start of injection in 2005. Moreover, the strong correlation between rapid injection rate changes and the subsequent seismicity sequence had a 3% probability of coinciding by chance based on tests with randomly determined onsets of injection rate changes [Goebel et al., 2015]. The White Wolf swarm occurred at the southern end of Kern County, the largest oilproducing (>75% of the state s oil production) and fluid-injecting (>80% of all injection 74 wells) county in CA [CA Department of Conservation, 2012]. The region experienced 75 76 77 one major earthquake since 1950, i.e the 1952, M w 7.3 Kern county event, and hosts many seismically active faults including the White Wolf, Wheeler Ridge and Pleito faults (Figure 1a).

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 7 78 79 80 81 82 83 84 85 86 87 88 89 Fluid injection rates at the southern end of the Central Valley increased rapidly from 20,000 to more than 100,000 m 3 /mo between 2001 and 2010 (Figure 1b). The majority of wastewater, i.e. 80 to 95%, was injected within the Tejon oilfield involving only three closely spaced wells. These wells started injecting in 2001 (WD01), 2004 (WD04) and 2005 (WD05), and injection fluids were generally contained below 950 m. Effective well-depths are reported between 1200 to 1500 m [CA Department of Conservation, 2012]. The injection wells targeted a 25 30 m thin, highly-permeable (0.2 1.0 10 12 m 2 ) stratigraphic zone within the Monterey formation [CA Department of Conservation, 2012]. This injection zone is comprised of turbiditic sand lenses with maximum lateral extents of 1 to 2 km (Figure S2). Moreover, the wellbores include a significant, horizontally drilled portion with 520 to 580 m long perforation zones that maximized injection rates. Injection into well WD05 occurred at consistently high rates of 57,000 m 3 /mo, start- 90 ing five months before the White Wolf swarm in September 2005. During these five 91 92 93 94 95 96 97 98 99 months, the wellbore accommodated more than 75% of the total injection activity of the entire study area. WD05 is located in an area of closely-spaced, north-west-striking, Early-Miocene, buried, normal faults that show evidence of local Holocene reactivation. Based on geological mapping, seismicity and well-log data, we identified a seismically active normal fault located between the WWF and injection site WD05 (Figure 1a, see Suppl. Material Section S1 & S2 and Fig. S1-S3 for details of fault identification). This fault is referred to as Tejon fault in the following. Both seismicity and well-log data suggest that the Tejon fault is shallow close to the injection site and deepens toward the northwest below the Wheeler Ridge fault before intersecting with the WWF (Figure

X - 8 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 100 101 102 1c). The horizontally-elongated perforation zone of well WD05, which extends directly east from the well-head, increased the probability of intersecting the Tejon fault with the borehole. 3. Seismicity analysis and hydrogeological modeling 103 104 105 106 107 108 109 110 111 112 The White Wolf swarm shows evidence of systematic event migration between injection sites and the M w 4.6 hypocenter (Figure 1c & 2). This systematic migration is best seen in relocated catalogs that include newly detected events using a template-matching method and may easily remain undetected in standard seismicity catalogs (Figure 2). Within the first two months of wastewater disposal in WD05, we observed some shallow seismicity at 4 km depth beneath the injection site and on the Tejon fault to the northwest, while little to no seismic activity occurred to the south and southeast. Within the following months, the area northwest of well WD05 in direction of the WWF became progressively active. Much of this seismicity was concentrated just up-dip of the intersection point between the Tejon and White Wolf fault between 70 to 150 days after injection (see animations 113 in the online Suppl.). In the two years prior to injection into WD05, seismicity rates 114 115 116 117 118 119 120 within a 10 km radius of the well were largely constant (Figure 3a). The start of injection also marked the onset of seismicity rate increase which peaked about 150 days later. The occurrence of the three M L >4 events was closely followed by a 30% decrease in injection rates in well WD05 and injection rates kept decreasing over the ensuing 4 years. In addition to seismicity rate changes, the frequency-magnitude-distribution (FMD) strongly deviated from expected behavior of Gutenberg-Richter-type FMDs with b-values close to unity (Figure 3b). During the period of peak injection in WD05 the b-value within

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 9 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 a 10 km radius was 0.6. This value was significantly lower than before (b=1) and after (b=0.9) peak injection activity. Moreover, we observed a long-term decrease in b-value coincident with a rate increase in cumulative injection rates in WD01, WD04 and WD05 (Figure S4). Using Monte Carlo simulations of earthquake magnitudes based on the observed a-value, magnitude of completeness of 1.5, we compute the 95% confidence interval of FMDs with b=1. The observed FMD exceeds the upper confidence bound, indicating a significantly larger proportion of large-magnitude events during the White Wolf swarm which is also confirmed by the corresponding moment magnitudes for events above M 4 (see Suppl. Section S6 for details on moment magnitude computations). Decreasing and low b-values may be characteristic for gradual fault activation processes, observed in laboratory experiments [Goebel et al., 2013] and during fluid injection [e.g. Maxwell et al., 2009; Skoumal et al., 2014; Huang and Beroza, 2015]. To further test if the 2005 White Wolf swarm was connected to wastewater disposal, we model permeability structure and pressure diffusion close to the injection site. Pressure diffusion was likely influenced by high permeability along the seismically active part of the Tejon fault located N-W of well WD05. Slip along this fault may have been pivotal in maintaining elevated permeability down to the depth of the White Wolf swarm. The existence of a zone with higher permeability than the surrounding lithology is supported by several observations such as: 1) The asymmetric seismicity distribution concentrated N-W of well WD05 with little to no seismicity to the south and east of the injection sites. Most of these events occurred at depths between the M w 4.6 hypocenter and the injection depth (Fig. S5); 2) The presence of background seismic activity on the upper portion of

X - 10 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 the Tejon fault revealing that part of the fault was seismically active prior to injection; 3) The presence of several mapped faults and fracture zones located N-W of the injection site between WD05 and WWF (Fig. S3). Our 3-D numerical diffusion model includes three principal stratigraphic zones, in addition to the Tejon fault which is implemented as a vertical zone of elevated permeability (Figure 4a). These three zones are: 1) The injection zone, i.e. a 20-30 m thin turbiditic sand lens in the Monterey formation with a lateral extent of up to 1.5 km, labeled as Transition zone in the industry data [CA Department of Conservation, 2012]; 2) the crystalline (gneissic) basement complex, and 3) the Monterey formation. Permeability is high within the sand lenses (i.e. 1 D) and very low ( 10 4 md) above injection depth which is one of the requirements for the selection of an injection site. Similarly, permeability is low ( 10 4 md) within the basement and Monterey formations outside of the injection zone (permeabilities are reported in milli-darcy, 1 md 10 15 m 2 ). Our hydrogeological model is based on the most complete available datasets within the upper 2 3 km of sedimentary basins and includes seismicity records, geologic mapping results, and industry data (i.e. well-logs, stratigraphic columns and interpreted reservoir structure, see Suppl. Section 1, 2 & 7). Below these depths few geophysical data is 160 available except for the seismicity record. To account for larger uncertainty at these 161 162 163 depths, we chose a coarse modeling-scale which primarily resolves the influence of large- scale structural heterogeneity such as large fault zones. More detailed information about the model set-up can be found in the Online Supplement, Section S7.

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 11 164 165 166 167 168 169 170 171 172 173 174 175 176 Our modeling results show that thin fault zones with higher permeability lead to fluid-pressure increase sufficient to trigger earthquakes at distances similar to the M w 4.6 hypocenter. We performed a detailed sensitivity analysis of pressure changes resulting from varying fault zone width and permeability. Permeability was varied over 7 orders of magnitude between 10 3 to 10 3 md to account for generally large uncertainties in permeability measurements [Manga et al., 2012], and fault width was varied between 100 and 1000 m (Figure 4b & SI Tab. S1). The modeled pressure changes experienced by a fault at 11 km distance span four orders of magnitude depending on the initial conditions, with significantly stronger effects of permeability changes compared to variations in fault zone width (Figure 4b). For fault zone permeability above 300 md and fault width below 800 m, we observe a pressure increase at the M w 4.6 hypocenter of at least 0.01 MPa (Figure 4b), which is sufficient to induce seismicity on faults favorably oriented to slip [Keranen et al., 2014; Hornbach et al., 2015]. Fault zones with lower permeability may 177 result in similar magnitude pressure changes if they are sufficiently narrow. The here 178 179 180 181 182 183 184 185 described seismogenic consequences of thin, high-permeability pressure channels are in agreement with previous studies of induced seismicity in Colorado and Arkansas [Hsieh and Bredehoeft, 1981; Zhang et al., 2013]. The migration pattern of seismic events relative to injection into WD05 can be used to constrain a plausible range of permeability values by correlating seismicity and modeled pressure front location along the Tejon fault (Figure 4c). The shape of the pore pressure front in a distance-time plot depends on the particular reservoir geometry and permeability values with a commonly observed square-root dependence of distance on time for simple

X - 12 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 186 187 188 189 190 191 192 193 194 radial symmetric models (Figure 2) [Talwani and Acree, 1984; Shapiro et al., 1997]. Using the more complex 3-D reservoir and fault geometries in our model, we find that fault zone permeability between 150 and 250 md and a width smaller than 500 m best agrees with the observed seismicity for a change in fluid-pressure of 0.01 MPa. This is within the general range of fault zone permeability values inferred from seismicity migration [e.g. Ingebritsen and Manning, 2010; Manga et al., 2012]. Much of the seismicity occurred close to the arrival of the initial pressure pulse of 0.01 MPa indicating that faults within the area may have been critically stressed, i.e. stress levels were within a narrow range below the shear strength (Figure 4c inset). 4. Discussion 195 196 197 198 199 200 201 202 203 204 205 206 The identification of human-induced earthquakes in tectonically active regions such as California is generally complicated by the many naturally occurring seismicity sequences. To address these challenges, we used detailed statistical analysis methods [Goebel et al., 2015] as well as geological and diffusion models to evaluate a potentially induced origin of an earthquake sequence close to the WWF in 2005. This sequence deviates from commonly observed tectonic sequences in the area by showing significantly elevated seismicity rates above the background associated with a rapid increase in injection rates. Moreover, the seismicity sequence showed evidence for deep migration within the crystalline basement between injection wells and the near-by White Wolf fault suggesting that wastewater disposal likely contributed to triggering the earthquake swarm. The recorded largest magnitude event (M max =4.6) of the sequence and the cumulative injection volume (V tot 1.8 10 6 m 3 ) fall well within the trend of V tot and M max reported by McGarr [2014],

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 13 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 and are similar to observations of V tot and M max in Timpson, Texas and Painesville, Ohio [McGarr, 2014]. This analysis assumes that injection in all three wells (WD01, WD04 and WD05) contributed to the induced seismicity sequence. Our results highlight that injection related earthquake triggering processes may involve multiple mechanisms. A plausible mechanism for the triggering of the White Wolf swarm is the diffusion of pressures from the 1.5 km deep injection site along the northwest-striking Tejon fault into the intersection zone with the WWF. This pressure channeling effect may have been further intensified if the WWF acted as flow barrier thereby trapping the pressure front within the damage zone of the Tejon fault resulting in more rapid pressure increase at the intersection between the two faults. Other triggering mechanisms may have included stress transfer at the front of the pressurized zone as well as injection-induced aseismic slip, which progressively became more seismogenic at larger depth within the basement complex. Shallow aseismic slip is well-documented in high-resolution, controlled injection experiments and may hide early fault activation processes [Cornet et al., 1997; Guglielmi et al., 2015]. A potentially injection-induced origin of the White-Wolf swarm is intrinsically connected to the specific geologic setting that accommodated pressure diffusion to seismogenic depth at the southern end of the Central Valley. More detailed assessments of the geologic setting close to injection wells are required to explain the lack of large-scale injection-induced earthquake activity in CA hydrocarbon basins [Goebel, 2015]. Cases of relatively deep induced seismicity far from injection sites have been reported in several other regions such as Oklahoma, Colorado and Arkansas, where induced earth-

X - 14 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 quakes occurred at 8 km depth and 7 to 35 km distance from the injection well [Hsieh and Bredehoeft, 1981; Horton, 2012; Keranen et al., 2014]. In addition to these large distances, there are several other factors that can complicate induced seismicity detection in CA hydrocarbon basins: 1) Faults may channel induced pressure changes and lead to localized pore-pressure increase; 2) Background seismicity rates are generally high so that small rate variations cannot be detected; 3) Small magnitude earthquakes close to injection sites can easily be missed because of sparse station coverage within hydrocarbon basins; and 4) Transient aseismic and seismic slip processes lead to a dynamic increase in permeability and progressive fault activation [Cornet et al., 1997; Bourouis and Bernard, 2007; Guglielmi et al., 2015; Wei et al., 2015]. All of these factors can potentially complicate mitigation strategies such as traffic light systems which rely on a systematic seismicity rate increase and event migration from the well. Based on our empirical results injection induced earthquakes are expected to contribute marginally to the overall seismicity in CA [see also Goebel, 2015; Goebel et al., 2015]. This is in line with physical models of crustal strength distributions which suggest a limit to the amount of strain energy that is available for shallow earthquake ruptures [e.g. Sibson, 1974]. Moreover, the frictional properties of shallow, sedimentary faults inhibit rupture growth and diminish the seismogenic impact of surficial pressure perturbation [e.g. Das and Scholz, 1983]. However, considering the numerous active faults in CA, the seismogenic consequences of even a few induced cases can be devastating.

5. Conclusion GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 15 249 Wastewater injection-induced earthquakes are rare in California compared to 250 widespread tectonic seismicity. Nevertheless, the proximity of high-rate injectors and 251 252 253 254 255 256 257 258 259 260 261 262 263 264 large active faults can cause noticeable earthquakes under certain geologic conditions. Our results suggest a connection between wastewater disposal and seismicity with events up to M w 4.6 at the southern end of the Central Valley. Wastewater injection within this region should be monitored carefully because of the presence of high-permeability fault structures that connect the injection site with the nearby WWF. The relatively shallow crystalline basement south of the WWF may increase the probability of inducing earthquakes, if fluids migrate beyond the intended geologic formations. The present example shows that injection-induced earthquakes may remain unidentified in tectonically active regions if only standard seismicity catalogs with comparably high magnitudes of completeness are analyzed. We present a pathway to more reliable identification of possibly induced earthquakes by extending the seismicity records to lower magnitudes and by analyzing waveform relocated catalogs as well as hydrogeological models. Such a detailed analysis of the available data may help recognize regions with increased induced seismicity potential and prevent injection-induced seismicity in CA in the future. Acknowledgments 265 266 267 268 The authors thank Semechah Lui, Preston Jordan, Bill Foxall, Emily Brodsky and members of the Induced Seismicity Consortium at USC for helpful comments and discussions. The manuscript benefitted from comments by Art McGarr and Michael Wysession. The well and operational data was supplied by the CA Department of Conservation

X - 16 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 269 270 271 272 273 (www.doggr.com), and the seismicity data by the Southern California Earthquake Center (doi: 10.7909/C3WD3xH1) and Advanced National Seismic Systems. This study was funded by NEHRP/USGS Grants G14AP00075 & G15AP00095 to Caltech, by the Southern CA Earthquake Center (SCEC) under contribution number 12017 & 15168 and the Induced Seismicity Consortium (ISC) at USC.

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 17 References 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 Bourouis, S., and P. Bernard (2007), Evidence for coupled seismic and aseismic fault slip during water injection in the geothermal site of Soultz (France), and implications for seismogenic transients, Geophys. J. Int., 169 (2), 723 732. CA Department of Conservation (2012), Division of Oil, Gas and Geothermal resources: Production history in California, (ftp://ftp.consrv.ca.gov/pub/oil/annual reports, accessed Sept. 26, 2014). Cornet, F., J. Helm, H. Poitrenaud, and A. Etchecopar (1997), Seismic and aseismic slips induced by large-scale fluid injections, Pure appl. geophys., 150, 563 583. Das, S., and C. Scholz (1983), Why large earthquakes do not nucleate at shallow depths, Nature, 305, 621 623. Ellsworth, W. L. (2013), Injection-induced earthquakes, Science, 341 (6142). Frohlich, C., and M. Brunt (2013), Two-year survey of earthquakes and injection/production wells in the Eagle Ford Shale, Texas, prior to the M w 4.8 20 October 2011 earthquake, Earth Planet. Sc. Lett., 379, 56 63. Goebel, T. H. W. (2015), A comparison of seismicity rates and fluid injection operations in Oklahoma and California: Implications for crustal stresses, The Leading Edge, 34 (6), 640 648, doi:10.1190/tle34060640.1. Goebel, T. H. W., D. Schorlemmer, T. W. Becker, G. Dresen, and C. G. Sammis (2013), Acoustic emissions document stress changes over many seismic cycles in stick-slip experiments, Geophys. Res. Letts., 40, doi:10.1002/grl.50507.

X - 18 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 Goebel, T. H. W., E. Hauksson, F. Aminzadeh, and J.-P. Ampuero (2015), An objective method for the assessment of possibly fluid-injection induced seismicity in tectonically active regions in central California, J. Geophys. Res., 120, 1 20, doi: 10.1002/2015JB011895. Guglielmi, Y., F. Cappa, J.-P. Avouac, P. Henry, and D. Elsworth (2015), Seismicity triggered by fluid injection induced aseismic slip, Science, 348 (6240), 1224 1226. Healy, J., W. Rubey, D. Griggs, and C. Raleigh (1968), The denver earthquakes, Science, 161 (3848), 1301 1310. Hornbach, M. J., et al. (2015), Causal factors for seismicity near Azle, Texas, Nature communications, 6. Horton, S. (2012), Disposal of hydrofracking waste fluid by injection into subsurface aquifers triggers earthquake swarm in central Arkansas with potential for damaging earthquake, Seism. Res. Lett., 83 (2), 250 260. Hsieh, P. A., and J. D. Bredehoeft (1981), A reservoir analysis of the Denver earthquakes: A case of induced seismicity, J. Geophys. Res., 86 (B2), 903 920. Huang, Y., and G. C. Beroza (2015), Temporal variation in the magnitude-frequency distribution during the Guy-Greenbrier earthquake sequence, Geophysical Research Letters, 42 (16), 6639 6646. Ingebritsen, S., and C. Manning (2010), Permeability of the continental crust: dynamic variations inferred from seismicity and metamorphism, Geofluids, 10 (1-2), 193 205. Kanamori, H., and E. Hauksson (1992), A slow earthquake in the Santa Maria Basin, California, Bull. Seism. Soc. Am., 82 (5), 2087 2096.

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 19 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 Keranen, K., M. Weingarten, G. Abers, B. Bekins, and S. Ge (2014), Sharp increase in central Oklahoma seismicity since 2008 induced by massive wastewater injection, Science, 345 (6195), 448 451. Keranen, K. M., H. M. Savage, G. A. Abers, and E. S. Cochran (2013), Potentially induced earthquakes in Oklahoma, USA: Links between wastewater injection and the 2011 M w 5.7 earthquake sequence, Geology, 41 (6), 699 702. Kim, W.-Y. (2013), Induced seismicity associated with fluid injection into a deep well in Youngstown, Ohio, J. Geophys. Res., 118 (7), 3506 3518. Manga, M., I. Beresnev, E. E. Brodsky, J. E. Elkhoury, D. Elsworth, S. Ingebritsen, D. C. Mays, and C.-Y. Wang (2012), Changes in permeability caused by transient stresses: Field observations, experiments, and mechanisms, Rev. Geophys., 50 (2). Martínez-Garzón, P., G. Kwiatek, H. Sone, M. Bohnhoff, G. Dresen, and C. Hartline (2014), Spatiotemporal changes, faulting regimes and source-parameters of induced seismicity: A case study from the Geysers geothermal field, J. Geophys. Res. Maxwell, S. C., et al. (2009), Fault activation during hydraulic fracturing, in SEG International Exposition and Annual Meeting, Houston, Texas. McGarr, A. (2014), Maximum magnitude earthquakes induced by fluid injection, J. Geophys. Res., 119 (2), 1008 1019, doi:10.1002/2013jb010597. Rubinstein, J. L., W. L. Ellsworth, A. McGarr, and H. M. Benz (2014), The 2001 present induced earthquake sequence in the Raton Basin of northern New Mexico and southern Colorado, Bull. Seismol. Soc. Am.

X - 20 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 Schoenball, M., L. Dorbath, E. Gaucher, J. F. Wellmann, and T. Kohl (2014), Change of stress regime during geothermal reservoir stimulation, Geophys. Res. Lett., 41 (4), 1163 1170, doi:10.1002/2013gl058514. Shapiro, S. A., E. Huenges, and G. Borm (1997), Estimating the crust permeability from fluid-injection-induced seismic emission at the KTB site, Geophys. J. Int., 131, F15 F18. Shearer, P., E. Hauksson, and G. Lin (2005), Southern California hypocenter relocation with waveform cross-correlation, part 2: Results using source-specific station terms and cluster analysis, Bull. Seismol. Soc. Am., 95 (3), 904 915. Sibson, R. H. (1974), Frictional constraints on thrust, wrench and normal faults, Nature, 249, 542 544. Skoumal, R. J., M. R. Brudzinski, B. S. Currie, and J. Levy (2014), Optimizing multistation earthquake template matching through re-examination of the Youngstown, Ohio, sequence, Earth Plant. Sc. Lett., doi:10.1016/j.epsl.2014.08.033. Talwani, P., and S. Acree (1984), Pore pressure diffusion and the mechanism of reservoirinduced seismicity, Pure and Applied Geophysics, 122 (6), 947 965. Wei, S., et al. (2015), The 2012 Brawley swarm triggered by injection-induced aseismic slip, Earth and Planetary Science Letters, 422, 115 125. Zhang, Y., et al. (2013), Hydrogeologic controls on induced seismicity in crystalline basement rocks due to fluid injection into basal reservoirs, Groundwater, 51 (4), 525 538.

X - 21 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION B A Days a>er InjecAon 2005, M4.7 Mainshock High Rate InjecAon Well Buried remobilized early Exposed Miocene normal flt. c ol f e W hit W M4.7 ult Fa Oil Field Boundary All Wells M4.3 M4.5 Tejon, North ult ge Fa ler Rid Whee 1952, M7.3 Kern County Eq. Wheeler Ridge c WD05 WD04 WD01 Tejon Pleito Fault y alle l V F SA t 4 6 lt a ntr Ce San Andreas Faul 2 8 10 12 Figure 1. WD05 c' SE 0 NW Fau White Wolf Magnitude 4 3 2 1 c' lt Fau ck rlo a G SAF 01+04+05 Seismically AcAve "Tejon Fault" Wheeler R idge Fau lt Seismically AcAve Fault Structure Approx. Depth of Basement Magnitude 4 3 2 1 C 0 2 4 6 8 10 12 14 16 18 Seismicity and injection activity within the study area centered at the Tejon oilfield. A) Fault traces (black and green lines), epicenters (dots), injection wells (blue triangles) and oilfield locations (blue lines). Seismicity is colored according to days after injection into WD05, gray dots show background seismicity. Earthquake locations are from the standard SCSN catalog. B) Injection (blue) and seismicity rates (gray) between 1980 and 2014. C) Seismicity cross-section within a 5 km zone between c and c in (A). Seismicity up to 300 days after start of injection is shown be colored dots (see color-bar in A), background events by gray dots and Tejon fault seismicity prior to injection by black dots. The gray rectangle qualitatively highlights a zone of high seismic activity that coincides with the Tejon fault. Earthquake locations are based on a high-quality waveform relocated catalog [Shearer et al., 2005]. D R A F T January 13, 2016, 8:59am D R A F T

X - 22 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION Epicentral Distance [km] SCSN Catalog Relocated Catalog r t Time a3er Injec6on [dy] Figure 2. Seismicity migration along the Tejon fault relative to the start of injection in WD05. We show both the SCSN (gray dots) and relocated earthquake catalogs (red circles). The latter includes new event detections using a template matching approach. Epicentral distances are reported relative to well-head location, error-bars show average, absolute location uncertainties. The dashed curve highlights the expected trend for a square-root dependence of distance on time characteristic for a diffusive process.

GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION X - 23 A x10 4 B Local Magnitude Cumul. Number Eqs. 20 # Bg.- rate ~19 #/yr 95% confidence b = 1 b = 0.6 95% confidence Time Windows in (b) Local Magnitude Figure 3. Fluid injection rates in well WD05, seismic activity and frequency-magnitude distributions (FMDs) of the likely induced White Wolf swarm. A) Injection rates (blue), cumulative number of earthquakes (black) and event magnitudes (circles) within a 10 km radius of injection in well WD05. B) FMD before (green), during (light blue) and after (dark blue) peak injection rates as well as 95% confidence interval for a FMD with b=1.

X - 24 GOEBEL ET AL.: EARTHQUAKE SWARMS AND FLUID INJECTION ~ 1-1.5 km Basement & other Units N Alluvium Santa Margarita TransiAon ~ 30 m Monterey ~ 300 m WD05 WD04 WD01 Sand Lense A M 4.6 w 8 km Tejon Fault 7.7 km not to scale Pathway width [m] Pathway width [m] P [MPa] p 0.01 0.05 0.1 Pathway Permeability [md] B P p [MPa] - 5 Figure 4. Distance from InjecAon [km] log Pathway Permeability [md] k = 250 md k = 200 md k = 150 md C Time ader InjecAon [dy] Hydrogeological modeling and predicted pore pressure increase caused by injection in well WD05. A) Schematic representation of model set-up including the highpermeability zone between injection-site and M w 4.6 event. The hydro-geological model is based on geologic mapping and industry data described in Section S1 & S2 in the online supplement. B) Model-dependent pore pressure change ( P p ) at distance of the M w 4.6 event ( 11 km) for different values of fault zone width and permeability. C) Observed hypocenter migration along the Tejon fault zone (black circles). The theoretical position of a 0.01 MPa pressure front is shown for a 500 m wide fault zone and three different permeability values (150, 200 and 250 md). The M w 4.6 event is shown by a red star. The inset shows distribution of pore pressures at each hypocenter for k=200 md and w=500 m, vertical lines are mean and standard deviation.