ArticlePDF Available

Static behaviour of induced seismicity

Authors:

Abstract and Figures

The standard paradigm to describe seismicity induced by fluid injection is to apply nonlinear diffusion dynamics in a poroelastic medium. I show that the spatiotemporal behaviour and rate evolution of induced seismicity can, instead, be expressed by geometric operations on a static stress field produced by volume change at depth. I obtain laws similar in form to the ones derived from poroelasticity while requiring a lower description length. Although fluid flow is known to occur in the ground, it is not pertinent to the behaviour of induced seismicity. The proposed model is equivalent to the static stress model for tectonic foreshocks generated by the Non- Critical Precursory Accelerating Seismicity Theory. This study hence verifies the explanatory power of this theory outside of its original scope.
Content may be subject to copyright.
!
1!
Static behaviour of induced seismicity
A. Mignan
Institute of Geophysics, Swiss Federal Institute of Technology Zurich, Switzerland
arnaud.mignan@sed.ethz.ch
Abstract: The standard paradigm to describe seismicity induced by fluid injection is
to apply nonlinear diffusion dynamics in a poroelastic medium. I show that the
spatiotemporal behaviour and rate evolution of induced seismicity can, instead, be
expressed by geometric operations on a static stress field produced by volume change
at depth. I obtain laws similar in form to the ones derived from poroelasticity while
requiring a lower description length. Although fluid flow is known to occur in the
ground, it is not pertinent to the behaviour of induced seismicity. The proposed model
is equivalent to the static stress model for tectonic foreshocks generated by the Non-
Critical Precursory Accelerating Seismicity Theory. This study hence verifies the
explanatory power of this theory outside of its original scope.
1. Introduction
Induced seismicity is a growing concern for the energy-sector industry relying
on fluid injection in the deep parts of the Earth’s crust [Ellsworth, 2013; Mignan et
al., 2015]. At the same time, fluid injection sites provide natural laboratories to study
the impact of increased fluid pressure on earthquake generation [Majer et al., 2007].
Induced seismicity is characterised by two empirical laws, namely (i) a linear
relationship between the fluid mass m(t) injected up to time t and the cumulative
number of induced earthquakes N(t) and (ii) a parabolic induced seismicity envelope
radius r 𝑚(𝑡)
!
with n a positive integer [Shapiro and Dinske, 2009]. These two
!
2!
descriptive laws can be derived from the differential equations of poroelasticity [Biot,
1941] under various assumptions. The full description of the process requires complex
numeric modelling coupling fluid flow, heat transport and geomechanics [Rutqvist,
2011]. These models, numerically cumbersome, can become intractable because of
the sheer number of parameters [Miller, 2015]. Attempts to additionally correct for
the known discrepancies between Biot's theory and rock experiments have led to a
large variety of model assumptions [Berryman and Wang, 2001], indicating that
poroelasticity results are ambiguous.
I will demonstrate that a simple static stress model can explain the two
empirical laws of induced seismicity without requiring any concept of poroelasticity.
The proposed theoretical framework hence avoids the aforementioned shortcomings
by suggesting an origin of induced seismicity that does not involve fluid flow in a
porous medium (although fluid flow indeed occurs). Historically, a similar static
stress model was proposed for the tectonic regime under the Non-Critical Precursory
Accelerating Seismicity Theory (N-C PAST) [Mignan et al., 2007; Mignan, 2012]. Its
application to induced seismicity data will allow a more fundamental investigation of
the relationship between static stress and earthquake generation. To test the model, I
will use data from the 2006 Basel Enhanced Geothermal System (EGS) stimulation
experiment including the flow rate of injected fluids [Häring et al., 2008] and the
relocated catalogue of induced seismicity [Kraft and Deichmann, 2014].
2. The Non-Critical Precursory Accelerating Seismicity Theory (N-C PAST)
The N-C PAST has been proposed to explain the precursory seismicity
patterns observed before large earthquakes from geometric operations in the
spatiotemporal stress field generated by constant tectonic stress accumulation
!
3!
[Mignan et al., 2007; Mignan, 2012]. In particular, it provides a mathematical
expression of temporal power-laws without requiring local interactions between the
elements of the system [Sammis and Sornette, 2002; Mignan, 2011]. Therefore
earthquakes are considered passive (static) tracers of the stress accumulation process,
in contrast with active earthquake cascading in a critical process (hence the term
"non-critical"). The concept of self-organized criticality [Bak and Tang, 1989] is
seldom used to explain induced seismicity [Grasso and Sornette, 1998]. Since there is
no equivalent of a mainshock in induced seismicity, the criticality versus non-
criticality debate has limited meaning in that case. However, the underlying process of
static stress changes considered in the N-C PAST can be tested against the observed
spatiotemporal behaviour of induced seismicity.
The N-C PAST postulates that earthquake activity can be categorized in three
regimes background, quiescence and activation depending on the spatiotemporal
stress field σ(r,t)
𝜎 𝑟, 𝑡 =
𝜎
!
, 𝑡 < 𝑡
!
!
!
!
!
!!
!
!
!
𝜎
!
+ 𝜏 𝑡 𝑡
!
+ 𝜎
!
, 𝑡
!
𝑡 < 𝑡
!
(1)
defined from the boundary conditions σ(r ! +, t) = σ
0
*
and σ(r = 0, t) = σ
0
+ 𝜏t +
σ
0
*
, with h the depth of the fault segment base, r the distance along the stress field
gradient from the fault’s surface projection, σ
0
< 0 the stress drop associated to a
hypothetical silent slip occurring at t
0
at the base of the fault, 𝜏 the tectonic stress rate
on the fault, σ
0
*
the crustal background stress, n = 3 the spatial diffusion exponent for
static stress and t
f
the mainshock occurrence time [Mignan et al., 2007] (Fig. 1a).
Background, quiescence and activation regimes are defined by event densities δ
b0
,
δ
bm
, and δ
bp
for |σ| σ
0
*
± Δσ
*
, σ < σ
0
*
- Δσ
*
and σ > σ
0
*
+ Δσ
*
, respectively, with the
boundary layer ±Δσ
*
the background stress amplitude range. By definition, δ
bm
< δ
b0
!
4!
< δ
bp
with each seismicity regime assumed isotropic and homogeneous in space (i.e.
role of fault network neglected). Correlation between earthquake productivity and
static stress changes is well established [King, 2007]. The distinction of three unique
seismicity regimes with constant event density, the main assumption of the N-C
PAST, is discussed later on.
Figure 1: Seismicity spatiotemporal behaviour described by the N-C PAST static
stress model (tectonic case [Mignan, 2012]): (a) Spatiotemporal evolution of the stress
field σ(r,t) generated by constant stress accumulation 𝜏 on a fault located at r = 0 (Eq.
1). Background, quiescence and activation seismicity regimes are described by
densities of events δ
b0
, δ
bm
, and δ
bp
for |σ| σ
0
*
± Δσ
*
, σ < σ
0
*
- Δσ
*
and σ > σ
0
*
+
Δσ
*
, respectively; (b) Temporal evolution of quiescence and activation envelopes r
*
(t)
!
5!
with σ(r
*
) = σ
0
*
± Δσ
*
(Eq. 2); (c) Rate of events µ(t) in a disc of constant radius
max(r*) (Eq. 3); (d) Cumulative number of events N(t) (Eq. 4) of power-law form
(Eq. 5). With t
0
= 0, t
mid
= 1, t
f
= 2, h = 1, 𝜏 = 0.1, σ
0
*
= 0, Δσ
*
= 10
-2
, δ
bm
= 0.001, δ
b0
= 0.1, δ
bp
= 1, n = 3, k = π, d = 2, Δt = 0.01.
In the tectonic case, static stress changes are underloading due to hypothetical
precursory silent slip on the fault at t
0
followed by overloading due to hypothetical
asperities delaying rupture on the fault after t
p
*
[Mignan, 2012]. The three seismicity
regimes are then defined as solid spatiotemporal objects with envelopes
𝑟
!
(𝑡
!
𝑡 < 𝑡
!
) =
!(!
!
!!)
!!
+ 1
!/!
1
!/!
𝑟
!
(𝑡
!
< 𝑡 < 𝑡
!
) =
!(!!!
!
)
!!
+ 1
!/!
1
!/!
(2)
by applying to Eq. (1) the boundary conditions σ(r
Q
*
, t) = σ(0, t
m
*
) = σ
0
*
- Δσ
*
and
σ(r
A
*
, t) = σ(0, t
p
*
) = σ
0
*
+ Δσ
*
, respectively. The parameters t
m
*
= t
mid
- Δσ
*
/𝜏 and t
p
*
= t
mid
+ Δσ
*
/𝜏 represent the times of quiescence disappearance and of activation
appearance, respectively, with σ(0, t
mid
) = σ
0
*
. The background seismicity regime is
defined by subtracting the quiescence and activation envelopes r
A
*
(t) and r
Q
*
(t) from a
larger constant envelope r
max
max(r
*
) (Fig. 1b). While trivial along 𝑟, concepts of
geometric modelling may be required to represent these seismicity solids in three-
dimensional Euclidian space [Gallier, 1999] in which the vector 𝑟 is possibly curved
[Mignan, 2011]. The non-stationary background seismicity rate µ(t) is then defined in
the volume of maximum extent r
max
by
𝜇 𝑡 =
𝛿
!!
𝑘𝑟
!"#
!
, 𝑡 < 𝑡
!
𝛿
!!
𝑘(𝑟
!"#
!
𝑟
!
(𝑡)
!
) + 𝛿
!"
𝑘𝑟
!
(𝑡)
!
, 𝑡
!
𝑡 < 𝑡
!
𝛿
!!
𝑘𝑟
!"#
!
, 𝑡
!
𝑡 𝑡
!
𝛿
!!
𝑘(𝑟
!"#
!
𝑟
!
(𝑡)
!
) + 𝛿
!"
𝑘𝑟
!
(𝑡)
!
, 𝑡
!
< 𝑡 < 𝑡
!
(3)
!
6!
with k a geometric parameter and d the spatial dimension. For the tectonic case in
which r
max
>> h, the volume is assumed a cylinder with k = π, d = 2 and
δ
the density
of epicentres in space (Fig. 1c). Finally, the cumulative number of events N(t) is
defined as
𝑁 𝑡 = 𝜇 𝑡 𝑑𝑡
!
!
!
(4)
which represents a power-law time-to-failure equation of the form
𝑁 𝑡 𝑡 + 𝑡
!
!
!!
(5)
the first term representing the linear background seismicity and the second term the
quiescence or activation power-law behaviour observed prior to some large
mainshocks (Fig. 1d) [Sammis and Sornette, 2002].
3. Application of the N-C PAST static stress model to induced seismicity
In the case of an EGS stimulation, the stress source is the fluid injected at
depth with overpressure
𝑃 𝑡, 𝑟 = 0 = 𝐾
!!(!,Δ!)
!
!
(6)
where K is the bulk modulus, ΔV the volume change per time unit and V
0
the
infinitesimal volume subjected to pressure effect per time unit at the borehole located
at r = 0. The injected volume V(t) is determined from the flow rate profile Q(t), as
𝑉 𝑡 = 𝑄 𝑡 𝑑𝑡
!
!
!
(7)
with t
0
the starting time of the injection. The change of volume is then defined as
Δ𝑉 𝑡, Δ𝑡 =
! ! !!(!!!!)
!
(8)
with Δt a time increment.
In the EGS case, r h with h the borehole depth and induced seismicity
defined as hypocentres. The spatiotemporal stress field σ(r,t) becomes
!
7!
𝜎 𝑟, 𝑡 =
𝜎
!
, 𝑡 < 𝑡
!
!
!
!
!!!
!
!
𝑃(𝑡, 𝑟 = 0) + 𝜎
!
, 𝑡 𝑡
!
(9)
with r the distance along the stress field gradient from the borehole, n = 3 the spatial
diffusion exponent for static stress and r
0
0 the infinitesimal radius of volume V
0
=
kr
0
d
/t
0
, t
0
= 1 being the time unit. Activation represents the case when fluids are
injected and quiescence when fluids are ejected (bleed-off). It follows that
𝑟
!
𝑡|Δ𝑉 0 =
!
!
!!!
!
!!
!
!
Δ𝑉(𝑡)
!/!
𝑟
!
𝑟
!
𝑡|Δ𝑉 < 0 =
!
!
!!!
!
!!
!
!
Δ𝑉(𝑡)
!/!
𝑟
!
(10)
which suggests that the spatiotemporal shape of the induced seismicity envelope
depends on the nth-root of the flow rate profile Q(t) (with n = 3 in the static stress
case). This parabolic relationship is similar to the generalized form r(t) m(t)
1/d
derived from nonlinear poroelasticity in a heterogeneous medium where m is the
cumulative mass of injected fluid and d the spatial dimension [Shapiro and Dinske,
2009]. The main difference between the two physical approaches is in the underlying
stress field, which is here static and in poroelasticity, dynamic and related to the
displacement gradient of the fluid mass [Rudnicki, 1986]. It is trivial to derive Eq.
(10) from Eq. (9) while numerous assumptions are necessary to obtain the parabolic
form m(t)
1/d
in nonlinear poroelasticity [Shapiro and Dinske, 2009].
The induced seismicity rate µ(t) is then defined by Eq. (3) but with r
*
from Eq.
(10), k = 4π/3 and d = 3, assuming a spherical spatial volume (i.e. isotropic stress
field). For the activation phase (i.e. stimulation period), it follows that
𝑁 𝑡 Δ𝑉(𝑡)
!
!
!!
(11)
or
𝑁 𝑡 𝑉(𝑡)
!
!
(12)
!
8!
The induced seismicity case d = n = 3 confirms the linear relationship between
cumulative injected volume and cumulative number of induced earthquakes N(t)
V(t) previously derived from poroelasticity [e.g., Shapiro and Dinske, 2009]. In
contrast with poroelasticity, this second law is a direct consequence of the first. The d
= n condition also yields the simplified form of Eq. (10)
𝑟
!
𝑡|Δ𝑉 0
!
!!
!!
!
!
Δ𝑉(𝑡)
!/!
𝑟
!
𝑡|Δ𝑉 < 0
!
!!
!!
!
!
Δ𝑉(𝑡)
!/!
(13)
where the one free parameter is the normalized background stress amplitude range
Δ𝜎
= Δ𝜎
/(𝐾𝑡
!
).
4. Application to the 2006 Basel EGS induced seismicity sequence
Figure 2 shows the flow rate Q(t) of injected fluids during the 2006 Basel EGS
stimulation experiment [Häring et al., 2008] and the spatiotemporal distribution of
relocated induced seismicity [Kraft and Deichmann, 2014] above completeness
magnitude M
c
= 0.8. The injection started at 18:00 on 2 December 2006 (t
0
) and
stopped at 11:33 on 8 December 2006 (t
1
) after which the well was bled-off (ΔV < 0)
(Fig. 2a). The N-C PAST thus predicts an activation envelope r
A
*
for t
0
t < t
1
and a
quiescence envelope r
Q
*
for t t
1
(Eq. 13). The activation and quiescence envelopes
are fitted to the Basel data using Δ𝜎
[10
-3
, 10
-1
] day
-1
(light curves) and Δt = 1/4
day. The results are shown in Figure 2b. The value Δ𝜎
= 0.007 day
-1
(dark curves)
provides the best fit to the data, defined from the best score S = (w
A
+w
Q
)/2 with w
A
and w
Q
the ratio of events of distance r r
A
*
and r r
Q
*
in the injection and bleeding-
off phases, respectively. Figure 2c shows S as a function of Δ𝜎
for Δt = {1/12, 1/8,
1/4} day, which indicates that the results remain stable for lower time increments.
!
9!
Figure 2: 2006 Basel EGS stimulation experiment data with activation and
quiescence envelope fits: (a) Flow rate Q(t) [Häring et al., 2008]; (b) Spatiotemporal
distribution of relocated induced seismicity [Kraft and Deichmann, 2014] with r the
distance from the borehole. The activation and quiescence envelopes r
A
*
(t) and r
Q
*
(t)
are defined from Eq. (13) with parameters Δ𝜎
= 0.007 day
-1
(dark curves) and Δt =
1/4 day. The light curves represent the range Δ𝜎
[10
-3
, 10
-1
] day
-1
in 0.1 increments
!
10!
in the log
10
scale. Points represent the induced earthquakes, which colour indicates
how they are declared; (c) Score S = (w
A
+w
Q
)/2 with w
A
and w
Q
the ratio of events of
distance r r
A
*
and r r
Q
*
in the injection and bleeding-off phases, respectively. The
vertical line represents Δ𝜎
= 0.007 day
-1
.
I evaluate δ
b0
= 10
-10
event/m
3
/day by counting all earthquakes declared in the
national Swiss catalogue (ECOS-09
1
) and located within 10 km of the borehole of
coordinates (7.594°E; 47.586°N) and depth 4.36 km. It means that ~1 tectonic
earthquake is expected in average in the space-time window considered. Due to the
low tectonic activity in the area, I approximate δ
b0
= δ
bm
= 0 event/m
3
/day (i.e., total
quiescence). The theory shows a good agreement with the observations with 97% of
the seismicity below r
A
*
during the injection phase (red points in Fig. 2b) and 98% of
the seismicity above r
Q
*
during the bleeding-off phase (orange to yellow points).
The density of events above r
Q
*
is however not δ
b0
but
𝛿
!
𝑡 𝑡
!
= 𝛿
!"
exp
!!!
!
!
(14)
which represents the temporal diffusion of induced seismicity with τ the average time
constant. Eq. (14) represents a relaxation process from the overloading state to the
background state. The results here suggest that only the events declared as
background (grey points) and quiescence events (blue points) are outliers. The
observed variations in r below r
A
*
and above r
Q
*
are not explained by the model,
which only predicts the behaviour of the activation and quiescence fronts. The
second-order variations may be due to anisotropic effects and for t > t(max(r
A
*
)) to
additional spatial diffusion effects.
!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!
1
!http://hitseddb.ethz.ch:8080/ecos09/!
!
11!
Figure 3 shows the 6-hour rate of induced seismicity µ(t) and the cumulative
number of induced events N(t), observed and predicted. With δ
b0
= δ
bm
= 0 and taking
into account induced seismicity temporal diffusion, the rate of induced seismicity
becomes
𝜇 𝑡 = max
!!
!
𝛿
!"
. Δ𝑡. 𝑟
(𝑡)
!
,
!!
!
𝛿
!"
. Δ𝑡. 𝑟
(𝑡𝑆
!
)
!
exp!(
!!!
!
!
) (15)
where δ
bp
= 4.68 10
-7
event/m
3
/day (production parameter) and τ = 1.18 day (diffusion
parameter) are obtained by maximum-likelihood estimation (MLE), set S
t
= {Δt, …,
iΔt, …} and
𝑟
𝑡 =
0 , 𝑡 < 𝑡
!
𝑟
!
(𝑡) , 𝑡
!
𝑡 < 𝑡
!
0 , 𝑡 𝑡
!
(16)
Eq. (15) infers that induced seismicity is fully explained by overloading, in agreement
with the observation of no causal relationships between events in the Basel sequence
[Langenbruch et al., 2011]. The predicted rate (Eq. 15) and predicted cumulative
number of events (Eq. 4) fit the data well, as shown in Figures 3a and 3b,
respectively. The role of temporal diffusion is observed after t
1
-Δt and is the only
contributor to induced seismicity after t
1
. Of three functional forms tested to describe
diffusion (exponential, stretched exponential and power law), the exponential (Eq. 14)
was verified to be the best model for the Basel case (following the formalism and tests
proposed by Clauset et al. [2009]).
5. Conclusions
I have demonstrated that the two principal induced seismicity descriptive laws
can be explained from simple geometric operations in a static stress field without
requiring any concept derived from poroelasticity. The two descriptive laws had been
previously obtained by considering the differential equations of poroelasticity [Biot,
!
12!
1941; Rudnicki, 1986] under different assumptions [Shapiro and Dinske, 2009],
which indicates that the static stress model defined from algebraic expressions
requires a lower description length [Kolmogorov, 1965]. This is crudely inferred here
from the difference between the lengths of the present demonstration and of published
poroelasticity demonstrations [e.g., Shapiro and Dinske, 2009].
Figure 3: Induced seismicity production time series, observed and predicted: (a)
Histogram of the observed 6-hour induced seismicity rate µ(t) with fit based on Eq.
(15) with MLE parameters δ
bp
= 4.68 10
-7
event/m
3
/day (production parameter) and τ
= 1.18 day (diffusion parameter); (b) Cumulative number of induced earthquakes N(t)
with fit based on Eq. (4) with µ(t) of Eq. (15).
!
13!
I also showed that the controlling parameter is the normalized background
stress amplitude range Δ𝜎
, which questions the usefulness of permeability and
diffusivity parameters in induced seismicity analyses and might explain why these
parameters remain elusive [Miller, 2015]. In that view, permeability could depend on
the “external loading configuration” instead of on the material itself, as recently
proposed in the case of the static friction coefficient [Ben-David and Fineberg, 2013].
Testing of the model on other induced seismicity sequences will determine if Δ𝜎
is
itself universal, region-specific or related to the static stress memory of the crust,
hence if Δ𝜎
depends or not on the tectonic loading configuration at EGS natural
laboratory sites. Similar questions apply to the earthquake production parameter δ
bp
and if the two parameters are independent or correlated.
The main assumption of the N-C PAST is to consider three unique seismicity
regimes (quiescence, background and activation) defined by the event productions δ
bm
< δ
b0
< δ
bp
. There are two possible physical alternatives to justify this choice: (1) it
represents the fundamental behaviour of the Earth crust, which would hence act as a
capacitor, with strain energy storage and δ
bp
analogues to electrical energy storage
and capacitance, respectively; (2) the proposed step function is a simplification of the
true stress-production profile, which remains unknown and is so far best characterized
by three regimes [e.g., King, 2007]. Both alternatives allow defining spatiotemporal
solids over which geometrical operations yield algebraic expressions of the induced
seismicity behaviour.
!
14!
References
Bak, P., and C. Tang (1989), Earthquakes as a Self-Organized Critical Phenomenon,
J. Geophys. Res., 94, 15635-15637
Ben-David, O., and J. Fineberg (2011), Static Friction Coefficient Is Not a Material
Constant, Phys. Rev. Lett., 106, 254301, doi: 10.1103/PhysRevLett.106.254301
Berryman, J. G., and H. F. Wang (2001), Dispersion in poroelastic systems, Phys.
Rev. E, 64, 011303, doi: 10.1103/PhysRevE.64.011303
Biot, M. A. (1941), General Theory of Three Dimensional Consolidation, J. Appl.
Phys., 12, 155-164, doi: 10.1063/1.1712886
Clauset, A., C. R. Shalizi, and M. E. J. Newman (2009), Power-Law Distributions in
Empirical Data, SIAM Review, 51, 661-703, doi: 10.1137/070710111
Ellsworth, W. L. (2013), Injection-induced Earthquakes, Science, 341, doi:
10.1126/science.1225942
Gallier, J. (1999), Curves and surfaces in geometric modeling, theory and algorithms,
Morgan Kaufmann, 491 pp.
Grasso, J.-R., and D. Sornette (1998), Testing self-organized criticality by induced
seismicity, J. Geophys. Res., 103, 29965-29987
Häring, M. O., U. Schanz, F. Ladner, and B. C. Dyer (2008), Characterisation of the
Basel 1 enhanced geothermal system, Geothermics, 37, 469-495, doi:
10.1016/j.geothermics.2008.06.002
King, G. C. P. (2007), Fault interaction, Earthquake Stress Changes, and the
Evolution of Seismicity, Treatise on Geophysics 4, Earthquake Seismology, G
Schubert (Ed.), Elsevier, Amsterdam, 225-255
Kolmogorov, A. N. (1965), Three approaches to the quantitative definition of
information, Problemy Peredachi Informatsii, 1, 3-11
!
15!
Kraft, T. and N. Deichmann (2014), High-precision relocation and focal mechanism
of the injection-induced seismicity at the Basel EGS, Geothermics, 52, 59-73, doi:
10.1016/j.geothermics.2014.05.014
Langenbruch, C., C. Dinske, and S. A. Shapiro (2011), Inter event times of fluid
induced earthquakes suggest their Poisson nature, Geophys. Res. Lett., 38, L21302,
doi: 10.1029/2011GL049474
Majer, E. L., R. Baria, M. Stark, S. Oates, J. Bommer, B. Smith, and H. Asanuma
(2007), Induced seismicity associated with Enhanced Geothermal Systems,
Geothermics, 36, 185-222, doi: 10.1016/j.geothermics.2007.03.003
Mignan, A., G. C. P. King, and D. Bowman (2007), A mathematical formulation of
accelerating moment release based on the stress accumulation model, J. Geophys.
Res., 112, B07308, doi: 10.1029/2006JB004671
Mignan, A. (2011), Retrospective on the Accelerating Seismic Release (ASR)
hypothesis: Controversy and new horizons, Tectonophysics, 505, 1-16, doi:
10.1016/j.tecto.2011.03.010
Mignan, A. (2012), Seismicity precursors to large earthquakes unified in a stress
accumulation framework, Geophys. Res. Lett., 39, L21308, doi:
10.1029/2012GL053946
Mignan, A., D. Landtwing, P. Kästli, B. Mena and S. Wiemer (2015), Induced
seismicity risk analysis of the 2006 Basel, Switzerland, Enhanced Geothermal
System project: Influence of uncertainties on risk mitigation, Geothermics, 53,
133-146, doi: 10.1016/j.geothermics.2014.05.007
Miller, S. A. (2015), Modeling enhanced geothermal systems and the essential nature
of large-scale changes in permeability at the onset of slip, Geofluids, 15, 338-349,
doi: 10.1111/gfl.12108
!
16!
Rudnicki, J. W. (1986), Fluid mass sources and point forces in linear elastic diffusive
solids, Mechanics of Materials, 5, 383-393
Rutqvist, J. (2011), Status of the TOUGH-FLAC simulator and recent applications
related to coupled fluid flow and crustal deformations, Computers & Geosciences,
37, 739-750, doi: 10.1016/j.cageo.2010.08.006
Sammis, C. G., and D. Sornette (2002), Positive feedback, memory, and the
predictability of earthquakes, PNAS, 99, 2501-2508, doi: 10.1073/pnas.012580999
Shapiro, S. A., and C. Dinske (2009), Scaling of seismicity induced by nonlinear luid-
rock interaction, J. Geophys. Res., 114, B09307, doi: 10.1029/2008JB006145
... Typically, fluid-induced seismicity is characterised by a time-varying seismicity rate related to the fluid injection rate [see, e.g. Shapiro et al., 2010;Dinske and Shapiro, 2013;Mignan, 2016;Mignan et al., 2017;van der Elst et al., 2016;Zoback, 2016, 2017]. Although the seismicity rate changes over time, the inter-arrival times between seismic events have been shown to be statistically independent [Langenbruch et al., 2011]. ...
... we use a f b as a generic statistical parameter with no preference for any underlying physical model [Mignan et al., 2017]. The post-injection phase identifies the phase with constant null flow rate, after the injection has been terminated, and is characterised by an exponential decay typical of a diffusion process [Mignan, 2015[Mignan, , 2016Mignan et al., 2017]. Although the Modified Omori Law is sometimes used to describe post-injection seismicity Barth et al., 2013], Mignan et al. [2017] have shown that an exponential function performs better than a power law for the six datasets presented in the Supplementary ...
... injection phase. The injection phase admits only positive fluid injection rates, and it is characterised by a linear relationship between V(t s ) and λ(t) (in line withShapiro et al. [2010];Dinske and Shapiro [2013];Hajati et al. [2015]; van der Elst et al.[2016];Mignan [2016]; ...
Article
Full-text available
In this study, we present a Bayesian hierarchical framework to model fluid-induced seismicity. The framework is based on a non-homogeneous Poisson process (NHPP) with a fluid-induced seismicity rate proportional to the rate of injected fluid. The fluid-induced seismicity rate model depends upon a set of physically meaningful parameters, and has been validated for six fluid-induced case studies. In line with the vision of hierarchical Bayesian modeling, the rate parameters are considered as random variables. We develop both the Bayesian inference and updating rules, which are used to develop a probabilistic forecasting model. We tested the Basel 2006 fluid-induced seismic case study to prove that the hierarchical Bayesian model offers a suitable framework to coherently encode both epistemic uncertainty and aleatory variability. Moreover, it provides a robust and consistent short-term seismic forecasting model suitable for online risk quantification and mitigation. Link to the article preview: http://onlinelibrary.wiley.com/doi/10.1002/2017GL075251/full
... A predictive model lies at the heart of any risk assessment. In the case of induced seismicity, a wide range of statistical and physics-based models exists 16,17 . Undoubtedly, more work is needed to develop, calibrate and validate new models; however, we believe that the missing link is the use of such models for deriving and monitoring quantitative risk thresholds. ...
... Both a fb and b can also be functions of time t (see Decision variable section). In this model, the injection or operation phase is described by a linear relationship between λ(t, m ≥ m 0 ) and  V (t) in line with previous studies [17][18][19] . It derives directly from the linear relationship between  V and overpressure 17 , hence assuming no change of injectivity during any given stimulation. ...
... In this model, the injection or operation phase is described by a linear relationship between λ(t, m ≥ m 0 ) and  V (t) in line with previous studies [17][18][19] . It derives directly from the linear relationship between  V and overpressure 17 , hence assuming no change of injectivity during any given stimulation. The post-injection phase is described by a pure exponential decay representative of a normal diffusion process 17 . ...
Article
Full-text available
The rise in the frequency of anthropogenic earthquakes due to deep fluid injections is posing serious economic, societal, and legal challenges to many geo-energy and waste-disposal projects. Existing tools to assess such problems are still inherently heuristic and mostly based on expert elicitation (so-called clinical judgment). We propose, as a complementary approach, an adaptive traffic light system (ATLS) that is function of a statistical model of induced seismicity. It offers an actuarial judgement of the risk, which is based on a mapping between earthquake magnitude and risk. Using data from six underground reservoir stimulation experiments, mostly from Enhanced Geothermal Systems, we illustrate how such a data-driven adaptive forecasting system could guarantee a risk-based safety target. The proposed model, which includes a linear relationship between seismicity rate and flow rate, as well as a normal diffusion process for post-injection, is first confirmed to be representative of the data. Being integrable, the model yields a closed-form ATLS solution that is both transparent and robust. Although simulations verify that the safety target is consistently ensured when the ATLS is applied, the model from which simulations are generated is validated on a limited dataset, hence still requiring further tests in additional fluid injection environments.
... The aim of the present article is to explain the Utsu aftershock productivity equation (Eq. 1) by applying a geometrical theory of seismicity (or " Solid Seismicity " ), which has already been shown to effectively explain other empirical laws of both natural and induced seismicity from simple geometric operations on a permanent static stress field (Mignan, 2012;2016). The theory is applied here for the first time to the case of aftershocks. ...
... The power of Eq. (5) is that it allows defining seismicity patterns in terms of " solids " described by the spatial envelope í±Ÿ * = í±Ÿ í¼Ž = ±Δí¼Š * . The spatiotemporal rate of seismicity is then a mathematical expression defined by the density of events δ times the volume characterized by í±Ÿ * (see previous demonstrations inMignan et al., 2007;Mignan, 2011;2012;2016where simple algebraic expressions were obtained). In the case of aftershocks, we define the static stress field of the mainshock by ...
... 1). This demonstration, combined to the previous ones made by the author to explain precursory accelerating seismicity and induced seismicity (Mignan, 2012;2016), suggests that most empirical laws observed in seismicity populations can be explained by simple geometric operations on a permanent static stress field. Although the Solid Seismicity Postulate (SSP) (Eq. ...
Article
Full-text available
The aftershock productivity law, first described by Utsu in 1970, is an exponential function of the form K=K0.exp({\alpha}M) where K is the number of aftershocks, M the mainshock magnitude, and {\alpha} the productivity parameter. The Utsu law remains empirical in nature although it has also been retrieved in static stress simulations. Here, we explain this law based on Solid Seismicity, a geometrical theory of seismicity where seismicity patterns are described by mathematical expressions obtained from geometric operations on a permanent static stress field. We recover the exponential form but with a break in scaling predicted between small and large magnitudes M, with {\alpha}=1.5ln(10) and ln(10), respectively, in agreement with results from previous static stress simulations. We suggest that the lack of break in scaling observed in seismicity catalogues (with {\alpha}=ln(10)) could be an artefact from existing aftershock selection methods, which assume a continuous behavior over the full magnitude range. While the possibility for such an artefact is verified in simulations, the existence of the theoretical kink remains to be proven.
... Deep fluid injection is a process used in different industrial contexts, as for example deep geothermal operations, conventional and non-conventional hydrocarbon extraction, as well as gas storage, CO2 sequestration, salt solution mining, sequestration of wastewater, etc. These activities can generate a seismicity provoked by the increase of deep fluid pressure (Mignan, 2016). As before, two types of seismicity can be observed, one, induced, directly linked to the injection, and the other, triggered, provoked by the remobilisation of a preexisting fault. ...
Book
Full-text available
Since the industrial revolution, global demand for fossil energy and raw materials has led to a considerable increase in the exploitation of underground resources. The demand for mineral resources even tends to accelerate: since the beginning of the last century to the first decade of 2000, the World extraction of raw materials has increased by 10. The part of mineral resources increased by a quarter to two thirds of the total and the OECD estimates that by 2030 the world production of raw materials could be multiplied by 15. Conventional and unconventional exploitation of hydrocarbons, energy storage, deep geothermal energy, geological sequestration of CO2 ... technological advances lead to constantly pushing back the limits of the exploited deep underground. However, the multiplicity, diversity and increasing scale of exploitation projects, as well as the new uses of the underground, increase the risk that these industrial activities may generate earthquakes, called "anthropogenic", that is, generated by man (as opposed to earthquakes of natural origin). Ineris's work on the know ledge review of anthropogenic seismicity was based on the identification and analysis of 260 case studies. Compared to natural seismicity, anthropogenic seismicity has certain specificities that should be considered at all stages of risk prevention. The Ineris report provides an update on hazard assessment methods, namely the probability that a seismic event of a given magnitude will occur at a given location. The different types of risks generated by anthropogenic seismicity are also analysed, as well as the main measures of prevention and mitigation of these phenomena.
... Physics-based models (e.g., Olivella et al., 1994;Bruel, 2005;Kohl and Megel, 2007;Baisch et al., 2010;Rinaldi et al., 2015;McClure and Horne, 2012;Wang and Ghassemi, 2012;Karvounis and Wiemer, 2015) do consider some underlying physical processes. But physicsbased models are currently too computationally demanding (Mignan, 2015), and often have too many free parameters to be robust and ready for real-time applications. Hybrid models are a compromise between forecast capabilities of physical models and computational efficiency of statistical models. ...
Technical Report
Full-text available
The Swiss Seismological Service at ETH Zurich (SED) wants to contribute to a sustainable and safe use of deep geothermal energy with this report on good practice on monitoring, assessment and management of induced seismicity related to the exploitation of deep geothermal energy. The report is aimed at a range of audiences: field operators, regulators at local, cantonal or federal level, insurance companies looking to assess the financial risk, as well as media and the general public that wish to be informed on the topic. In the good practice recommendations outlined in this report, we follow the concept of risk governance that considers “the totality of actors, rules, conventions, processes, and mechanisms concerned with how relevant risk information is collected, analysed, and communicated and how management decisions are taken” (IRGC, 2005, p. 80). Successful risk governance is an analytical-deliberative process in which field operators, independent risk analysts, regulators and stakeholders collaborate in managing risks risks (Trutnevyte & Wiemer, 2017). An important element of risk governance is that risks are considered consistently and transparently throughout all phases of the project. In this spirit, this good practice covers the initial phase of the project planning all the way through to the post-operation phase. In order to identify suitable workflows or risk governance processes, we recommend an initial evaluation of each project following the Geothermal Risk of Induced seismicity Diagnosis (GRID) approach, proposed by Trutnevyte & Wiemer (2017) for open or partly open geothermal systems. The GRID approach is based on a series of indicators that describe concern about induced seismicity hazard, risk, and social context. These indicators are generally available before the project is initiated, before the risk study is commissioned and before the first well is drilled. This good practice guide applies to essentially all current and future deep geothermal energy projects in Switzerland that inject or extract fluids from the deep underground (hereafter referred to as "open systems"). This excludes closed systems used for heat pumps, where no risk of inducing earthquakes exists.
... Brodsky & Lajoie 2013). Such relation has been theoretically formulated and established by the non-linear pore pressure diffusion concept (Shapiro & Dinske 2009), or alternatively, by the application of a static stress model based on the non-critical precursory accelerating seismicity theory (Mignan 2016) to induced seismicity. ...
Article
Full-text available
The Geysers geothermal field located in California, USA, is the largest geothermal site in the world, operating since the 1960΄s. We here investigate and quantify the correlation between temporal seismicity evolution and variation of the injection data by examination of time series through specified statistical tools (binomial test to investigate significant rate changes, cross correlation between seismic and injection data, b-value variation analysis). To do so, we utilize seismicity and operational data associated with 2 injection wells (Prati-9 and Prati-29) which cover a time period of approximately 7 years (from November 2007 to August 2014). The seismicity is found to be significantly positively correlated with the injection rate. The maximum correlation occurs with a seismic response delay of ∼2 weeks, following injection operations. Those results are very stable even after considering hypocentral uncertainties, by applying a vertical shift of the events foci up to 300 m. Our analysis indicates also time-variations of b-value, which exhibits significant positive correlation with injection rates.
Article
Full-text available
The Hengill geothermal field, located in southwest Iceland, is host to the Hellisheiði power plant, with its 40+ production wells and 17 reinjection wells. Located in a tectonically active area, the field experiences both natural and induced seismicity linked to the power plant operations. To better manage the risk posed by this seismicity, the development of robust and informative forecasting models is paramount. In this study, we compare the forecasting performance of a model developed for fluid‐induced seismicity (the Seismogenic Index model) and a class of well‐established statistical models (Epidemic‐Type Aftershock Sequence). The pseudo‐prospective experiment is set up with 14 months of initial calibration and daily forecasts for a year. In the timeframe of this experiment, a dense broadband network was in place in Hengill, allowing us to rely on a high quality relocated seismic catalog. The seismicity in the geothermal field is characterized by four main clusters, associated with the two reinjection areas, one production area, and an area with surface geothermal manifestations but where no operations are taking place. We show that the models are generally well suited to forecast induced seismicity, despite some limitations, and that a hybrid ETAS model accounting for fluid forcing has some potential in complex regions with natural and fluid‐induced seismicity.
Article
Full-text available
The seismic risk associated with deep fluid injection in Enhanced Geothermal Systems can be mitigated by stopping reservoir stimulation when the seismic risk becomes unacceptable or by reducing production flow rates when seismicity occurs during the operational phase. So far, none of these mitigation measures have been included in the Levelized Cost of Electricity. A meta-model is introduced that estimates the optimal price of electricity, based on an analytical geothermal energy model, and updates this cost to include the outlay for mandatory seismic risk mitigation measures. The proposed energy model computes both electricity production and heat credit. The costs added during reservoir stimulation are based on the probability of abandoning an injection well, based on a traffic-light system, defined as the ratio of scenarios that exceed a given seismic safety threshold in the risk space. In the production phase, the net energy generated is reduced by clipping the production flow rate so that the reservoir's overpressure does not exceed the regional minimum effective stress. Based on a generic geothermal triplet, we investigate the trade-off between heat credit and seismic risk mitigation cost. The added cost, mostly due to financial risk aversion, shifts the optimal site for a plant from between a few kilometers to tens of kilometers away from populated areas, for increasingly vulnerable building stocks.
Thesis
Full-text available
Recent years have seen a growing emphasis on developing models for earthquake forecasting and quantifying their predictive skills in order to augment our understanding of the processes leading to seismogenesis. Two extensively used models to describe the spatial and temporal distribution of earthquakes are Rate and State and Epidemic Type Aftershock Sequence model. They are respectively based on the static stress triggering hypothesis and empirically observed statistical laws (namely Omori law, Gutenberg Richter law and Productivity law). Both models have been successful to some extent in describing many empirical observations related to seismicity as well as in forecasting the rate of the future earthquakes. However, several key issues related to both remain unaddressed. For instance, (i) impact of secondary stress changes caused by small earthquakes is completely ignored while assessing the static stress triggering hypothesis as well as in forecasting the rate of future earthquakes based on Coulomb stress change from the past events; (ii) often simplistic assumptions are made about the orientation of the fault planes that cause static deformations as well as the recipient fault planes of these deformations; (iii) influence of uncertainties in the data are rarely considered while computing the static stress changes; (iv) parameters of the ETAS model are often assumed to be spatially homogeneous; (v) very few attempts are made to understand the physical origin of the ETAS parameters, and (vi) possibility of a self-consistent hybrid of the two models remains unexplored.We attempt to address these issues in a systematic manner with attention on statistical rigor, realistic synthetic tests and generalized data-driven approaches. We first conduct a rigorous test of the static stress triggering hypothesis with the aim to tackle the first three points outlined above. In particular, we investigate the correlation between the time variation of the seismicity rate and the sign (and amplitude) of Coulomb stress changes. We also quantify the Coulomb Index (CI), the fraction of events that received net positive Coulomb stress changes compared to the total number of events. We find compelling evidence supporting the static triggering (with stronger evidence after resolving the focal plane ambiguity) above significantly small (about 10 Pa) but consistently observed stress thresholds. Furthermore, we find evidence for the static triggering hypothesis to be robust with respect to the choice of the friction coefficient, Skempton’s coefficient, and magnitude threshold. However, our analysis suffers from the imprecise assumption that all earthquakes following a source event are its direct aftershocks. To ameliorate the effects of the preceding assumption on our analysis, we infer the full triggering genealogy of the ANSS catalog of the Californian earthquakes by defining and implementing an ETAS model with space varying parameters using the Expectation Maximization (EM) algorithm and spatial Voronoi tessellation ensembles. We use the penalized Log-Likelihood to rank the inverted models and select the best ones to eventually compute an ensemble model at any location. Prior to analyzing the results obtained from the application of the proposed method to earthquakes in and around California, we verified the reliability of this method using realistic synthetic catalogs. Results obtained on this earthquake catalog suggest that ETAS (productivity and background) parameters are far from uniform in the study region. We also find that the efficiency of earthquakes to trigger future ones (quantified by the branching ratio) positively correlates with the local surface heat flow measurements. In contrast, the rate of earthquakes triggered by far-field tectonic loading (or background seismicity rate) shows no such correlation, suggesting the relevance of dynamic triggering possibly through fluid-induced activation. Furthermore, we find the branching ratio and background seismicity rate to be uncorrelated with hypocentral depths, indicating that the seismic coupling remains invariant of hypocentral depths in the study region. Additionally, we find pieces of evidence suggesting that the earthquake triggering is mostly dominated by small earthquakes, convincing us further that the static stress change studies should not only focus on the Coulomb stress changes caused by specific moderate to large earthquakes but should also account for the secondary static stress changes caused by smaller events. We then quantify the forecasting skill of the ETAS model with spatially varying parameters by comparing the former to the forecasting skills of other smoothed seismicity models. Our results indicate that by taking account of both background and aftershock components of the seismicity rate, our model (ETAS-SV2) easily outperforms models that are based on just smoothing the declustered seismicity. Furthermore, we also find that when forecasting the rate of future M4.95 earthquakes, the smoothed seismicity model based only on a catalog that has been stochastically declustered using the ETAS model tends to outperform smoothed seismicity models based on declustered catalogs obtained from two widely used declustering methods. We then use the genealogy tree of earthquakes obtained from the spatially heterogeneous ETAS model to quantify the compatibility of the latter with the static stress triggering hypothesis. Using this tree, we obtain conditional Coulomb Indices (CI values conditioned on the probability of being a direct aftershock) for different source models available for the Landers earthquake. These conditional CI values are then compared to previously defined unconditional CI values, which serve as the null hypothesis. We find statistically significant enrichment in the evidence supporting the static triggering hypothesis when using the genealogy tree obtained from the improved ETAS model pointing towards at least partial compatibility of the former with the static stress triggering hypothesis. Finally, we estimate time-varying ETAS parameters in the Salton Sea geothermal area and Geysers geothermal field. Our observations suggest that the main fault segment in the Salton Sea geothermal area is more favorably oriented to the underlying background stress field than its remaining three subsidiary conjugate fault segments. Our results also allow us to evidence the influence of fluid injection (and/or extraction) on the background seismicity rate in the regions located around the Salton Sea geothermal facility and the Geysers geothermal facility. While in the case of the former we are not able to deduce if the fluid injection or extraction drives the human-induced seismic activity, our results clearly indicate that the fluid injection rate seems to dominate the seismicity triggering in the case of the latter. --> Recent years have seen a growing emphasis on developing models for earthquake forecastingand quantifying their predictive skills in order to augment our understandingof the processes leading to seismogenesis. Two extensively used models to describethe spatial and temporal distribution of earthquakes are Rate and State and Epidemic Type AftershockSequence model. They are respectively based on the static stress triggering hypothesisand empirically observed statistical laws (namely Omori law, Gutenberg Richter law and Productivitylaw). Both models have been successful to some extent in describing many empiricalobservations related to seismicity as well as in forecasting the rate of the future earthquakes.However, several key issues related to both remain unaddressed. For instance, (i) impact of secondarystress changes caused by small earthquakes is completely ignored while assessing thestatic stress triggering hypothesis as well as in forecasting the rate of future earthquakes basedon Coulomb stress change from the past events; (ii) often simplistic assumptions are made aboutthe orientation of the fault planes that cause static deformations as well as the recipient faultplanes of these deformations; (iii) influence of uncertainties in the data are rarely consideredwhile computing the static stress changes; (iv) parameters of the ETAS model are often assumedto be spatially homogeneous; (v) very few attempts are made to understand the physicalorigin of the ETAS parameters, and (vi) possibility of a self-consistent hybrid of the two modelsremains unexplored.We attempt to address these issues in a systematic manner with attentionon statistical rigor, realistic synthetic tests and generalized data-driven approaches.We first conduct a rigorous test of the static stress triggering hypothesis with the aim totackle the first three points outlined above. In particular, we investigate the correlation betweenthe time variation of the seismicity rate and the sign (and amplitude) of Coulomb stresschanges. We also quantify the Coulomb Index (CI), the fraction of events that received net positiveCoulomb stress changes compared to the total number of events. We find compelling evidencesupporting the static triggering (with stronger evidence after resolving the focal planeambiguity) above significantly small (about 10 Pa) but consistently observed stress thresholds.Furthermore, we find evidence for the static triggering hypothesis to be robust with respect tothe choice of the friction coefficient, Skempton’s coefficient, and magnitude threshold. However,our analysis suffers from the imprecise assumption that all earthquakes following a source eventare its direct aftershocks.To ameliorate the effects of the preceding assumption on our analysis, we infer the full triggeringgenealogy of the ANSS catalog of the Californian earthquakes by defining and implementingan ETAS model with space varying parameters using the Expectation Maximization (EM)algorithm and spatial Voronoi tessellation ensembles. We use the penalized Log-Likelihood torank the inverted models and select the best ones to eventually compute an ensemble model atany location. Prior to analyzing the results obtained from the application of the proposed methodto earthquakes in and around California, we verified the reliability of this method using realisticsynthetic catalogs. Results obtained on this earthquake catalog suggest that ETAS (productivityand background) parameters are far from uniform in the study region. We also find that theefficiency of earthquakes to trigger future ones (quantified by the branching ratio) positivelycorrelates with the local surface heat flow measurements. In contrast, the rate of earthquakestriggered by far-field tectonic loading (or background seismicity rate) shows no such correlation,suggesting the relevance of dynamic triggering possibly through fluid-induced activation.Furthermore, we find the branching ratio and background seismicity rate to be uncorrelatedwith hypocentral depths, indicating that the seismic coupling remains invariant of hypocentraldepths in the study region. Additionally, we find pieces of evidence suggesting that the earthquake triggering is mostly dominated by small earthquakes, convincing us further thatthe static stress change studies should not only focus on the Coulomb stress changes caused by specific moderate to large earthquakes but should also account for the secondary static stresschanges caused by smaller events.We then quantify the forecasting skill of the ETAS model with spatially varying parametersby comparing the former to the forecasting skills of other smoothed seismicity models. Ourresults indicate that by taking account of both background and aftershock components of theseismicity rate, our model (ETAS-SV2) easily outperforms models that are based on just smoothingthe declustered seismicity. Furthermore, we also find that when forecasting the rate of futureM4.95 earthquakes, the smoothed seismicity model based only on a catalog that has beenstochastically declustered using the ETAS model tends to outperform smoothed seismicity modelsbased on declustered catalogs obtained from two widely used declustering methods.We then use the genealogy tree of earthquakes obtained from the spatially heterogeneousETAS model to quantify the compatibility of the latter with the static stress triggering hypothesis.Using this tree, we obtain conditional Coulomb Indices (CI values conditioned on theprobability of being a direct aftershock) for different source models available for the Landersearthquake. These conditional CI values are then compared to previously defined unconditional values, which serve as the null hypothesis. We find statistically significant enrichment in the evidencesupporting the static triggering hypothesis when using the genealogy tree obtained fromthe improved ETAS model pointing towards at least partial compatibility of the former with thestatic stress triggering hypothesis.Finally, we estimate time-varying ETAS parameters in the Salton Sea geothermal area andGeysers geothermal field. Our observations suggest that the main fault segment in the SaltonSea geothermal area is more favorably oriented to the underlying background stress field thanits remaining three subsidiary conjugate fault segments. Our results also allow us to evidencethe influence of fluid injection (and/or extraction) on the background seismicity rate in the regions located around the Salton Sea geothermal facility and the Geysers geothermal facility.While in the case of the former we are not able to deduce if the fluid injection or extraction drives the human-induced seismic activity, our results clearly indicate that the fluid injection rate seems to dominate the seismicity triggering in the case of the latter
Article
Full-text available
Large earthquakes can be preceded by a period of accelerating seismic activity of moderate-sized earthquakes. This phenomenon, usually termed accelerating moment release, has yet to be clearly understood. A new mathematical formulation of accelerating moment release is obtained from simple stress transfer considerations, following the recently proposed stress accumulation model. This model, based on the concept of elastic rebound, simulates accelerating seismicity from theoretical stress changes during an idealized seismic cycle. In this view, accelerating moment release is simply the consequence of the decrease, due to loading, of the size of a stress shadow due to a previous earthquake. We show that a power law time-to-failure equation can be expressed as a function of the loading rate on the fault that is going to rupture. We also show that the m value, which is the power law exponent, can be defined as m = D/3, with D a parameter that takes into account the geometrical shape of the stress lobes and the distribution of active faults. In the stress accumulation model, the power law is not due to critical processes.
Article
The decay rate of aftershocks has been modeled as a power law since the pioneering work of Omori in the late nineteenth century. Considered the second most fundamental empirical law after the Gutenberg-Richter relationship, the power law paradigm has rarely been challenged by the seismological community. By taking a view of aftershock research not biased by prior conceptions of Omori power law decay and by applying statistical methods recommended in applied mathematics, I show that all aftershock sequences tested in three regional earthquake catalogs (Southern and Northern California, Taiwan) and with three declustering techniques (nearest-neighbor, second-order moment, window methods) follow a stretched exponential instead of a power law. These results infer that aftershocks are due to a simpler relaxation process than originally thought, in accordance with most other relaxation processes observed in Nature.
Article
The permeability structure resulting from high fluid pressure stimulation of a geothermal resource is the most important parameter controlling the feasibility and the viability of enhanced geothermal systems (EGS), yet is the most elusive to constrain. Linear diffusion models do a reasonably good job of constraining the front of the stimulated region because of the t1/2 dependence of the perturbation length, but triggering pressures resulting from such models, and the permeability inferred using the diffusivity parameter, drastically underestimate both permeability and pressure changes. This leads to incorrect interpretations about the nature of the system, including the degree of fluid pressures needed to induce seismicity required to enhance the system. Here, I use a minimalist approach to modeling and show that all of the observations from Basel (Switzerland) fluid injection experiment are well matched by a simple model where the dominant control on the system is a large-scale change in permeability at the onset of slip. The excellent agreement between observations and these simplest of models indicates that these systems may be less complicated than envisaged, thus offering strategies for more sophisticated future modeling to help constrain and exploit these systems.
Article
In early December 2006, a massive fluid injection was carried out at 5 km depth below the city of Basel, Switzerland, for geothermal reservoir enhancement. During the six-day stimulation, approximately 13,000 induced microearthquakes were detected by a borehole network. The largest of the induced earthquakes, which had a magnitude of ML 3.4, was strongly felt in the Basel area and led to the termination of the project after only six days of stimulation. We analyzed the approximately 3500 locatable events of this induced earthquake sequence, which is one of the most densely monitored deep fluid-injections in the world. The seismic monitoring system consisted of six borehole seismometers at depths between 300 m and 2700 m near the injection well and of numerous surface stations in the Basel area. In this article, we report on the analysis of the sequence using exclusively data from the down-hole instruments. We show how a refinement of arrival time picks by cross-correlation techniques and subsequent high-precision relocations lead to significant improvements of the hypocenter locations compared to routinely applied manual procedures. We also analyze focal mechanisms determined from both first-motion polarities and amplitudes of signals recorded by the borehole sensors alone and compare the results to the focal mechanisms of the larger events recorded also by the surface networks. Our findings indicate that the induced sequence consists to a large part of repeating earthquakes that repeatedly rupture the same fault patches, and that the activated seismogenic structure is not a single fault plane but a heterogeneous fault zone.
Article
We present a probabilistic seismic risk analysis of the 2006 Basel Enhanced Geothermal System (EGS) experiment. We combine induced seismicity time-dependent hazard with the RISK-UE macroseismic method and propose a logic tree approach to capture epistemic uncertainties. We find that the expected losses vary over several orders of magnitude for the tested parameters. It indicates that the previous Basel EGS seismic risk study (SERIANEX), which did not include epistemic uncertainties, led to subjective estimates. We address the issue of decision-making under uncertainty by discussing the role of model ambiguity in a simple traffic light system for EGS seismic risk mitigation.
Article
Various seismicity patterns before large earthquakes have been reported in the literature. They include foreshocks (medium-term acceleration and short-term activation), quiescence, doughnut patterns and event migration. The existence of these precursory patterns is however debated. Here, we develop an approach based on the concept of stress accumulation to unify and categorize all claimed seismic precursors in a same physical framework. We first extend the Non-Critical Precursory Accelerating Seismicity Theory (N-C PAST), which already explains most precursors, to additionally include short-term activation. Theoretical results are then compared to the time series observed prior to the 2009 Mw = 6.3 L'Aquila, Italy, earthquake. We finally show that different precursory paths are possible before large earthquakes, with possible coupling of different patterns or non-occurrence of any. This is described by a logic tree defined from the combined probabilities of occurrence of the mainshock at a given stress state and of precursory silent slip on the fault. In the case of the L'Aquila earthquake, the observed precursory path is coupling of quiescence and accelerating seismic release, followed by activation. These results provide guidelines for future research on earthquake predictability.
Article
The hypothesis that large earthquakes may be preceded by a period of accelerating seismicity, or Accelerating Seismic Release (ASR), was proposed about twenty years ago. A compilation of almost one hundred peer-reviewed publications on this topic since the late 1980s to the present day shows that the rate of ASR studies increased gradually until 2004 but decreased afterwards. This negative trend is amplified by a recent increase in the number of negative results and criticisms of the ASR hypothesis. The author suggests that much of the recent negativity regarding this topic is due to the formulation of this hypothesis as a power-law fit to cumulative seismicity series. This approach is intrinsically linked to the consensus for criticality, evident from an overview of the ASR literature, to explain the emergence of power-laws in earthquake populations. The holistic view of the earth's crust as a complex system restricts seismicity pattern analyses to the study of main features such as power-laws, while a reductionist view would allow for more refined ones. Such a paradigm shift, or ‘sea change’, might be under way in the ASR literature where in 2007 a new approach was proposed to explain the ASR power-law from combined concepts of elastic rebound and geometry. Reductionism versus holism is a fundamental problem that not only applies to the study of ASR but also to the broader field of earthquake physics and earthquake predictability science.
Article
Fluid injections into geothermal systems sometimes can produce significant seismic events with magnitudes of up to 4. However, in the case of hydraulic fracturing of hydrocarbon reservoirs, such events occur extremely seldom. In the last case, in contrast to the former one, the structure of rocks is being actively destroyed (e.g., opening of tensile fractures) and the fluid-rock interaction is strongly nonlinear (e.g., a strong increase of permeability). What is the role of this nonlinearity? We consider nonlinear pore pressure diffusion to explain features of seismicity. We formulate seismicity triggering front. Its propagation is sensitive to a grade of nonlinearity, to spatial dimension of diffusion, and to the injection rate. We show that a probability of an event with a magnitude larger than a given one increases proportionally to the injected mass. An increase of this probability with time is insensitive to the nonlinearity. We compare different borehole experiments. In some cases the injection produces clearly nonlinear impact on rocks. In others this impact is approximately linear. We find a well agreement with our theory. We observe an insensitivity of temporal increments of magnitude probability to the grade of nonlinearity. These increments are controlled by injection-rate increments. In contrast, nonlinear fluid-rock interactions are characterized by a strong dominance of small earthquakes. Defects activated by a nonlinear diffusion possibly obey Gutenberg-Richter statistics with anomalous high b values. In the case of a linear diffusion, magnitude distributions of events are probably inherited from pre-existing fracture systems.