Vous êtes sur la page 1sur 34

Encyclopedia of Earthquake Engineering

DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Seismic Sources from Landslides and Glaciers


Ewald Br
uckla*, Daniel Binderb and Stefan Mertlc
a
Department of Geodesy and Geoinformation, TU Wien, Vienna, Austria
b
Zentralanstalt f
ur Meteorologie und Geodynamik, Vienna, Austria
c
Mertl Research GmbH, Vienna, Austria

Synonyms
Earthquakes generated by landslides and glaciers; Seismic activity of landslides and glaciers

Introduction
With few exceptions, the discovery and analysis of global earthquakes with landslides and glaciers as
sources did not begin until the twenty-rst century. This may come as a surprise considering that the
magnitudes of such earthquakes have reached and even exceeded M = 5. In this essay, we describe the
specic characteristics of these seismic signals and their source locations. We describe the source
mechanisms and discuss the information that can be derived from interpretation of the data. In the case
of landslides, the source process represents a severe hazard, and the same may be true for glaciers.
Although seismic signals from landslides and glaciers are observed at intermediate scales, we will jump
from magnitudes of M ~ 5 to M < 0. The deployment of seismic stations on the target landslide or glacier
is necessary to observe such weak signals. Systematic research on these microearthquakes also did not
start until the twenty-rst century. We offer an overview on landslide and glacier processes relevant to the
generation of microearthquakes. Using selected case studies, we present the diversity of seismic signals
generated by landslide and glacier sources and emphasize the close relations to geodetic and hydrometeorological data. We relate the seismic data to landslide and glacier processes as far as possible at the
current state of knowledge. The potential usage of this seismic monitoring data for hazard estimation is
also addressed.

A New Class of Earthquakes


Seismic Signals from the 1980 Mount St. Helens Eruption and Landslide

Mount St. Helens (46.2 N, 122.2 W) has been the most active volcano in the Cascade Range during the
Holocene. Its most impressive active phase in historical times started with a series of small earthquakes in
March 1980. During the rst half of the following May, seismic activity increased dramatically, and a
bulge over 100 m grew outward at the mountains north ank. The cataclysmic Plinian-style eruption, VEI
5, occurred on May 18, 1980. Within 1520 s of a magnitude 5.1 earthquake, the volcanos bulge and
summit slid away in a huge landslide (Fig. 1a). Volcanic explosions interacted with the landslide process
(Fig. 1b). The landslide, comprising a volume of 2.8 km3, rushed rst down the slope to the north and bent
later to the west-northwest along the North Fork Toutle River with a speed briey exceeding 200 km/h. The
landslide covered an area of nearly 60 km2, with the most distal deposits traveling about 25 km (Fig. 1c).
The height of Mount St. Helens was reduced from about 2950 to 2550 m (Brantley and Myers 2000).
*Email: ewald.brueckl@geo.tuwien.ac.at
Page 1 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 1 The Mount St. Helens eruption and landslide 18 May 1980; (a) the bulge at the north ank; (b) development of the
landslide during eruption; (c) landslide deposit area (Modied after Voight 1981; Moore and Albee 1981, http://pubs.usgs.gov/
gip/msh/mudows.html)

Seismic signals from the May 18, 1980, eruption and landslide of Mount St. Helens have been recorded
by seismic stations at epicentral distances from 23 up to 133 (Fig. 2a). The seismograms show strong
surface waves in the period range of 100260 s corresponding to Ms = 5.9. The spectra fall off very
rapidly at periods shorter than 75 s. This makes these recordings signicantly different from typical
tectonic earthquake spectra (Fig. 2b). A cutoff period of ~5 s would be expected for a tectonic earthquake
of similar magnitude. The low-frequency content of the Mount St. Helens recordings suggests a very slow
source process. Kanamori and Given (1982) analyzed the radiation pattern of the Love and Rayleigh
waves and concluded that a subhorizontal, about southward-directed single force is an adequate source
mechanism and describes the radiation pattern much better than a conventional DC. The earthquake
attributed to the Mount St. Helens landslide resembled the rst example of a new class of global
earthquakes.

Page 2 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Mt. St. Helens May 18, 1980, M=5.9

MAJ

D=70.5

R1

Mammoth Lakes May 25, 1980, M=6.2

a
V

KON

R1

G1

D=67.4

KON

R1

D=77.34

G1

R1
R1

R1

MAJ

R1

D=73.9

R1
R

G1

NWA
V

G1

T
D=133.3

D=134.6

R1

NWA

R1
R

R1

V
R1
R
G1

G1

T
0

30 min

Fig. 2 Recordings from seismic observatories in Japan (MAJ), Norway (KON), and Australia (NWA), 30 s high-cut ltered of
(a) Mount. St. Helens eruption and landslide, (b) Mammoth Lakes earthquake (Modied after Kanamori and Given 1982); V,
R, and T are the vertical, radial and transverse components, R1 and G1 mark the rst Rayleigh and Love phases

The Single-Force Seismic Source


Earthquakes of the new class produce long-period surface waves detectable on a global scale with
wavelengths at least ten times the extent of the source area. The spatial migration of the source during
the event cannot be resolved by long period data, therefore we consider a single force F(t) varying in time
but xed in space as the seismic source. We refer to F(t) as force history. F(t) is the reaction force to the
acceleration or deceleration b(t) times a mass M on the Earths surface (Eq. 1).
Ft btM

(1)

The acceleration is caused by the drop of the frictional force at a sliding plane from static equilibrium to
its value during sliding. The excess gravitational forces result in the acceleration. When the slope of the
landslide trajectory reduces, frictional forces exceed the driving gravitational forces, and the landslide
decelerates. The momentum I(t) and velocity of the landslide v(t) follow from integration of Eq. 1 over
time (Eq. 2).
It vtM

Fsds

(2)

Page 3 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

We have I(T) = v(T) = 0, where T is the duration of the landslide. This condition imposes a constraint
on the derivation of the force history F(t). Kanamori and Given (1982) found that the data of the Mount
St. Helens eruption can be tted by a sinusoidal F(t) with a period of 240 s.

CSF: Centroid Single Force


The displacement multiplied by the landslide mass at time t follows from an integration of I(t) (Eq. 2). The
total displacement vector D times M is given by Eq. 3.
DM

Itdt

(3)

The product of D and M can be derived from seismic data. Satellite images or other topographic
information may also be used to break down the product into its factors D and M. Equations 1, 2, and 3 are
valid for any surface mass accelerated or decelerated by either gravitational, inertial, or frictional forces,
stress relief, or volcanic processes. Kawakatsu (1989) showed that D M represents the spatiotemporal
centroid single force vector CSF = D M, the analogue of the CMT for moment tensor seismic sources.
He also provided the formalism for normal-mode inversion.

Global Glacial Earthquakes


A search for earthquakes of the new class in Arctic and Antarctic areas in data provided by global
networks was carried out by Ekstrm et al. (2003) for the period 19992001 and extended to the period
19932008 by Tsai and Ekstrm (2007) and Nettles and Ekstrm (2010). About 243 earthquakes were
found in Greenland, 14 in Antarctica, and 1 in Alaska. The one in Alaska (Fig. 3) exemplarily shows the
difference of the new-class earthquake to a nearby tectonic earthquake of similar magnitude and may
explain why the new class of earthquakes has so long been undiscovered. CSF inversion was applied to
the events in Greenland and Antarctica by Tsai and Ekstrm (2007) and Nettles and Ekstrm (2010),
resulting in an improved location with an uncertainty of approximately 20 km. These events were
clustered at or near eight fast-moving outlet glaciers with speeds >800 m/year (Fig. 4). An antisymmetric
boxcar function (25 s of constant acceleration before the centroid focal time followed by 25 s of constant
deceleration) was assumed in order to derive reactive forces transmitted to the Earths crust by the
displacement of a hypothetical mass. The magnitudes of centroid single force vectors were in the range
10 1012  CSF 200 1012 m kg, and the magnitudes were 4.6  Ms  5.1. Most CSF vectors point in
the opposite direction of the general ow of the glaciers.
The frequency of glacial earthquakes in Greenland is subject to seasonal variations and synchronous
with maximum glacier ow velocities in JulyAugust and a corresponding minimum in January and
February (Nettles et al. 2008). A power law does not properly describe the magnitude/CSF frequency
relation. It seems that each outlet glacier has its typical earthquake, which depends on glacier size,
hydrological conditions, and the calving rate (Tsai and Ekstrm 2007).
Chen et al. (2011) detected 13 more new-class glacial earthquakes in Antarctica with epicenters at or
near the Ronne Ice Shelf, Ninnis Glacier, and Vanderford Glacier (Fig. 5). At the Ronne Ice Shelf, the
direction of the CSF vector was normal to the ice ow direction and parallel to the ice front; force direction
was parallel to the rift propagation direction. At Vanderford and Ninnis glaciers, CSF and local ice ow
directions coincided with only one exception. The magnitudes of the glacier earthquakes in Antarctica
were 4.2  Ms  4.9.

Page 4 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 3 Recordings of the Alaska glacial earthquake and a M 4.2 earthquake (Modied after Ekstrm et al. 2003)

Global Landslide Earthquakes


Ekstrm and Stark (2013) applied the detection and location method developed for the glacial earthquakes
to known large landslides or areas where a potential for giant landslides can be assumed. The landslide
force history F(t) was discretized into a sequence of overlapping isosceles triangles with a half-width of
1015 s. Momentum, displacement, and CSF were derived from F(t). Satellite images and other
topographic information were used to separate the factors D and M out of the product CSF. Figure 6
shows, as an example, the Hunza-Attabad landslide in North Pakistan with a total mass of 140 109 kg.
Ekstrm and Stark (2013) analyzed seismic recordings of 29 giant landslides in the Aleutian and Coast
Ranges, the Rocky Mountains, the Andes, the Alpine Himalayan orogeny, Taiwan, New Guinea, and
Antarctica, with the Mount St. Helens 1980 landslide included. The magnitudes of these events range
from Ms = 4.65.6. The CSF can be calculated from the data provided by Ekstrm and Stark (2013). CSF
or log10(CSF) signicantly correlates with the landslide mass M and the magnitude Ms (Fig. 7).

Mechanisms of Giant-Single Force Sources


Landslide: Rotating Slider Block Model
Planar slider block models have been proven to support understanding of earthquake processes considerably. When considering the application of a planar slider block to earthquake processes, the slider block
is loaded by tectonic strain, and after strain release the slider block comes to rest. For landslides, the
situation is different; on an inclined planar sliding plane the slider block accelerates constantly and the
travel distance is not conned. A rotational slider block overcomes this problem and connes the travel
distance (e.g., Br
uckl and Parotidis 2005).
Page 5 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 4 Glacial earthquakes in Greenland (Modied after Tsai and Ekstrm (2007) and Nettles and Ekstrm (2010))

Figure 8 shows the geometry of a rotational slider block. The dynamic system of a landslide is
approximated by a mathematical pendulum with the length R and the mass M concentrated to a point.
The position of the mass is dened by the angle a times R. Before the landslide happens at a = a0, the
driving gravitational force M g sin(a) is in equilibrium with the frictional force m0 M g cos(a), which
always acts in opposition to the direction of movement (Eq. 4). A decrease of the friction coefcient
from m0 to m initiates the landslide. The excess driving force is compensated for by the inertial force
M d2a/dt2 R (Eq. 5). We only look at one half oscillation of the slider block.
M g sin a0  M g cos a0m0 0

(4)

M g sin a  M g cos am M R d2 a=dt2

(5)

We consider a rst-order approximation for small amplitudes (sin(a) ~ a, cos(a) ~ 1) and travel
distance D ~ R (a  a0). Introduction of these approximations and subtraction of Eq. 4 from Eq. 5 gives
R=gd2 a=dt2 a  a0 m0  m Dm

(6)

Page 6 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 5 Glacial earthquakes in Antarctica (Modied after Nettles and Ekstrm (2010) and Chen et al. (2011))

The solution of the inhomogeneous differential equation for the initial conditions a = a0 and
d/dt(a) = 0 is given by Eqs. 7 and 8.
at a0 Dm1  cos o0t

(7)

o0 2p=T0 sqrtg=R

(8)

The position angles a of the slider block at start time t0, stop time t1 = T0/2, and the time of maximum
velocity tm = T0/4 are a0, a1 = a0 + 2 Dm, and am = (a0 + a1)/2 = m = atan(CSFz/CSFh). The
travel distance D and maximum magnitude of the landslide force (positive for t = t0, negative for t = t1)
are given by Eqs. 9 and 10.
D CSF=M Ra1  a0 2 RDm 2Dm gT0=2p2

(9)



Fmax M g Dm g CSF=2R p2 CSF= 2T0=22

(10)

The half-period T0/2 of the rotational slider block is the duration of the slide or the landslide force
history T. We introduce a calibration factor k < 1 (Eq. 11) to account for the observation of the landslide
force history F(t) not starting and ending with the extreme value predicted by the slider block model.

Page 7 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

KBL-IU
BHZ

ABKTII
LHZ00

KBL-IU
LONG

ABKTII
TRAN
UOSSII
LHZ00

KBL-IU
TRAN

ARUII
LHZ00

KZAKN
LHZ

GNIIU
LHZ00

AAKII
LHZ00

KIVII
LHZ00

AAKII
LONG

KIVII
TRAN

AAKII
TRAN

ANTOIU
LHZ00

EKS2KN
LHZ

200 100 0

100 200 300 400 500 600


seconds

Force [1012 N]

ANTOIU
TRAN

200

400

600 800
seconds

1000 1200

0.3
0.2
0.1
0.0
0.1

North
East
Up

0.2
0

10

20

30 40
Time [s]

50

60

70

Fig. 6 Hunza-Attabad landslide 04 January 2010 (Modied after Ekstrm and Stark 2013); (a) observed and synthetic traces;
(b) landslide force history; (c) source (red outline) and deposition area (blue) of the landslide; src and dpo are centroids of the
landslide mass M before and after the slide, D the trajectory derived from satellite image; given CSF = M D = 14*1013 kg m
the landslide mass is M ~ 14*1010 kg; note the damming of the Hunza river

Page 8 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 7 Relations between landslide data (Ekstrm and Stark 2013); (a) Mass CSF, (b) Ms log10(CSF)

Fig. 8 Rotational slider block model (for description of symbols see section Landslide: Rotating Slider Block Model)

T0=2 k T

(11)

A calibration factor of k = 0.68 brings Fmax, calculated in Eq. 10, into a nearly perfect 1:1 relation
with the Fmax derived from the inverted forcetime function (Ekstrm and Stark 2013; Fig. 9a). Figure 9b
shows the dependence of am and a0 on CSF. Over the range of CSF (approximately three decades), the
average sliding angle reduces from ~25 to ~10 . There is also a systematic increase of ratio (a0am) to
a0 with CSF from ~28 % to ~55 %.

Calving Glacier: Toppling of an Iceberg


The large outlet glaciers of Antarctica and Greenland lose mass mainly by calving. One of the bestmonitored sites where calving takes place is the front of the Jakobshavn Glacier near Ilulissat, West
Greenland. On 28 May 2008, a gigantic sequence of calving events was lmed by a team of Chasing Ice
(http://earthsky.org/earth/video-largest-glacier-calving-ever-caught-on-lm). An area of the size of Manhattan calved successively over a time span of 75 min. One can observe a great number of slabs of about
1000 m in height toppling from unstable to stable equilibrium, thus pushing the iceberg mlange of
previously calved icebergs away from the ice front out to sea. This toppling of large icebergs has been
considered a candidate for a seismic source of global scale (Tsai et al. 2008).
Figure 10 illustrates the geometry, kinematics, and acting forces during the toppling of an iceberg,
leaving it in unstable equilibrium after calving from the front of an ice stream. The calving front may or

Page 9 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 9 Relations derived from the rotational slider block model; (a) maximum force observed maximum force calculated; (b)
a, a0 log10(CSF)

may not be grounded. The side lengths of the cross section are H (the height before toppling) and L. The
width of the iceberg is W. With H > L, the iceberg is initially in an unstable position. Toppling of the
iceberg implies not only a rotation of ~90 but also acceleration off the glacier front. This inertial force
causes a reaction force at the contact area of the iceberg with the glacier, which is then transmitted via the
glacier ice to the glacier bed and represents the main part of the glacial seismic source. A rst-order
approximation of CSF:
CSF  MH  L=2, with M H L W di

(12)

The density of ice is di and that of seawater dw. The characteristic time scale of toppling can be
estimated from the period of an iceberg interpreted as a mathematical pendulum:
T  2p sqrtdi H=dw  di g

(13)

Page 10 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

glacie

glacier be
d
Antarctica

r bed
Green
land

b
100 m
1000 m

Fig. 10 Toppling (capsizing) of an iceberg acting as single force seismic source; (a) kinematics and acting force (toppling may
also occur by counterclockwise rotation); (b) toppling of an iceberg during the 28 May 2008 calving event at Jakobshavn
glacier

Equations 12 and 13 do not take into consideration the remaining kinetic energy the iceberg has after
reaching its stable position at the horizontal travel distance (H-L)/2. Furthermore, the friction between the
iceberg and the glacier ice front along with the interaction of the iceberg with seawater and the mlange of
smaller icebergs and fragments in front of the calving front is not taken into account. However, assuming
the very reasonable dimensions H = 800 m, L = 100 m, and W = 400 m, CSF ~ 1013 kg m (Eq. 12),
T ~ 150 s (Eq. 13), and Ms = 4.8 according to the relation between log10(CSF) and Ms (Fig. 7b).

Microearthquakes Generated by Landslides


The dynamic collapse of a slope within several tens of seconds is only representative of how a landslide
can end in the most extreme cases. However, the majority of landslides do not develop into dynamic and
catastrophic failure but reach more or less stable nal states after phases of varying levels of activity.
A deeper understanding of landslide processes and evolution is of fundamental importance for early
warning and hazard estimation. Seismic activity connected to these processes can be observed by local

Page 11 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 11 Landslide processes and evolution; (a) reduction of rock mass strength; (b) temporal evolution of displacement

seismic networks deployed on the landslide mass. We term microearthquakes generated by landslides
LMEs (landslide microearthquakes) in the following.

Landslide Processes and Evolution


We assume brittle rheology of the rock material constituting a landslide as a precondition for seismic
activity. However, under some circumstances, the landslide mass may behave as ductile, and even a brittle
deformation will not necessarily act as a seismic source. In the following, we present a simplied scheme
of landslide processes and possible evolutions to dynamic failure or stabilization (Fig. 11). This schema
should provide a backbone for the interpretation of the seismic data provided by the case studies presented
in sections Pre- and Postfailure Processes at Steep Slopes to Flow.
Slopes of a given inclination and height can only exist if the stability is guaranteed by the initial strength
of the rock material. The possible physical states are depicted in Fig. 11a in a Mohr diagram by the three
Mohrs circles (M1, M2, M3) and Mohr-Coulomb failure criteria representing the initial and degraded
strengths of the rock mass forming the slope. The Mohrs circles symbolize the states of stress, which are
relevant for the stability of the slope. The initial rock strength is given by a Mohr-Coulomb failure
criterion for the shear stress, with C0 and f0 the initial cohesion and the initial friction angle. Increasing
pore pressure p shifts the Mohrs circles to the left and nearer to the initial rock strength. In an extreme
case of high pore pressure, a dynamic rupture could be induced; however, this is not the only situation that
reduces the slope stability. Subcritical crack growth (Br
uckl and Parotidis 2005) supported by deep
chemical or thermomechanical weathering (Gischig et al. 2011) may progressively damage the rock
material and degrade the rock strength. Ultimately, cohesion (C0) is completely lost and the friction angle
diminished considerably from f0 to f. This failure criterion of the degraded rock strength intersects
Mohrs circle M1 and is tangential to M2, which means that in the rst case a dynamic failure has already
occurred. In the second case, the slope is in an equilibrium state of driving (gravitational) and resistive
(frictional) forces. Creep of the whole slope may reduce the inclination and therefore the driving forces,
and this situation is symbolized in Mohrs circle M3. It is well below the failure criterion of the degraded
rock. The slope comes to an ultimate stabilization by this process.
The temporal evolution of the deformation or displacement of a landslide mass is schematically shown
in Fig. 11b. The initial phase corresponds to the progressive damage from initial to degraded rock strength
(Fig. 11a), where the whole rock mass of the landslide is degraded and deformed. However, deformation
concentrates successively to one basal or several main sliding zones or surfaces. Toward the end of the
initial phase, the further development of the slope comes to a bifurcation. In the case where Mohrs circle
M1 represents the critical state of stress in the slope, the initial phase leads directly to a dynamic slope
failure. In the case of Mohrs circle M2-type critical state of stress, a phase of quasistationary creep,

Page 12 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 12 Monitoring sites in the European Alps

eventually interrupted by episodic phases of high creep or sliding velocities, follows. However, we cannot
exclude that episodic high activity is the precursor for a later bifurcation between quasistationary creep
and dynamic failure, as indicated in Fig. 11b by a second grey dot. The longer the quasistationary phase
lasts, the more the driving forces and the radius of the Mohrs circle reduce, so that in the nal state of
stress represented by M3 the slope stabilizes. Decrease of the pore pressure p by improved drainage of the
slope due to increased permeability of the rock mass may support stabilization.

Rockfall
Rockfall is a phenomenon occurring on slopes where, at least locally, the inclination equals or exceeds
that of talus of the same rock material. It may be the consequence of weathering but also an indication of
increased activity in a landslide. The seismic signature of rockfall may be a sequence of many irregular
impulsive signals moving down the slope. Figure 13 shows seismic recordings from the tumbling of a
medium-sized boulder (~2 m diameter) at the Gradenbach site (Fig. 12), images of the boulder itself and
the site of heaviest impact, and the seismic location of that place (Br
uckl et al. 2013).
As a second example, the seismic signature of a rockfall at Steinlehnen (Fig. 12) is documented in
Fig. 14. This rockfall was more of a small rock avalanche, which left behind a path resolved by groundbased InSAR monitoring. The whole event lasted around 30 s, and location applied to several time
windows allowed for the tracing of the path of this rockfall. There is fair agreement between both tracing
methods (Weginger 2012).

Pre- and Postfailure Processes at Steep Slopes

We denote steep slopes as slopes that are steeper than the initial angle of friction or the inclination of
talus built up by the same rock material. The rock material of these slopes needs cohesion to balance the
gravitational forces. In the case of cohesion dropping to zero because of progressive damage (Fig. 11a),
the slope will immediately transit from this initial phase to dynamic failure (Fig. 11b). Identication and
monitoring of precursors of this kind of dynamic rupture would be of great scientic and practical
importance.

Page 13 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

2 sec

PDF

8500

seismic station
8000

0.95

7500

0.9

7000

0.85

[m]

4
6500

1
0.8

6000

0.75

5500
0.7
5000
0.65
4000

4500

5000

5500

6000

6500

7000

[m]

Fig. 13 Rockfall at the Gradenbach landslide (Bruckl et al. 2013); (a) main boulder; (b) recordings by the local network;
(c) main impact; (d) location by the seismic data (normalized PDF) and actual position of the impact (star)

The retreat of cliffs along the British and French Channel Coasts by sea erosiondriven rockfalls is an
ongoing process. Amitrano et al. (2005) report seismic monitoring of a cliff collapse at Mesnil-Val on the
Normandy coast, France. Precursory seismic activity was recorded about 1 h before the collapse.
Frequency of detected seismic signals and the seismic energy obey inverse Omoris laws. The data
suggests that a warning based on increased activity could be given ~10 min before the collapse and a
reliable prediction of the rupture time about 1 min in advance, provided data acquisition and continuous
updating of calculations are done in real time.
The village of Randa, Matter Valley, Swiss Alps (Figs. 12, 15), was threatened by rock- and icefalls
several times in its history. We anticipate a look from rock to ice and icefalls because of the similar
phenomenology behind these processes in rock and ice. Icefalls were released by the Weisshorn hanging
glacier ve times in the last 35 years. The break-offs in 1973 and 2005 were monitored by geodetic and
seismic methods. The velocity of the glacier front developed according to an inverse Omori law (Faillettaz
et al. 2011) toward the instant of rupture. Log-periodic oscillations appeared superimposed on the general
trend. Starting 30 days before rupture, three phases characterized by different cumulative size-frequency
distribution of the signal energies and interevent waiting-time distribution behavior were identied and a
correlation of seismic activity with the log-periodic oscillations established. These phases represent the
development from random crack generation over crack concentration at a fracture zone to the nal
rupture. The ndings could be the basis of an efcient in-time warning.
From the detour to ice and a hanging glacier, we revert to rock and landslides. On 18 April and 9 May
1991, two successive rockslides took place near Randa (Fig. 15). The failure volumes involved about

Page 14 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015
GB-SAR raw data image (time in CET)
13.06.10 22:26:31 - 13.06.10 22:32:52, t2 = 1000

SL 03

DB-SAR: differential displacement [mm]

1400

Y [m]

1200
SL03
1000
SL01
800
SL02

600

400

SL0* seismic station

800

600

400

200

2
SL 01

SL 02

3
0
X [m]

200

400

600

800

60 sec

Fig. 14 Rockfall at the Steinlehnen landslide (Weginger 2012); (a) differential ground based InSAR image shows the trace of a
rockfall (black arrow); yellow arrow represents the path of the rockfall from the seismic data shown in (b)

Fig. 15 View from east to the glacier break-off and rockslide sites near Randa; location of the prole in Fig. 16 is shown by a
dashed red line

20 and 10 106 m3 massive para- and orthogneisses (Eberhardt et al. 2004). The average slopes of the
failure planes were roughly 60 and 50 (Fig. 16). The rock masses dropped down 200 and 300 m and
produced Ms 4.9 and 4.8 earthquakes detected by Ekstrm and Stark (2013). Deformation of the slope
above the actual scarp continued after the rockfall sequence of the year 1991 with ~1020 mm/year
Page 15 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

2500

Scarp (May 9, 1991)


2250

iss

gne

a
Par

Scarp (April 18, 1991)

is
gne

ho

Or t
2000

PDF
3
10
10
10

ss

ei
agn
r
a
P

10
10

1750

ss
nei s)
hog neis
t
r
G
O
nda
(Ra

4
5
6
7

current topography

1500

200 m

topography before
rockslide

Fig. 16 Cross section through the Randa rockslide site (Modied after Eberhardt et al. 2004); cumulative PDF of LMEs is
shown within the detail marked by black rectangle (Modied after Spillmann et al. 2007); the location of a possible next scarp
is marked by red dots

(20042008). A seismic monitoring campaign was started in January 2002 and continued until July 2004
(Spillmann et al. 2007), where a monitoring network was deployed above the scarp (Fig. 15). Two types of
waveforms classied as LMEs were observed: 94 single events comparable to the recordings of small
detonations and 129 multiple events with a more complex source time function. Frequency content of the
signals reached up to 100 Hz for the surface sensors and up to 200 Hz for the borehole sensors. Most of the
seismic activity was concentrated on a volume of low seismic velocities (<= 1500 m/s) near the scarp,
but also deeper-reaching zones of increased seismic activity were identied, possibly outlining the scarp
of a future rock slide (Fig. 16). Hydrological and meteorological conditions had no impact on the seismic
activity and the geodetically observed displacement rates. Most magnitudes were in the range 2.0 <
Mw < 0.5. It is questionable if the observed seismic activity and geodetically observed displacements
are postfailure relaxation processes or if they belong to the initial phase of a third, larger rock slide
at Randa.

DSGSD: Deep-Seated Gravitational Slope Deformation


Deep-seated gravitational slope deformation (DSGSD) is a phenomenon frequently observed in the hard
rock slopes of alpine regions. Typically, the deformation comprises the whole ank of a valley, and the
maximum thickness of the deformed rock mass exceeds 100 m. The mechanisms of DSGSD are manifold,
and deformation styles such as Bergzerreiung, Talzuschub, sagging or Sackung (Zischinsky 1969),
and kink band slumping (Goodman and Kieffer 2000) can be summarized under DSGSD. The initiation of
DSGSD at alpine slopes mostly dates back to the retreat of the ice age glaciers, when the lateral support of
the valley ank by the ice pressure was lost. This conception implies a long initial phase of mass
movement, during which progressive damage degraded the rock strength almost to its residual value
(see Fig. 11a; Br
uckl and Parotidis 2005). DSGSDs may be in a quasistationary creep phase or may

Page 16 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 17 View from south to the Schilienne DSGSD and the active zone (area colored in red); the locations of the 2 seismic
micro-arrays RUI and THE are marked by symbols

exhibit episodic high deformation rates in parts of the slope or its entirety interrupting longer phases of
little activity (Fig. 11b). There is currently no general answer as to whether these processes eventually lead
to stabilization or a transition to dynamic rupture.
In the following, we consider the seismic signals recorded at three DSGSD examples in crystalline
rocks of the European Alps: Schilienne (Helmstetter and Garambois 2010), Steinlehnen (Weginger
2012), and Gradenbach (Br
uckl et al. 2013; Mertl 2015). The locations of these landslides are shown in
Fig. 12; Figs. 17, 18, and 19 supply information about the morphology, the seismic monitoring network,
and partly the geology and structure of these landslides. The average slopes of these DSGSDs vary
between 28 (Gradenbach) and 35 (Steinlehnen).
At the Schilienne DSGSD, a presently very active zone is located on top of the steeper lower part of the
slope (marked in Fig. 17). The velocity of this active zone reached 1.4 m/year in 2008. Until April 2009,
rockfalls were the most frequent seismic sources related to the dynamics recorded at the Schillienne
DSGSD. The cumulative frequency of >3000 events obeys a power law distribution over 2.5 decades of
the amplitude range with an exponent corresponding to a b-value of b = 1.1.
Figure 20 shows typical waveforms of LMEs at the Schilienne DSGSD. One type corresponds to
impulsive sources similar to small explosions, while another event type shows mostly emergent onsets
and represents source-time functions lasting ~5 s, either relatively broadband or low frequency. The
LMEs cluster either in the high-velocity area or the head scarp. It is also possible that some hypocenters
are located at or near a basal shear zone. Rockfall and LME frequencies precede accelerations of the active
zone by 45 days. Some but not all rainfall events trigger increased rockfall and seismic activity
(Helmstetter and Garambois 2010).
Seismic monitoring campaigns at Steinlehnen and Gradenbach had the opportunity to observe episodic
phases of high landslide velocities interrupting longer periods of quasistationary creep. During June 2010,
the Steinlehnen DSGSD was monitored by a ground-based InSAR system installed opposite the
Steinlehnen DSGSD. An acceleration phase of the active slab was observed with total displacements
Page 17 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 18 Steinlehnen DSGSD (Weginger 2012); (a) view from east to the slope with contour of the active zone; triangles mark
the locations of the seismic monitoring stations, the extent of the slab accelerating during the observation period is colored in
ochre; (b) geological cross section A-B (Modied after Zangerl et al. 2007)

reaching up to 0.9 m. Up to ve seismic monitoring stations were operating at the Steinlehnen slope
during this period (Fig. 18a). Besides the rockfall (see Fig. 14), Weginger (2012) identied three classes
(n, ln, i in Fig. 21a) of LMEs that had their sources within the area of the accelerated slab. This
classication was based on characteristic properties of the seismic signals (e.g., duration, frequency
spectral characteristic, amplitude, envelopes, cross correlations, polarization). Figure 21b shows velocity
and displacement of the highly active slab for the period of 1028 June 2010. The high-velocity phase
starts on 20 June and lasts until 25 June. The cumulative frequencies for LME types n, ln, and i rise from
13 June onward, and some rockfalls also occur (Fig. 21c). A sudden increase of the cumulative rockfall
frequency accompanies the onset of the high-velocity phase between 20 and 21 June, and rockfall ceases
completely thereafter. The cumulative frequencies of the events n, ln, i increase more or less continuously
with the highest increments during the high-velocity phase. No extraordinary rainfall was reported to

Page 18 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 19 Gradenbach DSGSD (Br


uckl et al. 2013): (a) Orthophoto with contour of the active zone and seismic stations
GB01GB07; the main scarp, the highly fractures zone in the uppermost area of the mass movement, and the erosional zone at
the toe of the mass movement are well visible in the orthophoto; (b) cross section A-B; the depth of the basal sliding surface or
zone was determined by drilling (boreholes E2 and GB) and refraction seismic measurements (c) displacement history derived
from terrestrial geodetic measurements 19691991, photogrammetric evaluation of the aerial photos 1962 and 1996, and GPS
since 1999

precede or accompany the period of high velocities of the active slab, and no external trigger of this phase
is known so far. The incipient seismic and rockfall activity on 13 June was most likely the only precursor
of the high-velocity phase in the following week.
The episodic highsliding velocity phase at the Gradenbach DSGSD occurred during spring of 2009. In
contrast to the Steinlehnen event, the whole landslide became active due to inltration of melting snow
and precipitation (Br
uckl et al. 2013). Five different LME types (AA, A, B, D, tremor; Fig. 22a) besides
rockfall (rf) were identied using criteria comparable to those used for the data from the Steinlehnen site
(Mertl 2015). The cumulative frequencies of these events behave distinctly. Event types A and D start
with the onset of snow melt (Fig. 22bd) and high inltration. The tremor type was only observed at that
time. Thereafter, a quiet phase that lasted over 1 month followed. The beginning of the episodic phase of
high creep velocity was accompanied by type B microearthquakes and some rockfall. The frequency of

Page 19 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

100
[Hz]
50

33

34

35

36

37

0
10

20

30

40

50

100
50

30

29

31

32

33

0
10

20

30

40

50

100
50

204

205

206

207

208

170

180

190

200

210

220

230

100
50

135

136

137

138

139
[sec]

0
110

120

130

140

150

160
[sec]

Fig. 20 Typical waveforms and spectrograms from LMEs recorded at the Schilienne DSGSD (Modied after Helmstetter
and Garambois 2010); (a) impulsive LME; (b) small explosion; (c) emergent and broad band LME; (d) emergent and low
frequency LME

Fig. 21 LMEs observed at the Steinlehnen DSGSD, 1028 June 2010 (Weginger 2012); (a) typical waveforms and spectrograms of LME types n (emerging signal, few seconds length), ln (irregular sequence of n-type signals), i (irregular sequence of
impulses over a period of several seconds); (b) displacement and velocity; (c) development of cumulative frequency of event
types n, ln, i, and rf (rockfall)

Page 20 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 22 LMEs observed at the Gradenbach DSGSD, MarchOctober 2009 (Bruckl et al. 2013; Mertl 2015); (a) typical
waveforms and spectrograms of LME types AA (impulsive onset with short lower frequency coda), A (like AA but with long
coda), B (emergent onset, low frequency, typical duration 2030 s), D (emergent onset but signicantly higher frequency
content and shorter duration than B), tremor (emergent low frequency signal, duration >1 min, superimposed by irregular
spikes), rf (rockfall); (b) displacement and velocity; (c) snow cover water equivalent and precipitation; (d) development of
cumulative frequency of event types

type A and D events rose 12 weeks later and lasted for the rest of the episodic high-creep phase.
Magnitudes of the LMEs range from 2.8 to 0.9 with a b-value of b = 1.7.
Though the geodetically observed movements covered the whole landslide mass, the spatial distribution of the seismic activity during the episodic high-creep phase is concentrated to limited areas (Fig. 23).
Event types A, D, and tremor, which coincide with the onset of snow melt and high inltration, are most
likely very shallow events. The waveforms and coincidence in time with the onset of high creep velocities
may be indications that the hypocenters of B-type LMEs are at the level of the basal sliding zone.

Page 21 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 23 Spatial distribution of seismic activity MarchOctober 2009 at the Gradenbach DSGSD; colored contour plot
visualizes the frequency of possible epicenters with normalized PDF > 0.7, event types A, AA, D; blue area marks area of
event type B epicenters (Br
uckl et al. 2013; Mertl 2015)

Asperities must have been overcome before the whole landslide mass could increase in velocity (Mertl
2015).

Flow

A ow is a spatially continuous movement in which surfaces of shear are short lived, closely spaced, and
not usually preserved. Whereas in DSGSD the original structure of the rock mass is partly preserved, in a
ow it is completely destroyed and mixed up. The ratio of mean thickness to length of a ow is <0.01 and
signicantly lower than for DSGSDs. When considering the scheme of landslide processes and evolution
presented in Fig. 11a, it is evident that the initial rock (or soil) strength has been completely reduced to its
residual values. Changing pore pressure may keep the ow either active or temporarily inactive (Mohrs
circles M2 and M3 in Fig. 11a). Flow here represents quasistationary creep including episodic high
activity, depending on inltration and the pore pressure (Fig. 11b). The distribution of velocities in space
and time within the landslide mass justies its macroscopic description as a ductile medium. However, as
the following two examples will prove, brittle deformations or interactions between pore uid and the
rock/soil mass also exist and resemble seismic sources.
The Super-Sauce landslide is a ow-type mass movement situated in the Barcelonnette basin in the
southern French Alps (Fig. 12). It developed from heterogeneous soft Jurassic black marks (Fig. 24). The
average slope of the ow is ~23 . The landslide triggered around 1960, ruling out a causal connection to
the retreat of ice age glaciers. From 1996 to 2007, the velocities covered the range of 230 mm/day.
Seismic monitoring at the Super-Sauce landslide was carried out between 14 and 24 July 2008 (Walter
et al. 2012). Besides rockfalls, 34 LMEs were located and labeled as slide quakes. They have an
impulsive onset, last 25 s, and cover a wide frequency band (1080 Hz). Later-arriving phases have
lower frequencies (Fig. 25a). The epicenters are concentrated around the high-velocity area of the
landslide. Magnitudes between 3.2  M  1.3 were detected following a power law distribution
with a b-value of b = 0.84. The hypocenters were probably located within the uppermost few meters and
not at the basal sliding zone or surface. The highest magnitudes of slide quakes followed few hours after a
strong rainfall event; the temporal distribution was otherwise uniform. Single miniarrays also detected

Page 22 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 24 Super-Sauce ow, red dotted line outlines landslide area (Walter et al. 2012)

very high-frequency events (5150 Hz) with durations of 220 s (Fig. 25b). These LMEs were concentrated around a stable crest within the high-velocity zone of the ow and related to ssure development in
the uppermost layer of the ow. Laboratory tests proved that the ow material can have brittle rheology.
The second example of a ow is the Slumgullion landslide (Fig. 26). The landslide occurs in Tertiary
volcanic rocks and comprises sandy, silty clay with scattered patches of boulder debris, clay, and pond and
stream sediments. The landslide consists of an active upper part, which moved over an older, inactive
lower part in its path. The average slope of the active part is 13 . The Slumgullion moves primarily by
sliding along discrete bounding faults. Additionally, differential internal processes take place at faults and
fractures within the landslide body.
During 1826 August 2009, a seismic monitoring campaign was carried out at the Slumgullion
landslide (Gomberg et al. 2011). Geodetic and hydrometeorological monitoring supplemented the
campaign. During the monitoring campaign, the displacement trend was ~10 mm/day. One stronger
period of rainfall occurred during the observation period. Several seismic signals were detected and
related to LMEs: tremor and harmonic slide quakes (Fig. 27a, b). The dominant energy of tremor is
distributed broadly above 3050 Hz and diminishes toward 125 Hz. High amplitudes of tremor envelopes
last <10s and follow irregularly in intervals of several tens of seconds. Harmonic slide quakes last ~2 s,
their fundamental frequency is ~12 Hz, and no harmonics are visible throughout the whole duration. This
LME type occurs more frequently during daytime, and an increase in sliding velocity is also observed
during the day. Tremor and harmonic slide quakes most probably originate at the left side-bounding
strike-slip faults of the landslide. The stickslip behavior of the side-bounding strike-slip faults is
explained by a transient dilatant strengthening.

Page 23 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 25 Typical waveforms and sonograms of LMEs observed at Super-Sauce debris/mud ow (Modied after Walter
et al. 2012); (a) slide quakes; (b) ssure development

Fig. 26 Slumgullion ow (Gomberg et al. 2011)


Page 24 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

P05
P03
O11
O07
O04
N10
N05
N03
M11
M05
M03
L06
L03
K11
K05
J10
J06
J02
I07
H08
G09
G05
F08
F02
E07
E03
D06
D02
C05
B08
B04
A11
A07
A03

20

40

60

80

100

120

140

Seconds from 08/25/09 07:48:00

A03
A05
A07
A11
B02
B06
C03
C05
C07
D02
D06
E03
E05
E07
E02
E06
G03
G05
G07
H02
H06
I03
I05
I07
J02
J04
J06
J08
K03
K05
K07
L02
L04
M03
M05
M07
N02
N04
O03
O05
P02
P04

Seconds from 08/25/2009 17:10:46.73

Fig. 27 Typical LME waveforms observed at Slumgullion ow (Modied after Gomberg et al. 2011); (a) envelopes of tremor;
(b) harmonic slide quake

Microearthquakes Generated by Glaciers


Ductile and Brittle Deformations of Glaciers
In contrast to landslides, which are intrinsic transient phenomena, glaciers or ice sheets can reach a longlasting stationary state in steady climatic conditions. Snow is transformed to ice and added to the glacier in
the accumulation area and removed in the ablation area. Glacier ow transports ice from the accumulation
area to the ablation area, thereby keeping the total mass as well as the thickness of the glacier constant
under stationary conditions. The glacier ow is driven by gravitational forces, controlled by the rheology
of the ice and physical conditions that allow for sliding of the glacier ice over its bed (Cuffey and Paterson
2010).
We consider an innite slab of ice with constant thickness H resting on a plane glacier bed with the
slope a. Shear stress parallel to the glacier bed is proportional to sin(a) and increases linearly with the
overburden (Fig. 28a). According to Glens law, the corresponding shear strain rate is proportional to the
nth power of the shear stress, with n ~ 3.
Integration yields the velocity prole, which is an (n + 1)th-order parabola. For cold glaciers, the ice
temperature is below melting point at the base, and the ice may be frozen to the glacier bed (Fig. 28b). For
temperate glaciers, a thin coat of water between ice and the bedrock is assumed, and sliding occurs over
the thin viscous water layer (Fig. 28c). Sliding at the glacier bed is controlled by obstacles, which are
overcome either by enhanced ow due to stress concentration at the obstacles or regelation. All processes
related to glacier ow described so far are either ductile (internal deformation of the ice, viscous ow of
the water coat at the glacier bed, enhanced ow over obstacles) or correspond to another steady-state
process (regelation) and cannot produce earthquakes.
However, the rheology of ice changes signicantly in the case of the hydrostatic pressure component
p of the state of stress becoming negative. In this case, the ice behaves in a brittle manner and may radiate
seismic energy. First, we consider tensile crack nucleation and propagation without the inuence of water.

Page 25 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 28 Planar slab of ice resting on a sloping plane; (a) shear stress parallel to surface; (b) velocity prole of a cold
glacier no basal sliding; (c) velocity prole of a temperate glacier basal sliding

Fig. 29 Crevasse formation due to thickness adaption to increasing slope; stars mark possible locations of GMEs

Glacier thickness adapts to varying slopes of the glacier bed by keeping the basal shear stress approximately constant. Therefore, the thickness is inversely proportional to sin(a). Neglecting accumulation or
ablation over a relatively short distance of an increasing slope, the velocity must increase to keep the ux
of ice constant (Fig. 29). Longitudinal tensile stress must develop to extend the glacier ice moving from
the gentle to the steeper slope. Surface crevasses develop to the depth at which the hydrostatic pressure
equals the longitudinal tensile stress component. On alpine glaciers, surface crevasse depths seldom
exceed 2530 m. After reaching the zone of constant steeper slope, the tensile stress component
disappears and the crevasses close by ductile deformation.
Water signicantly inuences the mechanics of landslides and glaciers. The fundamental difference
between landslides and glaciers with respect to the effect of water is that buoyancy is lower than
hydrostatic pressure in landslides and higher in glaciers due to the comparative densities of water
(~1000 kg m3), rock (~2500 kg m3), and ice (~900 kg m3). Therefore, a water column of sufcient

Page 26 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 30 Hydraulically induced fracturing in a glacier (Modied after Walter 2009; Walter et al. 2010); stars mark possible
locations of GMEs; (a) opening of internal and basal crevasses due to high water pressure; meltwater inux from active
moulins or basal channels (b) closure of internal and basal crevasses due to decreasing water pressure; the collapse of a basal
cavity generated during the high water pressure phase acts as seismic source

height can introduce a tensile stress regime and lead to crack development and propagation at all glacier
depths. Possible scenarios are shown in Fig. 30.
Thus far, we have only considered hard-bedded glaciers where the glacier bed consists of hard rock and
is not deformed by the glacier ow. However, it is well known that fast-moving outlet glaciers in Alaska,
Antarctica, and Greenland rest on a basal layer of soft glacial till, which makes up a considerable part of
the basal sliding (Alley et al. 1987). The mechanical strength of glacial till obeys the Mohr-Coulomb
failure criterion and may be comparable to the masses of ow-type landslides treated in section Flow.
Although steady shear deformation may dominate, experience from seismic monitoring on landslides
shows that stickslip deformation may also occur. At the glacier beds of polythermal and temperate
glaciers, there may also be asperities or sticky spots, which induce stickslip basal sliding under the
otherwise nonseismic deformation conditions at the glacier bed.
In the following, we term seismic activity connected to the various processes of glacier ow and
observed by local networks GMEs (glacial microearthquakes).

Glacial Lake Outburst Floods at an Alpine Glacier


As an example of seismic activity of an alpine glacier, we review observations made at the polythermal
Gornergletscher, Monte Rosa Massif, Switzerland (Fig. 12). At the conuence of Gornergletscher and
Grenzgletscher, there is the ice-dammed Gorner lake (Fig. 31). The lake drains during annual glacial lake
outburst oods (GLOFs), which were monitored in the years 2004 and 2006 with seismic and geodetic
networks (Walter 2009; Walter et al. 2010). During the 2004 and 2006 eld seasons, about 35000 and
50000 events were detected. Generally, diurnal variations in the seismicity were observed, with the peak
in the early afternoon and much lower GME activity during the nighttime and morning hours. GME
clusters were recorded at three different depth levels: SURF_A and SURF_B at the depth of surface
crevasses, INT at ~100 m depth, and BAS just above the glacier bed (Fig. 31b). The SURF_A, SURF_B,
and INT GMEs have magnitudes between 2.3  Mw  1.6; the BAS GMEs reach Mw = 0.7.
With the exception of BAS GMEs, fault planes are steeply dipping, and strike directions of the interpreted
fault planes follow the general trend of surface crevasses (Fig. 31a). The radiation pattern of the SURF_A
Page 27 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 31 GMEs observed at the Gornergletscher before and during the GLOFs 2004 and 2006 (Walter 2009); (a) location of the
GME clusters; fault plane strike of the SURF A, SURF B, and INT GME clusters agree with the orientation of crevasses; dotted
line outlines approximately the Gorner lake at the maximum water levels; (b) focal depths of the different GME clusters

and INT LMEs can best be explained by the generation of tensile cracks, but the SURF_B events more
likely result from a DC mechanism. The BAS GMEs occur on subhorizontal fault planes and are well
explained by tensile cracks. The source mechanisms of the SURF_A, INT, and BAS cluster correspond to
the processes schematically outlined in Figs. 29 and 30. The BAS events occurred mainly during
otherwise quiescent night and early morning hours. This observation supports the idea of a collapsing
basal cavity due to the reduced meltwater access and basal water pressure (Fig. 30b). The occurrence of
DC-source mechanisms with the SURF_B GMEs may be due to the particular dynamic situation near the
Gorner lake and the GLOF.

Arctic Glacier Prone to Surges

The temperate Bering Glacier ows from the St. Elias mountain range to its terminus on the south-central
coast of Alaska and has a history of dramatic surges. Seismic monitoring was carried out between the
equilibrium line altitude and the terminus (Fig. 32a) and covered the early melting period from 20 April
until 19 June 2007 (West et al. 2010). Yet another surge started in 2008 (Burgess et al. 2012).
Around 160,000 events were detected during the observation period. Location of the GMEs was not
attempted. The waveforms of the GMEs (Fig. 32b) show a bimodal distribution and range from highfrequency signals (2035 Hz) to low-frequency signals (615 Hz). The high-frequency GMEs have
impulsive source functions and were interpreted as brittle failures. In an analogy to waveforms of volcanic
earthquakes, the low-frequency tremor GMEs were explained as a uid-driven crack model with a
resonant water-lled cavity as the seismic source. Hybrid GMEs were interpreted as hydrofracturing
followed by the rush of water into the new opening.
Relative maxima of the daily rate of low-frequency events precede the corresponding maxima of the
high-frequency events by 16 days (Fig. 32c). The interpretation of this pattern is that the peaks of the
low-frequency events reect the subglacial rush in of water leading to a decoupling of the glacier from its
bed and consequently enhanced ow velocities. The enhanced glacier motion is accompanied by an
increase of brittle failures. Velocity data would be essential to further develop this interpretation.

Page 28 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Fig. 32 Seismic monitoring at the Bering glacier (Modied after West et al. 2010); (a) location of seismic stations (red
triangles); (b) waveforms; (c) temporal evolution of daily rates of high and low frequency event

Stickslip Movement of an Antarctic Ice Stream


Bindschadler et al. (2003) reported on major West Antarctic ice streams discharging in sudden periods of
rapid motion, in particular Whillans Ice Stream (WIS) (Fig. 33a), which has an area of 200 km  100 km
with an average ice thickness of ~600 m. During December 2010 and January 2011 and again during
December 2011, Winberry et al. (2013) deployed a GPS and seismic network on the WIS in order to
monitor the slip nucleation and the subsequent period of rapid motion (Fig. 33b). Figure 33c shows the
displacements for a time span of ~1 h, covering the end of a preceding interevent period, the nucleation
phase, the rupture, the main-slip phase, and the transition to the following interevent period. Clear
evidence for the nucleation phase is shown only at one GPS station (station 1), where the rupture occurs
within ~5 min at all stations according to the rupture propagation speed of >500 m/s. The main-slip phase
lasts ~30 min and terminates almost simultaneously at all stations.
GMEs recorded at the individual stations correlate well in time and frequency domain with the ice
stream velocities (Fig. 34). At station 1, seismicity increases at the beginning of the nucleation phase but
only begins at the other stations around the rupture time. Two different styles of GMEs or microseismicity
can be distinguished. One type (Fig. 34a) is built by a sequence of fewer than a hundred to several
thousand individual impulsive events. The wide distances between the seismic stations did not allow for a

Page 29 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

b
1
2

0.4
4
0.3

Nucleation
Phase

Main-Slip
Phase

Interevent Period

Rupture
Interevent
Period

0.2

Displacement (m)

0.1

0
10

10
20
30
40
50
Minutes Since Slip Initiation

60

Fig. 33 Geodetic and seismic monitoring of an episodic high ow velocity phase at the Whillan Ice Stream (WIS); (a) location
map; (b) zoom of the monitoring area with locations of monitoring-stations; (c) displacements versus time since slip initiation;
numbers and colours assign displacement data to stations in (b) (Modied after Winberry et al. 2013)

Fig. 34 Waveforms observed at stations 14 at WIS; (a) impulsive, high frequency waveforms; (b) low frequency (Modied
after Winberry et al. 2013)
Page 30 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

precise location; however, the relative timing of P and S wave arrivals and the lack of crevasses is
consistent with source locations at or near the glacier bed. The individual events can be grouped into
families with a high degree of similarity between each other. The same event families persist in subsequent
high-slip-rate phases. This observation is an indication of the temporal stability of subsurface asperities
representing the source regions of the individual events. Other locations show in general a similar
behavior but are characterized by lower frequencies, indicating larger source areas. The lack of microseismic activity or GMEs at some locations may be explained by the presence of more continuous fault
gouge and higher water pressure at the glacier bed providing basal lubrication.
The second style of microseismicity is the emergent tremor (Fig. 34b). It is continuous and cannot be
resolved into individual events. The overall spectral power of this tremor correlates well with the glacial
ow velocities. Seismic energy during the tremor is concentrated in the horizontal components. A source,
radiating shear energy from near the glacier bed, is consistent with this observation. The spectra reveal
coherent gliding spectral lines, some of which are overtones of fundamental frequencies. Periodic
stickslip failure of small asperities could be the source mechanism of these seismic signals. Winberry
et al. (2013) argue that the highly repeatable WIS system represents an excellent natural laboratory which
could provide insight into friction and sliding complementary to laboratory or other eld studies (e.g.,
Gomberg et al. 2011).

Summary
A new class of global earthquakes (Ms > 4.5) was discovered and related to giant and catastrophic
landslides or phenomena occurring near the front of fast-moving outlet glaciers in Greenland and
Antarctica. The unique characteristics of these earthquakes are the depletion of short-period energy, the
elongated source time functions, and their radiation patterns, which can be explained by a single surface
force better than by seismic moments. Efcient detection and location methods were developed and force
histories F(t) along with the centroid single force (CSF) derived. The CSF is the product of the total
moving mass M and the displacement D of the centroid. It represents a major constraint on the whole
source process.
Originally, the new-class global earthquakes with sources on glaciers were interpreted as stickslip
ow behavior of large portions of outlet glaciers in Greenland and Antarctica. However, the postulated
movements of the ice masses could not to date be conrmed geodetically. An alternative source
mechanism which generates sufcient CSF and provides the appropriate duration of the source function
is the capsizing (or toppling) of large icebergs at the front of the outlet glaciers during major calving
events. The basic mechanics of this process are well understood, and correlations with glacier ow,
seasonal variations of iceberg production, and glacier retreat support the explanation of the majority of
these earthquakes coming about through this source process.
Related to landslide-generated global earthquakes, additional information from satellite images or local
reconnaissance helps to separate CSF into mass and total displacement. The inclination of the CSF vector
corresponds to the Fahrbschung; the duration of the landslide force history constrains the sliding
velocity. Application of a rotational slider block model allows, e.g., for estimates of the reduction of
the friction coefcient from static to dynamic conditions.
Microearthquakes related to the processes in landslide masses (LMEs) that evolve in a more steady
form or those in a post- or prefailure state have magnitudes of M < 0. Seismic networks deployed on or
very near to the landslide mass are necessary to capture these signals. A rich variety of waveforms can be
observed and classied by attributes derived from the frequency content, the duration, or the shape of the
envelope. Currently, no denite assignment of the different waveforms to landslide types and processes
Page 31 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

can be established. Impulsive events have been observed on slopes with steep or medium inclinations
(>20 ); tremor or harmonic LMEs dominate the seismic activity of gently sloping ows. Most seismic
activity on landslides is rather shallow. However, the location of a few events at the basal sliding zone
cannot be excluded. Rockfall and LME frequency correlate frequently with precipitation and precede
geodetically observed acceleration. An increase in seismic activity preceding acceleration without
discernible external triggers has also been observed. The general pattern of rockfall and LME activity
suggests that the DSGSD- and ow-type landslides are near their yield stress not only at the basal sliding
zone but also in their entire moving mass. Further progress in the understanding of LMEs would be
supported by denser seismic networks including borehole seismometers as a standard. Long-term seismic
monitoring covering seasonal variations and episodic high-velocity phases combined with complementary monitoring (e.g., geodetic, meteorological, hydrological, geomorphologic) are preconditions for
testing rened hypotheses about LMEs and their relation to landslide processes and evolution.
Monitoring of glacier ow-related microearthquakes (GMEs) by seismic networks deployed on
glaciers and supplemented by geodetic and meteorological devices provides a deeper insight into the
stress distribution within the glacier ice, glacier hydraulics, and processes at the glacier bed or within a
basal layer of soft sediments. The latter aspect can be exemplarily studied at the Whillans Ice Stream
(WIS) in Antarctica. The ice stream exhibits a stickslip behavior, controlled by tides with a CSF
corresponding to Ms = 5.6. The duration of this process is above the seismically observable range, but
GMEs generated by this process yield unique information about the spatial distribution and mechanical
behavior of asperities or sticky spots in the basal layer of soft sediments. Research in this eld does not
only support the understanding of glacier sliding but addresses generally the friction behavior of fault
gouges, e.g., the basal sliding zones of landslides or active tectonic faults.
Finally, we touch on how the potential monitoring and analyzing of landslide- and glacier-generated
earthquakes contribute to prediction and early warning of catastrophic events. Global landslide or glacial
earthquakes always represent the nite state of a development and may message a disaster. Prediction and
early warning is not provided by the detection of these earthquakes; however, instantaneous location and
quantication by seismic methods may support mitigation measures and warnings against subsequent
hazards like damming of the valley (see Fig. 6c) or increased density of icebergs. LMEs have been proven
to indicate a dangerous acceleration of a landslide, usually earlier than the geodetically observable
displacement. GMEs precede glacier break-off and glacial lake outburst oods (GLOFs). The results
from basic research on LMEs and GMEs lead to a recommendation of seismic monitoring as an essential
component of an integrated early warning system for landslides and glacial hazards.

Cross-References
Earthquake Location
Earthquake Mechanism Description and Inversion
Mechanisms of Earthquakes in Iceland
Non-Double-Couple Earthquakes
Rockfall Seismicity Accompanying Dome Building Eruptions
Volcanic Tremor

Page 32 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

References
Alley RB, Blankenship DD, Bentley CR, Rooney ST (1987) Till beneath Ice Stream B. Till deformation:
evidence and implications. J Geophys Res 92(B9):89218929
Amitrano D, Grasso JR, Senfaute G (2005) Seismic precursory patterns before a cliff collapse and critical
point Phenomena. Geophys Res Lett 32, L08314. doi:10.1029/2004GL022270
Bindschadler RA, King MA, Alley EB, Anandakrishnan S, Padman L (2003) Tidally controlled stickslip
discharge of a West Antarctic ice stream. Science 301(5636):10871089. doi:10.1126/
science.1087231
Brantley SR, Myers B (2000) Mount St. Helens from the 1980 eruption to 2000. U.S. Geological Survey
Fact Sheet, FS-036-00, 2 p. http://pubs.er.usgs.gov/publication/fs03600
Br
uckl E, Parotidis M (2005) Prediction of slope instabilities due to deep-seated gravitational creep. Nat
Hazards Earth Syst Sci 5:155172. SRef-ID: 1684-9981/nhess/2005-5-155
Bruckl E, Brunner FK, Lang E, Mertl S, M
uller M, Stary U (2013) The Gradenbach observatorymonitoring deep-seated gravitational slope deformation by geodetic, hydrological, and seismological
methods. Landslides. doi:10.1007/s10346-013-0417-1
Burgess EW, Forster RR, Larsen CF, Braun M (2012) Surge dynamics on Bering Glacier, Alaska, in
20082011. Cryosphere 6:12511262. doi:10.5194/tc-6-1251-2012
Chen X, Shearer PM, Walter F, Fricker HA (2011) Seventeen Antarctic seismic events detected by global
surface waves and a possible link to calving events from satellite images. J Geophys Res 116, B06311.
doi:10.1029/2011JB008262
Cuffey KM, Paterson WSB (2010) The physics of glaciers, 4th Edition. Academic Press, ISBN
978012333694614, p 704
Eberhardt E, Stead D, Coggan JS (2004) Numerical analysis of initiation and progressive failure in natural
rock slopes the 1991 Randa rockslide. Int J Rock Mech Min Sci 41:6987
Ekstrm G, Stark CP (2013) Simple scaling of catastrophic landslide dynamics. Science
339(6126):14161419. doi:10.1126/science.1232887
Ekstrm G, Nettles M, Abers GA (2003) Glacial earthquakes. Science 302:22624. doi:10.1126/
science.1088057
Faillettaz J, Funk M, Sornette D (2011) Icequakes coupled with surface displacements for predicting
glacier break-off. J Glaciol 57(203):453460
Gischig V, Moore JR, Evans KF, Amann F, Loew S (2011) Thermomechanical forcing of deep rock slope
deformation: 2. The Randa rock slope instability. J Geophys Res 116:F04011. doi:10.1029/
2011JF002007
Gomberg J, Schulz W, Bodin P, Kean J (2011) Seismic and geodetic signatures of fault slip at the
Slumgullion Landslide Natural Laboratory. J Geophys Res 116, B09404. doi:10.1029/2011JB008304
Goodman RE, Kieffer DS (2000) Behaviour of rock in slopes. J Geotech Geoenviron 128(8):675684
Helmstetter A, Garambois S (2010) Seismic monitoring of Schilienne rockslide (French Alps): analysis
of seismic signals and their correlation with rainfalls. J Geophys Res 115, F03016. doi:10.1029/
2009JF001532
Kanamori H, Given JW (1982) Analysis of longperiod seismic waves excited by the May 18, 1980,
eruption of Mount St. Helens a terrestrial monopole? J Geophys Res 87:54225432. doi:10.1029/
JB087iB07p05422. ISSN: 01480227
Kawakatsu H (1989) Centroid single force inversion of seismic waves generated by landslides. J Geophys
Res 94:12,36312,374. doi:10.1029/89JB01118. ISSN: 01480227
Mertl S (2015) Characterization of local seismic events on the deep-seated events gravitational slobe
deformation Gradenbach. PhD thesis, Vienna University of Technology, p 176
Page 33 of 34

Encyclopedia of Earthquake Engineering


DOI 10.1007/978-3-642-36197-5_385-1
# Springer-Verlag Berlin Heidelberg 2015

Moore J, Albee W (1981) Topographical and structural changes, March-July 1980 photogrammetric
data. In: Lipman P, Mullineaux D (eds) The 1980 eruptions of Mount St. Helens Washington.
Geological Survey Professional Paper 1250, pp 123134
Nettles M, Ekstrm G (2010) Glacial earthquakes in Greenland and Antarctica. Annu Rev Earth Planet
Sci 38:467491. doi:10.1146/annurev-earth-040809-152414
Nettles M, Larsen TB, Elsegui P, Hamilton GS, Stearns LA, Ahlstrm AP, Davis JL, Andersen ML, de
Juan J, Khan SA, Stenseng L, Ekstrm G, Forsberg R (2008) Step-wise changes in glacier ow speed
coincide with calving and glacial earthquakes at Helheim Glacier, Greenland. Geophys Res Lett
35, L24503. doi:10.1029/2008GL036127
Spillmann T, Maurer H, Green AG, Heincke B, Willenberg H, Husen S (2007) Microseismic investigation
of an unstable mountain slope in the Swiss Alps. J Geophys Res 112, B07301. doi:10.1029/
2006JB004723
Tsai VC, Ekstrm G (2007) Analysis of glacial earthquakes. J Geophys Res 113:F03S22. doi:10.1029/
2006JF000596
Tsai VC, Rice JR, Fahnestock M (2008) Possible mechanisms for glacial earthquakes. J Geophys Res
113, F03014. doi:10.1029/2007JF000944
Voight B (1981) Time scale for the rst moments of the May 18 eruption. In: Lipman P, Mullineaux
D (eds) The 1980 eruptions of Mount St. Helens Washington. Geological Survey Professional Paper
1250, pp 6886
Walter F (2009) Seismic activity on Gornergletscher during Gornersee outburst oods. PhD thesis, DISS.
ETHNo 18184, ETH Zrich, p 135
Walter F, Dreger DS, Clinton JF, Deichmann N, Funk M (2010) Evidence for near-horizontal tensile
faulting at the base of Gornergletscher, a Swiss Alpine Glacier. Bull Seismol Soc Am 100(2):458472.
doi:10.1785/0120090083
Walter M, Arnhardt C, Joswig M (2012) Seismic monitoring of rockfalls, slide quakes, and ssure
development at the Super-Sauze mudslide, French Alps. Eng Geol 128:1222
Weginger S (2012) Detektion und Lokalisierung seismischer Signale zur berwachung der
Massenbewegung Steinlehnen. Master thesis, Vienna University of Technology, p 96
West ME, Larsen C, Truffer M, ONeel S, LeBlanc L (2010) Glacier microseismicity. Geology
38(4):319322. doi:10.1130/G30606.1
Winberry JP, Anandakrishnan S, Wiens DA, Alley RB (2013) Nucleation and seismic Tremor associated
with the glacial earthquakes of Whillans Ice Stream, Antartica. Geophys Res Lett 40:312315.
doi:10.1002/grl.50130
Zangerl C, Eberhardt E, Schnlaub H, Anegg J (2007) Deformation behavior of deep-seated rockslides in
crystalline rock. In: Rock mechanics: proceedings of the 1st Canada-US rock mechanics symposium,
Vancouver, 2731 May 2007. doi: 10.1201/NOE0415444019-c112
Zischinsky U (1969) ber Sackungen. Rock Mech 1:3052

Page 34 of 34

Vous aimerez peut-être aussi