ArticlePDF Available

Slip pulse and resonance of Kathmandu basin during the 2015 Mw 7.8 Gorkha earthquake, Nepal imaged with geodesy

Authors:

Abstract

Detailed geodetic imaging of earthquake rupture enhances our understanding of earthquake physics and induced ground shaking. The April 25, 2015 Mw 7.8 Gorkha, Nepal earthquake is the first example of a large continental megathrust rupture beneath a high-rate (5 Hz) GPS network. We use GPS and InSAR data to model the earthquake rupture as a slip pulse of ~20 km width, ~6 s duration, and with peak sliding velocity of 1.1 m/s that propagated toward Kathmandu basin at ~3.3 km/s over ~140 km. The smooth slip onset, indicating a large ~5 m slip-weakening distance, caused moderate ground shaking at high >1Hz frequencies (~16% g) and limited damage to regular dwellings. Whole basin resonance at 4-5 s period caused collapse of tall structures, including cultural artifacts. Copyright © 2015, American Association for the Advancement of Science.
directly. The diluted filtrate could be augmented
with TPGS-750-M to the original level (2 wt %) and
reused, thereby creating little to no wastewater
stream. The environmental factor (E factor) (21),
ametricofgreennessthat has previously been
applied to micellar catalysis (22), is very low (E
factor = 3).
We used ICP to analyze the palladium content
(<10 ppm) of a product formed via the technology
presented here, and we compared the result with
that quantified following atraditionalSuzuki-
Miyaura coupling in organic solvent (fig. S4). Re-
sidual palladium in the product derived from a
standard coupling in dioxane was far higher than
that observed using our nanoparticle approach.
Prospects for incorporating this water-based
nanomicelle-nanometal technology into a one-
pot sequence of reactions are shown in Fig. 5A.
Heteroaryl iodide 4, containing carbamate and
trimethylsilyl (TMS) protecting groups, was gen-
erated in situ and then subjected to cross-coupling
with alkenyl tetrafluoroborate salt 5,usingthe
Feppm Pd nanoparticle protocol. The coupling
product 6was then exposed to aqueous base to
remove the TMS groups and effect elimination to
7, followed by butoxycarbonyl (Boc) deprotection
to 8. Final aryl amination with bromobenzene to
9provided entry to the bioactive class of 2,4,5-
substituted pyrazol-3-one compounds in a one-
pot sequence with an overall isolated yield of
68% (23).
In addition, testing the potential for this mixed-
metal catalyst system to effect other important
Pd-catalyzed reactions, such as Sonogashira cou-
plings, was carried out (in the absence of added
copper), following the example illustrated in
Fig. 5B. The prognosis for a similar outcome is
good.
REFERENCES AND NOTES
1. C. C. C. Johansson Seechurn, M. O. Kitching, T. J. Colacot,
V. Snieckus, Angew. Chem. Int. Ed. 51, 50625085 (2012).
2. H. Li, C. C. C. Johansson Seechurn, T. J. Colacot, ACS Catal. 2,
11471164 (2012).
3. J. Kooroshy, C. Meindersma, R. Podkolinski, M. Rademaker,
T. Sweijs, A. Diederen, M. Beerthuizen, S. de Goede, Scarcity
of Minerals: A Strategic Security Issue (No. 02/01/10, The
Hague Center for Strategic Studies, The Hague, Netherlands,
2009).
4. F.-S. Han, Chem. Soc. Rev. 42, 52705298 (2013).
5. J. C. Tellis, D. N. Primer, G. A. Molander, Science 345, 433436
(2014).
6. S. K. Gurung, S. Thapa, A. Kafle, D. A. Dickie, R. Giri, Org. Lett.
16, 12641267 (2014).
7. Y. Zhou, W. You, K. B. Smith, M. K. Brown, Angew. Chem.
Int. Ed. 126, 35433547 (2014).
8. M. R. Netherton, G. C. Fu, Adv. Synth. Catal. 346, 15251532
(2004).
9. G. D. Allred, L. S. Liebeskind, J. Am. Chem. Soc. 118,
27482749 (1996).
10. B. H. Lipshutz et al., J. Org. Chem. 76, 43794391 (2011).
11. A. J. J. Lennox, G. C. Lloyd-Jones, Chem. Soc. Rev. 43,
412443 (2014).
12. N. A. Isley, F. Gallou, B. H. Lipshutz, J. Am. Chem. Soc. 135,
1770717710 (2013).
13. S. J. Lee, T. M. Anderson, M. D. Burke, Angew. Chem. Int. Ed.
49, 88608863 (2010).
14. G. A. Molander, Y. Yokoyama, J. Org. Chem. 71, 24932498
(2006).
15. S. Darses, J.-P. Genet, Chem. Rev. 108, 288325 (2008).
16. T. H. Bointon et al., Nano Lett. 14, 17511755 (2014).
17. S. L. Buchwald, C. Bolm, Angew. Chem. Int. Ed. 48, 55865587
(2009).
18.C.L.Chun,D.R.Baer,D.W.Matson,J.E.Amonette,R.L.Penn,
Environ. Sci. Technol. 44,50795085 (2010).
19. Y.-P. Sun, X. Q. Li, J. Cao, W.-X. Zhang, H. P. Wang, Adv. Colloid
Interface Sci. 120,4756 (2006).
20. I. Thomé, A. Nijs, C. Bolm, Chem. Soc. Rev. 41, 979987
(2012).
21. R. A. Sheldon, Green Chem. 9, 1273 (2007).
22. B. H. Lipshutz, N. A. Isley, J. C. Fennewald, E. D. Slack, Angew.
Chem. Int. Ed. 52, 1095210958 (2013).
23. P. N. Dube et al., Chem. Biol. Drug Des. 84, 409419 (2014).
ACKNOW LEDGM ENTS
We thank Novartis for financial support; J. Feng for technical
assistance; M. Cornish for acquisition of the AFM images;
S. Kraemer for obtaining the cryo-TEM, SEM, and EDX data; and
J. Matthey for providing Pd salts. Parts of this work were carried
out in the Characterization Facility, University of Minnesota, which
receives partial support from NSF through the Materials Research
Science and Engineering Center (MRSEC) program. This work also
made use of the University of CaliforniaSanta Barbara (UCSB)
Materials Research Laboratory Central Facilities, supported by
NSFs MRSEC program under award no. DMR-1121053. ICP-MS
analyses were provided by J. Reilly (Novartis, Cambridge, MA). We
also acknowledge support from NIH in the form of a Shared
Instrument Grant to UCSB(1S10OD012077-01A1). A preliminary
patent covering this chemistry has been filed by the University of
CaliforniaSanta Barbara. The exp erimental data r eported in this
paper are availab le in the supplemen tary materials.
SUPPLEMENTARY MATERIALS
www.sciencemag.org/content/349/6252/1087/suppl/DC1
Materials and Methods
Figs. S1 to S18
Tables S1 to S15
References (2434)
13 June 2015; accepted 30 July 2015
10.1126/science.aac6936
NATURAL HAZARDS
Slip pulse and resonance of the
Kathmandu basin during the 2015
Gorkha earthquake, Nepal
J. Galetzka,
1
*D. Melgar,
2
J. F. Genrich,
1
J. Geng,
3
S. Owen,
4
E. O. Lindsey,
3
X. Xu,
3
Y. Bock,
3
J.-P. Avouac,
5,1
L. B. Adhikari,
6
B. N. Upreti,
7
B. Pratt-Sitaula,
8
T. N. Bhattarai,
9
B. P. Sitaula,
9
A. Moore,
4
K. W. Hudnut,
10
W. Szeliga,
11
J. Normandeau,
12
M. Fend,
12
M. Flouzat,
13
L. Bollinger,
13
P. Shrestha,
6
B. Koirala,
6
U. Gautam,
6
M. Bhatterai,
6
R. Gupta,
6
T. Kandel,
6
C. Timsina,
6
S. N. Sapkota,
6
S. Rajaure,
6
N. Maharjan
6
Detailed geodetic imaging of earthquake ruptures enhances our understanding of earthquake
physics and associated ground shaking.The 25 April 2015 moment magnitude 7.8 earthquake in
Gorkha, Nepal was the first large continental megathrust rupture to have occurred beneath a
high-rate (5-hertz) Global Positioning System (GPS) network. We used GPS and interferometric
synthetic aperture radar data to model the earthquake rupture as a slip pulse ~20 kilometers in
width, ~6 seconds in duration, and with a peak sliding velocity of 1.1 meters per second, which
propagated toward the Kathmandu basin at ~3.3 kilometers per second over ~140 kilometers.
The smooth slip onset, indicating a large (~5-meter) slip-weakening distance, caused moderate
ground shaking at high frequencies (>1 hertz; peak ground acceleration, ~16% of Earths
gravity) and minimized damage to vernacular dwellings. Whole-basin resonance at a period of
4 to 5 seconds caused the collapse of tall structures, including cultural artifacts.
The shape of the slip-rate time function
(STF) during a seismic rupture provides
critical insight into the constitutive fault
properties. The abruptness of the slip on-
set determines the high-frequency con-
tent of the STF, and hence the intensity of the
near-field ground motion (1), whereas the tail,
which discriminates pulse-like and crack-like rup-
tures (2), has a low-frequency signature. Therefore,
resolving the STF with band-limited strong-motion
recordsisdifficult.Combininghigh-rateGlobal
Positioning System (GPS) waveforms (3,4), which
capture both dynamic and permanent deforma-
tion,overcomesthislimitation.
The 25 April 2015 moment magnitude (M
w
)
7.8 earthquake in Gorkha, Nepal resulted from
the unzipping of the lower edge of the locked
portion of the Main Himalayan Thrust (MHT)
fault, along which the Himalayan wedge is thrust
over India (5). The earthquake nucleated ~80 km
northwest of Kathmandu and ruptured a 140-km-
longsegmentofthefault(Fig.1A),withahypo-
central dept h of ~15 km and a dip angle of 7°
to 12° (5,6). The MHT accommodates most of
the convergence between India and southern
Tibet, with a convergence rate between 17 and
21 mm/year (7). For the 2015 event, which re-
sulted in over 8000 deaths (mostly in Kathmandu
and adjacent districts), modified Mercalli inten-
sities (MMIs) reported by the National Society
for Earthquake TechnologyNepal (NSET) (8)
reached up to IX (violent shaking) and exceeded
VI (strong shaking) over an area 170 km by 40 km.
Kathmandu has been struck by repeated earth-
quakes in the past, with major destruction [MMI >
X (extreme shaking)] in the years 1255, 1344, 1408,
1681, 1833, and 1934 (911). These earthquakes
all occurred close to Kathmandu and have been
SCIENCE sciencemag.org 4SEPTEMBER2015VOL 349 ISSUE 6252 1091
RESEARCH |REPORTS
on September 20, 2015www.sciencemag.orgDownloaded from on September 20, 2015www.sciencemag.orgDownloaded from on September 20, 2015www.sciencemag.orgDownloaded from on September 20, 2015www.sciencemag.orgDownloaded from on September 20, 2015www.sciencemag.orgDownloaded from
assigned magnitudes between M
w
7.5 and 8.4.
During the Gorkha earthquake, damages in the
Kathmandu basin were probably amplified by site
effects, as has happened in past events (12,13).
The basin is filled with 500 to 600 m of fluvio-
lacustrine sediments resting on a metamorphic
basement (14).
The damage to the most vulnerable vernac-
ular dwellings in Kathmandu, which rarely ex-
ceed four stories, was much less than expected
in view of the 2015 earthquakes magnitude
and its proximity to Kathmandu. In contrast,
some taller structures were more severely af-
fected, such as the 60-m-tall Dharahara tower,
which collapsed even though it had partially sur-
vived an M
w
8.1 to 8.4 earthquake in 1934. The
1934 event caused much more extensive de-
struction to vernacular dwellings in Kathmandu
than the 2015 event did: 20% of the buildings in
Kathmandu were destroyed in 1934 versus less
than 1% in 2015 (15).Theseobservationsreflect
the combined effects of the earthquake source
characteristics and local geological conditions, in
addition to the evolution of building practices.
The 2015 Gorkha earthquake ruptured a sub-
horizontal portion of the MHT that lies directly
beneath a network (16) of continuous GPS (cGPS)
stations, which record data at a high rate of
five samples per second, and one accelerometer
station (17) (Fig. 1A). In addition, surface displace-
ments were measured with interferometric syn-
thetic aperture radar [InSAR (18,19)] (fig. S1).
Although a number of recent earthquakes have
been documented with similar techniques (20,21),
the Gorkha event is the first occurrence of a large
continental thrust earthquake to be recorded by
high-rate cGPS stations very close to and com-
pletely encompassing the rupture area. The
combination of these measurements provides
the opportunity to image the kinematics of the
source process and the strong ground motion
that led to the particular pattern of structural
damage observed during this earthquake.
The records of seismic displacements and ac-
celerations (Fig. 2 and fig. S2) show southward
motion of up to 2 m, with a rise time on the
order of 6 s. The pulse is particularly clear at
cGPS station KKN4, located on bedrock just
north of Kathmandu and only ~13 km above the
fault. The displacement at this station started
about 25 s after the onset of the rupture, corre-
sponding to 15 s after the P-wave arrival (Fig. 2);
it reached its final static value by about 32 s,
based on the origin time of 06:11:26.270 UTC
determined by the U.S. Geological Survey (USGS)
from the arrival of radiated direct Pwaves (6).
The records indicate a pulse-like rupture (22),
with slip on any given portion of the fault oc-
curring over a short fraction of the total ~70-s
earthquake source duration (5). Given the ~78-km
distance of KKN4 from the epicenter, the pulse
must have propagated at ~3 km/s, a value con-
sistent with waveform modeling and back pro-
jection of high-frequency seismic waves recorded
at teleseismic distances (5). Surface velocities
reached values of ~0.7 m/s. In addition to the
pulse recorded at KKN4, the cGPS station NAST
within the Kathmandu basin detected strong
oscillations of about 3- to 4-s periods lasting for
~20 s (Fig. 2 and Fig. 3A). The Gorkha earthquake
1092 4 SEPTEMBER 2015 VOL 349 ISSUE 6252 sciencemag.org SCIENCE
1
Department of Geology and Planetary Sciences, California
Institute of Technology (Caltech), Pasadena, CA 91125, USA.
2
Berkeley Seismological Laboratory, University of California
(UC)Berkeley, Berkeley, CA 94720, USA.
3
Cecil H. and Ida M.
Green Institute of Geophysics and Planetary Physics, Scripps
Institution of Oceanography, UCSanDiego,LaJolla,CA92037,
USA.
4
Jet Propulsion Laboratory (JPL), Caltech, Pasadena, CA
91109, USA.
5
Department of Earth Sciences, University of
Cambridge, Cambridge CB2 3EQ, UK.
6
Department of Mines
and Geology, Lainchour, Kathmandu, Nepal.
7
Nepal Academy of
Science and Technology, Khumaltar, Lalitpur, Nepal.
8
Department of Geological Sciences, Central Washington
University (CWU), Ellensberg, WA 98926, USA.
9
Tri-Chandra
Campus, Tribhuvan University, Ghantaghar, Kathmandu, Nepal.
10
U.S. Geological Survey (USGS), Pasadena, CA 91106, USA.
11
Pacific Northwest Geodetic Array and Department of
Geol ogical Sciences, CWU, Ellensberg, WA 98926, USA.
12
UNAVCO,
Boulder, CO 80301, USA.
13
Département Analyse et Sureveillance
de lEnvironnement (DASE), Commissariat à lEnergie Atomique
(CEA), 91297 Bruyères-le-Châtel, Arpajon, France.
*Present address: UNAVCO, Boulder, CO 80301, USA.
Corresponding author. E-mail: avouac@gps.caltech.edu
84.0° 86.0°
85.0°
27.0°
28.0°
29.0°
1m
INDIA
NEPAL
CHINA
5km
15km
25km
KKN4
NAST
SNDL
CHLM
KATNP
KIRT
DAMA
SYBC
RMTE
BELT
BESI
DMAU
GHER
0
2
4
6
Slip (m)
27.0°
28.0°
29.0°
0 km 100 km
5km
15km
25km
INDIA
NEPAL
CHINA
Kathmandu
−8
−4
0
4
8
Stress drop (MPa)
12/05/15
Mw 7.3
Fig. 1. Cumulative slip distribution of and static stress drop due to the Gorkha earthquake. (A)
Slip inversion results for the M
w
7.8 Gorkha event. The red star is the hypocenter. Dashed contours are
depths to the fault. Orange diamonds are 5-HzcGPS stations, and white diamonds are low-rate (1/30-Hz)
stations. The green triangle is the strong-motion station. Kathmandu is represented by the blue square.
The black arrows indicate the coseismic offsets measured at the sites (the values and uncertainties are
given in table S1). Vectors with less than 10 cm of displacement are not shown. (B)Staticstressdrop
predicted by the model of Fig. 1A. Green circles are aftershocks with local magnitudes greater than four,
recorded and located by the Nepal National Seismic Center. Focal mechanisms (yellow and white circles)
represent the global centroid-moment tensor solutions for aftershocks with magnitudes greater than six.
RESEARCH |REPORTS
must have excited a resonance of the Kathmandu
basin as a whole. The resonance is evident in
the response spectra from these stations and in
data from the accelerometer station, KATNP
(Fig. 3, G to I).
To determine the kinematics of the seismic
rupture, we carried out a formal inversion of
time-dependent slip on the fault (23,24) and
compared the recorded waveforms with for-
ward predictions, assuming a propagating slip
pulse with varied characteristics. We assumed
a planar fault geometry with a strike of 295° and
a dip of 11°, in accordance with the teleseismic
W-phase moment tensor solution calculated by
the USGS (6). We tested shallower dips up to
7° but found that 11° provided a better fit to
the data. The fault was discretized into 10 km by
10 km subfault segments. We jointly inverted the
three-component 5-Hz GPSderived velocity wave-
forms, the GPS static offsets, and the InSAR line
of sight (LOS) static displacements measured
between 22 February and 3 May (fig. S1). The
GPS displacement time series shows large post-
seismic motion at only one station (CHLM), with
a magnitude of less than 2 cm in both the hor-
izontal and vertical directions over the week af-
ter the earthquake. Therefore, we neglected the
contribution of postseismic deformation to the
LOS displacements. The model fits both data
sets closely (Fig. 1A), with an 86% reduction in
variance for the InSAR and GPS coseismic dis-
placements and a 74% reduction in variance
for the GPS velocity waveforms (figs. S2 and
S4). The model indicates a predominantly uni-
lateral rupture to the southeast, with a peak
slip of ~6.5 m on a large asperity to the north of
Kathmandu. The event duration was 65 s (fig.
S4), with the peak moment release at 23 s when
the slip pulse was less than 10 km north of
Kathmandu (movie S1); the peak slip rate was
1.1 m/s. Most of the slip was concentrated within
a narrow region between the 10- and 20-km fault-
depth contours. We found a large asperity with
3.0 m of slip, located east of the main asperity
and between 20 and 23 km below the surface.
The rupture velocity of the propagating slip pulse,
indicated by the onset of slip in our best-fitting
model, was ~3.2 km/s and had a maximum al-
lowed velocity of 3.3 km/s (fig. S4). This velocity
corresponds to ~95% of the shear wave speed at
the depth where the majority of the slip occurred
(15 km), according to the local velocity model
used to calculate the Greensfunctions(tableS2),
which indicates a very fast rupture propagation.
The slip tapered at 17 to 20 km depth along the
edge of the locked zone of the MHT.
The inversion that we performed includes a
large number of parameters, which would allow
for a relatively complex rupture history. However,
the resulting model is simple, with essentially a
single propagating slip pulse. The spatiotemporal
evolution of the slip pulse matches well with the
location of the sources of high-frequency seismic
waves (0.5 to 2 Hz) derived from back projection
of the teleseismic waveforms (5)(movieS1).
We calculated the static stress change on the
fault plane due to the earthquake (Fig. 1B). This
SCIENCE sciencemag.org 4SEPTEMBER2015VOL 349 ISSUE 6252 1093
−2
0
North
KKN4 (m)
−1
0
NAST (m)
−1
1
KATNP (m /s 2)
0.0
0.5
East
0.0
0.5
−1
1
020 40 60
0.0
1.5
Up
020 40 60
Seconds after P-wave arrival
0
1
020 40 60
−1
1
Fig. 2. Records of ground displacements and accelerations during the Gorkha earthquake.
Shown are displacement waveforms at cGPS stations KKN4 and NAST (five samples per second)
and acceleration waveforms at strong-motion station KATNP (Fig. 1).
Fig. 3. Evidence for resonance of the
Kathmandu basin. (Ato C) Three compo-
nents of ground velocity observed at two
high-rate GPS stations (KKN4 and NAST)
and one strong-motion station (KATNP) in
the Kathmandu region. KKN4 is located on
hard rock northwest of Kathmandu, whereas
the other two stations are located on soft
sediment in the basin.The GPS is differen-
tiated to velocity, and the strong-motion
data are integrated after high-pass filtering at 0.02 Hz. (Dto F) Ground-motion amplification observed at the two basin stations. Plotted is the ratio of the
amplitude spectra of the basin stations to the amplitude spectra of the reference bedrock station, KKN4. (Gto I) Five-percent damped velocity response
spectra for all three stations. (J) Close-up map showing the location of the basin and bedrock stations.
RESEARCH |REPORTS
calculation showed loading of the fault around
the main asperity where most of the aftershocks
occurred, including the M
w
7.3 aftershock of 12
May, as expected for aftershocks triggered by
coseismic stress transfer (25). The model pre-
dicted a pattern of uplift in the Kathmandu basin
and subsidence at the front of the high range (fig.
S4), approximately opposite to the pattern ob-
served in the interseismic period, as expected
from simple models of the seismic cycle on the
MHT (26,27).
The record at station KKN4 should be a close
representation of the STF, as it lies only about
13 km above the propagating slip pulse and is
not affected by the site effects observed at the
stations in the Kathmandu basin. We conducted
synthetic tests with the same Earth structure
model used in the inversion (table S1) to assess
the distortion and smoothing introduced by the
elastic half-space response (fig. S5). We found a
vertical velocity amplitude of about 70% of the
peak slip rate on the fault directly beneath the sta-
tion, along with a well-preserved temporal shape.
Furthermore, the tests demonstrated that the
smooth onset of slip is not an artifact of the trans-
fer through the elastic medium, represented by
the elastodynamic Greensfunctions.Theshapeof
the slip pulse can also be retrieved from the GPS
records at NAST and the strong-motion vertical
records at KATNP, which are less affected by site
effects than the horizontal records (Fig. 1). All
three records indicate a pulse ~6 s in duration.
The shape of the pulse fits the regularized Yoffe
function (28), yielding a smooth rise, with an
acceleration time to the peak slip rate of t
s
=1.7s,
a rise time of t
R
= 3.3 s, and a total effective du-
ration of t
eff
= 6.7 s. The slip-rate pulse derived
from the inversion also fits well, using the same
values of t
s
and t
R
and peak slip-rate of ~0.9 m/s
(Fig. 4). We compared the recorded waveforms
with predictions from a suite of forward models
to test the robustness of our results. We used the
static slip model in these tests, deduced from the
inversion of the GPS static and InSAR measure-
ments (fig. S7). We assumed a propagating slip
pulse and a regularized Yoffe STF with varying
characteristics. We varied the rupture velocity
between 2.8 and 3.6 km/s and the rise time be-
tween 2 and 10 s (fig. S8). By inverting synthetics
calculated from forward modeling, we also tested
the resolution power of the inversion and the
limitedbiasintroducedbytheregularization
applied to the inversion (24) (figs. S10 and S11).
Together, these tests demonstrated that the du-
ration of the slip pulse was probably less than
10 s, the time to the peak slip rate could not have
been shorter than 1 s (we would otherwise have
observed a much larger amplitude at high fre-
quencies), and the average propagation rate of
the slip pulse was not less than ~3.0 km/s over
the first 30 s (until KKN4, NAST, and KATNP
recorded a pulse signal).
Tinti et al. (28) analyzed how the shape of
the STF relates to the characteristics of the
friction law that governs the dynamics of the
rupture. Based on this rationale (their equations
6 and 11), we estimated the slip-weakening dis-
tancetobe~5m(forapeakslipof6.5m).This
distance is large compared with those estimated
from kinematic and dynamic modeling of seis-
mic ruptures (29,30), which tend to be over-
estimated (1) and are typically on the order of
0.5to1m.Thelargevalueweobtainedispos-
sibly related to the earthquakeshavingoccurred
close to the brittle-ductile transition at the lower
edge of the locked portion of the MHT. The mod-
eled smooth onset of the STF and the related large
slip-weakening distance provide an explanation
oftherelativelylowamplitudeofshakingatfre-
quencies above 1 Hz. The observed slip-weakening
behavior does not require the slip-weakening
friction law to be in effect: A fault obeying the
rate-and-state friction law can show an appar-
ent slip-weakening behavior with an effective
critical distance that is several orders of mag-
nitude larger than the critical distance entering
the friction law (31). Aspects of the rupture kine-
matics and ground strong motion observed during
the Gorkha event may also be due to hanging wall
effects, the importance of which could be assessed
through dynamic modeling of the rupture (32,33).
Our study provides insight into the main
factors that determined the damage sustained
during the Gorkha earthquake. Although the
hypocenter was ~80 km away from the city, the
main asperity that radiated most of the energy
was much closer, just north of the basin and at
a relatively shallow depth. Comparison of the
waveforms recorded within the sedimentary
basin at NAST and KATNP (Fig. 3) with the
bedrock records at KKN4 shows prominent
differences, even though the stations are less
than 13 km apart. The waveforms recorded at
the bedrock station KKN4 were simple, mostly
dominated by the single pulse, whereas within
the basin, peak horizontal ground velocities of
0.5 to 0.8 m/s [considered severe to violent (34)]
were sustained for 20 s at KATNP and 40 s at
NAST. The ratio of the amplitude spectra of the
basin waveforms to those recorded at the bed-
rock station (Fig. 3, D to F) indicates an ampli-
fication of long-period energy between 1 and 9 s,
with horizontal-direction amplitudes in the basin
six to seven times as large as those at the bedrock
station. The response spectra (Fig. 3, G to I) show
that, within this amplified period band, the 4-s-
period shaking was the strongest at the basin
stations.
The 4-s peak in the response spectra agrees
with the observation that the source time func-
tion beneath Kathmandu had a duration of ~6
to 7 s. The net effect of this long source duration
with a slow onset time was to produce radiated
energy depleted in the high-frequency component
(fig. S11). This explains why vernacular dwellings
with only a few stories were not severely affected,
despite the anticipated short-period site effects
from microzoning (13). Furthermore, high-frequency
intensity measures such as peak ground accel-
erations (Fig. 2) were modest (~1.6 m/s
2
, MMI =
VI), whereas longer-period intensity measures
such as peak ground velocity (Fig. 3) were very
large (80 cm/s, MMI = IX). Kathmandu was faced
with a combination of source and site effects. The
rupture directivity focused radiated seismic ene rgy
toward the city; the smooth onset and 6- to 7-s
duration of the pulse excited a resonance of the
Kathmandu basin, producing a protracted dura-
tion of violent shaking at a period of around 4 s.
1094 4SEPTEMBER2015VOL 349 ISSUE 6252 sciencemag.org SCIENCE
84.0°
27.0°
84.0 85.0° 86.0° 87.0°
28.0°
0 km 100 km
Kathmandu
5km
15km
25km
t=27.0s
0.00 0.25 0.50 0.75 1.00
Slip rate (m/s)
0.67a
1.03b
Slip rate (m/s)
0.98c
Ti m e( s)
0.53d
10 20 30 40
Time (s)
0.43e
Slip model
Yof fe
abcde
Fig. 4. Slip-pulse kinematics during the Gorkha earthquake. (A) Snapshot of the slip rate on the
MTH at 27 s after the origin time, during propagation of the seismic rupture from the model in Fig. 1.
The red star is the hypocenter, and dashed contours represent the depth to the fault. The white
circles are the centers of five subfaults used to compare against theoretical regularized Yoffe source
time functions (28). (B) STFs at the five locations from (A). Plotted are the inverted slip rates and the
regularized Yoffe functions measured from the vertical velocity at KKN4, scaled to the maximum
observed slip rate at each point, which is indicated numerically. Time is relative to the hypocentral
origin (28.147°N, 84.708°E; 25 April 2015, 06:11:26.270 UTC).
RESEARCH |REPORTS
REFERENCES AND NOTES
1. M. Guatteri, P. Spudich, Bull. Seismol. Soc. Am. 90,98116
(2000).
2. X. Lu, N. Lapusta, A. J. Rosakis, Proc. Natl. Acad. Sci. U.S.A.
104, 1893118936 (2007).
3. R. M. Nikolaidis et al., J. Geophys. Res. Solid Earth 106,
2189721916 (2001).
4. G. L. Emore, J. S. Haase, K. Choi, K. A. Larson, A. Yamagiwa,
Bull. Seismol. Soc. Am. 97, 357378 (2007).
5. J.-P. Avouac, L. Meng, S. Wei, W. Wang, J.-P. Ampuero,
Nat. Geosci. (2015).
6. USGS, M7.8 - 36km E of Khudi, Nepal (2015); http://
earthquake.usgs.gov/earthquakes/eventpage/us20002926.
7. T. Ader et al., J. Geophys. Res. Solid Earth 117, B04403
(2012).
8. NSET, Gorkha Earthquake: Intensity Maps (2015); www.nset.
org.np/eq2015/intensity_maps.php.
9. J. L. Mugnier, P. Huyghe, A. P. Gajurel, B. N. Upreti, F. Jouanne,
Tectonophysics 509,3349 (2011).
10. R. Bilham, Curr. Sci. 69, 101128 (1995).
11. M. R. Pant, Ādarśa2,2960 (2002).
12. Kathamandu Valle y Earthquake Risk Management Project,
The Kathmandu Valley Earthquake Risk Management
Action Plan (NSET, Lalitpur, Nepal, 1998); www.
preventionweb.net/english/professional/publications/
v.php?id=1496.
13. Y. R. Paudyal, R. Yatabe, N. P. Bhandary, R. K. Dahal,
Earthq. Eng. Eng. Vib. 11, 257268 (2012).
14. S. Moribayashi, Y. Maruo, J. Jap. Soc. Eng. Geol. 21,8087
(1980).
15. B. S. J. B. Rana, Nepal Ko Maha Bhukampa (1990 Sala) Nepal's
Great Earthquake (Sahyogi Prakashan, Kathmandu, 1935).
16. The Ne pal Geodetic Array (www.tectonics.caltech.edu/
resources/kmlnepal.html) was deployed through a collabo-
ration between the Caltech Tectonics Observatory (USA),
the Department of Mines and Geology (Nepal), and the
Department Analyse et Surveillance de lEnvironnement
(CEA, France).
17. USGS, NetQuakes: Station KATNP_NQ_01, 25 April 2015;
http://earthquake.usgs.gov/monitoring/netquakes/station/
KATNP_NQ_01/20150425061138/.
18. E. Lindsey et al., Nepal Earthquake: Line of Sight Deformation
from ALOS-2 Interferometry, 7 May 2015; http://topex.ucsd.
edu/nepal.
19. European Space Agency, Nepal Earthquake: Sentinel-1 InSAR
Analysis, April 2015; http://insarap.org/.
20. C. Ji, K. M. Larson, Y. Tan, K. W. Hudnut, K. H. Choi, Geophys.
Res. Lett. 31, L17608 (2004).
21. H. Yue et al., J. Geophys. Res. Solid Earth 118, 54535466
(2013).
22. T. H. Heaton, Phys. Earth Planet. Inter. 64,120
(1990).
23. D. Melgar, Y. Bock, J. Geophys. Res. Solid Earth 120,
33243349 (2015).
24. Materials and methods are available as supplementary
materials on Science Online.
25. R. S. Stein, Nature 402, 605609 (1999).
26. R. Bilham, K. Larson, J. Freymueller, Nature 386,6164
(1997).
27. R. Cattin, J.-P. Avouac, J. Geophys. Res. Solid Earth 105,
1338913407 (2000).
28. E. Tinti, E. Fukuyama, A. Piatanesi, M. Cocco, Bull. Seismol.
Soc. Am. 95, 12111223 (2005).
29. S. Ide, M. Takeo, J. Geophys. Res. Solid Earth 102,
2737927391 (1997).
30. T. Mikumo, K. B. Olsen, E. Fukuyama, Y. Yagi, Bull. Seismol.
Soc. Am. 93, 264282 (2003).
31. A. Bizzarri, M. Cocco, J. Geophys. Res. Solid Earth 108, 2373
(2003).
32. J. E. Kozdon, E. M. Dunham, Bull. Seismol. Soc. Am. 103,
12751289 (2013).
33. D. D. Oglesby, R. J. Archuleta, S. B. Nielsen, Science 280,
10551059 (1998).
34. C. B. Worden, M. C. Gerstenberger, D. A. Rhoades, D. J. Wald,
Bull. Seismol. Soc. Am. 102, 204221 (2012).
ACKNOWL EDGME NTS
The GPS data are available from the UNAVCO website. The InSAR
data are available at http://topex.ucsd.edu/nepal/. The Nepal
Geodetic Array was funded by Caltech and DASE (to J.-P.A.) and
by the Gordon and Betty Moore Foundation, through grant GBMF
423.01 to the Caltech Tectonics Observatory; support was
maintained by NSF grant EAR-1345136. A. Miner and the Pacific
Northwest Geodetic Array (PANGA) at CWU are thanked for
technical assistance with the construction and operation of the
Tribhuvan University (TU)CWU network. Additional funding for the
TU-CWU network came from the United Nations Development
Programme and the Nepal Academy for Science and Technology.
The high-rate data were recovered thanks to (i) a rapid
intervention funded by NASA (USA) and the Department of Foreign
International Development (UK) and (ii) engineering services
provided by UNAVCO via the GAGE (Geodesy Advancing
Geosciences and EarthScope) Facility, with support from NSF and
NASA under NSF Cooperative Agreement no. EAR-1261833. We
also thank Trimble Navigation and the Vaidya family for supporting
the rapid response. The accelerometer record at KATNP was
provided by USGS. We thank A. Nathan (U.S. Embassy in
Kathmandu), S. Hough, D. Given, I. Flores, and J. Luetgert for
contributions to the installation of this station. Research at
UCBerkeley was funded by the Gordon and Betty Moore
Foundation through grant GBMF 3024. A portion of this work
was carried out at JPL under a contract with the NASA.
The GPS data were processed by the Advanced Rapid Imaging and
Analysis Center for Natural Hazards (JPL) and the Scripps Orbit
and Permanent Array Center. The effort at the Scripps Institution
of Oceanography was funded by NASA grants NNX14AQ53G and
NNX14AT33G. Advanced Land Observing Satellite2 data were
provided by the Japan Aerospace Exploration Agency under
investigations 1148 and 1413. J.-P.A. thanks the Royal Society for
support. We thank D. Dreger for discussion and W. Mooney for
comments. J.-P.A led the study and wrote the article. D.M.
performed the kinematic modeling and wrote the article. Y.B.
supervised the high-rate data processing and wrote the article.
J.Ga. led the field operations. J.Ge. conducted the high-rate data
processing. S.O., A.M., W.S., and J.F.G. conducted the low-rate
data analysis to estimate coseismic offsets. E.O.L. and X.X.
conducted the InSAR data processing. L.B. helped to organize the
field operations. All other authors contributed to building and
servicing the GPS stations and to the post-earthquake data
recovery. All authors edited the article.
SUPPLEMENTARY MATERIALS
www.sciencemag.org/content/349/6252/1091/suppl/DC1
Materials and Methods
Figs. S1 to S11
Tables S1 and S2
Movie S1
References (3545)
25 May 2015; accepted 29 July 2015
Published online 6 August 2015
10.1126/science.aac6383
SYNT HETIC B IOLOGY
Complete biosynthesis of opioids
in yeast
Stephanie Galanie,
1
Kate Thodey,
2
Isis J. Trenchard,
2
Maria Filsinger Interrante,
2
Christina D. Smolke
2
*
Opioids are the primary drugs used in Western medicine for pain management and
palliative care. Farming of opium poppies remains the sole source of these essential
medicines, despite diverse market demands and uncertainty in crop yields due to weather,
climate change, and pests. We engineered yeast to produce the selected opioid
compounds thebaine and hydrocodone starting from sugar. All work was conducted in a
laboratory that is permitted and secured for work with controlled substances. We
combined enzyme discovery, enzyme engineering, and pathway and strain optimization
to realize full opiate biosynthesis in yeast. The resulting opioid biosynthesis strains
required the expression of 21 (thebaine) and 23 (hydrocodone) enzyme activities from
plants, mammals, bacteria, and yeast itself. This is a proof of principle, and major hurdles
remain before optimization and scale-up could be achieved. Open discussions of options
for governing this technology are also needed in order to responsibly realize alternative
supplies for these medically relevant compounds.
Opioids are an important class of medicines
that include the analgesic morphine and
the antitussive codeine. The World Health
Organization (WHO) classifies these com-
pounds as essential medicines because of
their utility in treating severe pain, in pain man-
agement, and in palliative care (1). In the de-
veloping world, there are shortages of painkillers;
the WHO has estimated that 5.5 billion people
have low to nonexistent access to treatment for
moderate or severe pain(2).
All natural opiates (e.g., morphine and codeine)
and semisynthetic opioids (e.g., oxycodone, hydro-
codone, and hydromorphone) are currently de-
rived from the opium poppy (Papaver somniferum).
Approximately 100,000 ha of opium poppy are
cultivated annually to yield poppy straw contain-
ing more than 800 tons of opiates, primarily mor-
phine and thebaine, to meet licit medical and
scientific demand (3). The majority of poppy-
derived morphine and thebaine is chemically
converted into higher-value compounds, includ-
ing codeine, oxycodone, and hydrocodone. Indus-
trial poppy farming is susceptible to environmental
factors such as pests, disease, and climate, which
can introduce instability and variability into this
geographically concentrated supply chain, result-
ing in pressure to diversify supply (4). Despite
diverse market demands and increasing supply
risks, poppy farming remains the sole source of
opioids, in part because chemical synthesis of
these complex molecules is not commercially
competitive. Approximately 30 chemical syntheses
SCIENCE sciencemag.org 4SEPTEMBER2015VOL 349 ISSUE 6252 1095
1
Department of Chemistry, Stanford University, Stanford, CA
94305, USA.
2
Department of Bioengineering, Stanford
University, Stanford, CA 94305, USA.
*Corresponding author. E-mail: csmolke@stanford.edu
RESEARCH |REPORTS
DOI: 10.1126/science.aac6383
, 1091 (2015);349 Science et al.J. Galetzka
Gorkha earthquake, Nepal
Slip pulse and resonance of the Kathmandu basin during the 2015
This copy is for your personal, non-commercial use only.
clicking here.colleagues, clients, or customers by , you can order high-quality copies for yourIf you wish to distribute this article to others
here.following the guidelines can be obtained byPermission to republish or repurpose articles or portions of articles
): September 20, 2015 www.sciencemag.org (this information is current as of
The following resources related to this article are available online at
http://www.sciencemag.org/content/349/6252/1091.full.html
version of this article at: including high-resolution figures, can be found in the onlineUpdated information and services,
http://www.sciencemag.org/content/suppl/2015/08/05/science.aac6383.DC1.html
can be found at: Supporting Online Material
http://www.sciencemag.org/content/349/6252/1091.full.html#ref-list-1
, 11 of which can be accessed free:cites 36 articlesThis article
http://www.sciencemag.org/cgi/collection/geochem_phys
Geochemistry, Geophysics subject collections:This article appears in the following
registered trademark of AAAS. is aScience2015 by the American Association for the Advancement of Science; all rights reserved. The title CopyrightAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005.
(print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by theScience
on September 20, 2015www.sciencemag.orgDownloaded from
... The planar fault of 66 km × 21 km is further discretized into 152 subfaults, each with a 3 × 3 km 2 area. We adopt the linear multi-time-window (MTW) approach to account for the potential variations in rupture velocity and local slip rates (Hertzell and Heaton, 1983;Galetzka et al., 2015;Liu et al., 2019). In particular, the slip rate used as the source time function (STF) of each subfault is discretized into 10 isosceles triangles, each with a 1 s half-duration and an overlap by 50% duration, allowing the maximum duration of 11 s. ...
... 2c, 4a). On the other hand, a slower rupture velocity is allowed if the early time windows of the subfault STF are solved as with little slip (Galetzka et al., 2015;Yue et al., 2017). This is the case during the southward rupture, in which some subfault STFs with large slips are delayed by 1-2 s (Fig. S7). ...
... Moreover, the ground motions recorded at the MOXI site commonly display higher amplitudes than the synthetic waveforms from all the aforementioned inversions. It is noteworthy that the MOXI site is a confirmed bedrock station with the same construction standards as the reference site of CMONOC , which is rarely engaged with the GNSS monument amplification (Galetzka et al., 2015;Hodgkinson et al., 2020). The fitting to abnormally large amplitude of the MOXI recording in the inversion tends to have a large increase in overall slip and significant amplification for the modeling waveforms of other hrGNSS sites, which is verified in the hrGNSS-only inversions for which the waveforms from the MOXI site are upweighted by a factor ranging from 2 to 8 (Fig. 8). ...
Article
Impulse motion characterized by a large amplitude in the fault-normal direction can be observed at near-fault strong motion sites during strike-slip earthquakes. The large pulse, which always causes high intensity and stronger damage to structures close to faults, is usually attributed to the directivity effect of rupture propagating along strike and the proximity to the fault. We present an analysis of such a large directivity pulse captured by the near-fault high-rate Global Navigation Satellite System (GNSS) during the 2022 Mw 6.7 Luding, China, earthquake—the largest event ever observed by space geodesy on the seismically active Xianshuihe fault in the eastern Tibetan Plateau. We invert the displacement waveforms and offsets derived from the continuous and campaign GNSS for the rupture kinematics. The inferred slip model reveals a rupture zone of 30 km in length above 15 km depth along the Moxi segment, yielding a seismic moment of 1.1×1019 N·m and a source duration of 13 s. The high-rate GNSS (hrGNSS) waveforms suggest an asymmetric bilateral rupture: most slips with long rise time are concentrated on the southern part of the ruptured fault, whereas a short-duration pulse-like slip rate with low final slip propagates during the northward rupture. We found that the directivity pulse observed by the nearest hrGNSS site is controlled primarily by the sharp pulse-like slip rate and rapid rupture velocity approximating the local S-wave velocity. Along with additional local amplification, this large directivity pulse may be responsible for the heavy damage in Moxi town close to the northern ruptured fault.
... The main purpose of the installation of these sites is to identify the extent of the region of strain accumulation along with its rate (Yadav et al. 2019). Such a network also provides valuable information on the earthquake source parameters in case of occurrence of a strong earthquake (Galetzka et al. 2015;Yadav et al. 2017). For this purpose, the geodetic grade multi-frequency GPS receiver and choke ring antenna have been used. ...
Article
Full-text available
Slow slope movement is very common in the Himalayan region. In most cases, the topsoil, weathered rock mass, debris of paleo- landslide or of glacial origin, etc. slowly slides until it stabilizes or fails. In some cases, even the bedrock may be involved in sliding. Here we report four such sliding slopes on which GPS were actu- ally installed for tectonic deformation monitoring in the Himalayan region after a careful investigation but they recorded slow slope sliding. The slow sliding motion on these slopes varied from 10-15 mm/year to 5-6 m/year. The slow slides did not show any anomalous seasonal variation due to hydrological loading. Amongst the four, the Joshimath slide exhibited a large variation in sliding motion which varied from less than a mm/day to 15 mm/day during 2022-2023, causing cracks on the ground and buildings and panic amongst the residents. Even the highly non- linear motion with a sudden change in sliding motion at Joshimath appears to be of viscous creep type. The exact time of initiation of this episode of slide is not known but the available InSAR analyses imply that it is moving at least since 2018. Its motion does not seem to be affected by the February 7, 2021 floods, however, whether the anthropogenic activities around the sliding zone influenced its motion and initiation, remains to be debated.
Article
Full-text available
In Geophysics, topographic factors are observations that can be directly measured, but they are often ignored to simplify the model. Studying the coseismic deformation caused by earthquakes helps accurately determine the epicenter's parameterization. It provides a reference for the reasonable layout of coseismic observation stations and GNSS observation stations. After the Mw7.8 earthquake in Nepal in 2015, GCMT, USGS, GFZ, CPPT, and other institutions released their epicenter parameter. However, according to their parameters, the coseismic displacements simulated by the spectral-element method are quite different from the GNSS observations. Firstly, this paper inverts the geometric parameters of the seismogenic fault with Nepal’s coseismic GNSS displacement. The spectral-element method determines the source's location and depth under the heterogeneous terrain and outputs the source parameters. Among the results of many studies, the surface source is more consistent with the generation mechanism of large earthquakes. Secondly, this paper calculates the fault slip distribution of this earthquake using SDM (Steepest Descent Method) based on GNSS and InSAR data, which is divided into 1500 subfaults, and the moment tensor of each subfault is calculated. This paper investigates the distribution characteristics of the coseismic deformation field of the 2015 Mw 7.8 earthquake in Nepal under three different models. The results show that the influence of topographic factors is ~ 20%, and the influence of heterogeneous factors is ~ 10%. This paper concludes that the influence of topographic factors is much more significant than that of heterogeneous factors, and the influence of both should be addressed in coseismic deformation calculations.
Article
This study attempts to interrogate the upper mantle deformation pattern beneath the Kumaon-Garhwal region, located in the western Himalaya, using shear wave splitting (SWS) analysis of core-refracted (XK(K)S) phases recorded at 53 broadband stations. The fast polarisation azimuths (FPAs) revealed by 338 well constrained measurements are dominantly clustered around ENE-WSW, with a few along the NE and E-W directions. The delay times vary from 0.2 to 1.4 s, with an average of 0.6 s that is smaller than that for the Indian shield (∼0.8 s), central and eastern Himalayas. The northern part of the lesser Himalaya shows a slightly smaller delay time compared to the southern part, which is attributed to the weakening of azimuthal anisotropy caused by the dipping of the Indian lithosphere. In order to understand the crustal contribution, its anisotropy is measured by analysing the splitting of Ps conversions from the Moho (Pms), akin to that of the XK(K)S phases. However, reliable results for crustal anisotropy could be obtained only at 10 stations. The average delay time due to crustal anisotropy is 0.47 s, with a variation from 0.2 to 0.9 s. Although the dominant period of Pms is smaller than that of SK(K)S, crustal anisotropy contributing to splitting of the latter phases cannot be ruled out. The orientation of FPAs obtained from Pms phases is found to be parallel or sub-parallel to those from XK(K)S phases, suggesting a similar deformation mechanism in the mid- to lower-crust and upper mantle. On the basis of FPAs derived from XK(K)S measurements, the Kumaon-Garhwal Himalaya (KGH) region can be divided into four sub-regions. In the western and eastern parts, the FPAs are mostly aligned along NE and ENE-WSW, and NE, respectively. In the central and south-eastern parts, their orientation is along ENE-WSW and NW, respectively. The strong ENE-WSW orientation in the central part could result from a slightly variable anisotropy in the crust to the upper part of the lithosphere or basal topography causing deflection of mantle flow. Also, the NW orientation in the south-eastern part of KGH is associated with a shallow source within the lithosphere. Application of the spatial coherency technique to single-layered anisotropic parameters results in a depth of 220-240 km, implying that the dominant source of anisotropy could lie in the upper mantle.
Article
Full-text available
The destructive 2015 Mw7.8 Gorkha earthquake occurred in the Main Himalayan Thrust due to the collision of the Indian and Asian plates, which provides a unique opportunity to understand the deep dynamic processes and seismogenic mechanisms of strong earthquakes. We construct a regional-scale shear-wave velocity model of the crust and uppermost mantle using ambient seismic noise and teleseismic surface wave at periods of 5–100 s around the Gorkha earthquake region. The new shear-wave velocity model exhibits prominently lateral heterogeneities in the Gorkha earthquake areas. We observe a high-velocity (high-V) zone around the Gorkha main shock in the Main Himalayan Thrust, indicating the existence of a high-strength asperity that sustains the stress accumulating. The aftershocks are primarily located in the low-velocity (low-V) anomalies and enclosed by two high-V anomalies, which appear to act as structural barriers that influence the spread of the aftershocks. Prominent low-Vanomalies from the lower crust to the mantle lithosphere are observed along the north–south trending rifts, suggesting the hot materials upwelling due to the tearing of the northward subducting Indian lithosphere. These observations may indicate that seismic velocity heterogeneity could play an essential role in earthquake initiation and the rupture process.
Preprint
Full-text available
Earthquakes are one of the natural hazards that have caused many casualties and financial losses around the world. This is why earthquake risk analysis studies need to be conducted more seriously. Iran is located in one of the seismogenic regions of the world, the Himalayan-Alpine belt, which is subject to many earthquakes every year. The Arias intensity, as one of the important seismic parameters, helps in seismic hazard analysis, and can be used to estimate structural performance, slope stability, and liquefaction during an earthquake. In this study, the tectonics and seismotectonics of the region of Khoy in northwestern Iran were investigated, and the location of future earthquake was determined. Next, Arias intensity’s maps for return periods of 475, 975 and 2475 years were prepared. The effect of shear wave velocity (Vs 30 ) was also considered during the analyses. The most dangerous area, who’s located at one of future earthquake locations, and also has the potential of resonance and liquefaction, is the northeast of the Salmas fault. Evidently, other than the distance from the fault, the intensity of the earthquake is affected by changes in Vs 30 . It means, in some cases, at more distant points from the seismic source, the intensity even increases. This shows the importance of the effects of geological conditions on the estimation of intensity.
Article
Full-text available
Iron–nickel bimetallic nanoparticles (Fe–Ni BMNPs) are prepared by combining two different metals by using the bottom-up approach. The resulting material has entirely different properties as compared to both the metals. The product is examined by using different analytical instruments such as.; scanning electron microscope (SEM), transmission electron microscope (TEM), X-ray diffraction (XRD), MDIJADE, ORIGIN pro to characterize their morphology, crystallinity and elemental composition and the final data has been statistically analyzed. SEM findings show that most nanoparticles are irregular in form and range in size from 10 nm to 100 nm. The findings of the TEM verified that the particles between 10 nm and 50 nm are irregular in size shape. The products acquired utilized as a fuel additive to monitor oil effectiveness by studying various parameters. The degradation of methylene blue dye depends directly on the concentration of the nanocatalyst. Different parameters also use the freshly prepared bimetallic nanocatalyst to investigate the efficacy of the kerosene fuel. By adding a tiny quantity of the nanocatalyst, the value of the flash point and fire point is significantly reduced. The nanocatalyst does not affect the cloud point and pour point to a large extent. The bimetallic nanocatalyst therefore has very excellent catalytic characteristics.
Article
The Granada Basin in southeast Spain is an area of moderate seismicity. Yet, it hosts some of the highest seismic hazards in the Iberian Peninsula due to the presence of shallow soft sediments amplifying local ground motion. In urban areas, seismic measurements often suffer from sparse instrumentation. An enticing alternative to conventional seismometers is the Distributed Acoustic Sensing (DAS) technology that can convert fiber-optic telecommunication cables into dense arrays of seismic sensors. In this study, we perform a shallow structure analysis using the ambient seismic field interferometry method. We conduct a DAS array field test in the city of Granada on the August 26th and 27th, 2020, using a telecommunication fiber. In addition to the existing limitations of using DAS with unknown fiber-ground coupling conditions, the complex geometry of the fiber and limited data recording duration further challenge the extraction of surface-wave information from the ambient seismic field in such an urban environment. Therefore, we develop a processing scheme that incorporates a frequency-wavenumber (f − k) filter to enhance the quality of the virtual shot gathers and related multi-mode dispersion images. We are able to employ this dataset to generate several shear-wave velocity (VS) profiles for different sections of the cable. The shallow VS structure shows a good agreement with different geological conditions of soil deposits. This study demonstrates that DAS could provide insights into soil characterization and seismic microzonation in urban areas. In addition, the results contribute to a better understanding of local site response to ground motion.
Article
Full-text available
Interseismic strain across the Himalaya is associated with intense microseismic activity. In this study we analyze in detail this seismicity to explore in more details how it relates to Himalayan tectonics. We use the Double Difference Relocation Method to relocate local earthquakes recorded by National Seismological Centre in the period between 1995 and 2003. We also determined fault plane solutions for 10 earthquakes based on waveforms modeling and first P-motion data to complement the existing dataset of focal mechanisms. The results depict a narrow belt of intense seismic activity, at depth between 10 and 20km, which can be traced all along the topographic front of the Higher Himalaya in Nepal. This zone coincides with the zone of interseismic strain build up at the down dip end of the locked portion of the Main Himalayan Thrust fault. These earthquakes probably activate minor faults within the Himalayan wedge, in a volume which experiences stress build up in the interseismic period. The intense seismic activity generally cease abruptly as the elevation of topography gets higher than 3500 m. This elevation also coincides with a change of tectonic regime. Earthquakes within the seismicity belt at front of the Higher Himalaya are of thrust type indicating N-S shortening. Where the topographic elevation is higher than 3500 m focal mechanisms indicate E-W extension. The effect of topography on the regional stress field thus provides a simple explanation for the distribution of focal mechanisms and the seismicity cut-off at the 3500m elevation contour line.
Article
A complete set of closed analytical expressions is presented in a unified manner for the internal displacements and strains due to shear and tensile faults in a half-space for both point and finite rectangular sources. These expressions are particularly compact and systematically composed of terms representing deformations in an infinite medium, a term related to surface deformation and that is multiplied by the depth of observation point. Several practical suggestions to avoid mathematical singularities and computational instabilities are also presented. The expressions derived here represent powerful tools both for the observational and theoretical analyses of static field changes associated with earthquake and volcanic phenomena.
Article
Repeat-pass Interferometric Synthetic Aperture Radar (InSAR) provides spatially dense maps of surface deformation with potentially tens of millions of data points. Here we estimate the actual covariance structure of noise in InSAR data. We compare the results for several independent interferograms with a large ensemble of GPS observations of tropospheric delay and discuss how the common approaches used during processing of InSAR data affects the inferred covariance structure. Motivated by computational concerns associated with numerical modeling of deformation sources, we then combine the data-covariance information with the inherent resolution of an assumed source model to develop an efficient algorithm for spatially variable data resampling (or averaging). We illustrate these technical developments with two earthquake scenarios at different ends of the earthquake magnitude spectrum. For the larger events, our goal is to invert for the coseismic fault slip distribution. For smaller events, we infer the hypocenter location and moment. We compare the results of inversions using several different resampling algorithms, and we assess the importance of using the full noise covariance matrix.
Article
[Full text available at http://rdcu.be/dBc0] Large earthquakes are thought to release strain on previously locked faults. However, the details of how earthquakes are initiated, grow and terminate in relation to pre-seismically locked and creeping patches is unclear. The 2015 M w 7.8 Gorkha, Nepal earthquake occurred close to Kathmandu in a region where the prior pattern of fault locking is well documented. Here we analyse this event using seismological records measured at teleseismic distances and Synthetic Aperture Radar imagery. We show that the earthquake originated northwest of Kathmandu within a cluster of background seismicity that fringes the bottom of the locked portion of the Main Himalayan Thrust fault (MHT). The rupture propagated eastwards for about 140 km, unzipping the lower edge of the locked portion of the fault. High-frequency seismic waves radiated continuously as the slip pulse propagated at about 2.8 km s-1along this zone of presumably high and heterogeneous pre-seismic stress at the seismic-aseismic transition. Eastward unzipping of the fault resumed during the Mw 7.3 aftershock on 12 May. The transfer of stress to neighbouring regions during the Gorkha earthquake should facilitate future rupture of the areas of the MHT adjacent and updip of the Gorkha earthquake rupture.
Article
A complete set of closed analytical expressions is presented in a unified manner for the internal displacements and strains due to shear and tensile faults in a half-space for both point and finite rectangular sources. Several practical suggestions to avoid mathematical singularities and computational instabilities are presented. -from Author
Article
Dynamic simulations of earthquakes on dipping faults show asymmetric near-source ground motion caused by the asymmetric geometry of such faults. The ground motion from a thrust or reverse fault is larger than that of a normal fault by a factor of 2 or more, given identical initial stress magnitudes. The motion of the hanging wall is larger than that of the footwall in both thrust (reverse) and normal earthquakes. The asymmetry between normal and thrust (reverse) faults results from time-dependent normal stress caused by the interaction of the earthquake-generated stress field with Earth's free surface. The asymmetry between hanging wall and footwall results from the asymmetric mass and geometry on the two sides of the fault.
Article
Rapid near-source earthquake source modeling relying only on strong motion data is limited by instrumental offsets and magnitude saturation, adversely affecting subsequent tsunami prediction. Seismogeodetic displacement and velocity waveforms estimated from an optimal combination of high-rate GPS and strong motion data overcome these limitations. Supplementing land-based data with offshore wave measurements by seafloor pressure sensors and GPS-equipped buoys can further improve the image of the earthquake source and prediction of tsunami extent, inundation and runup. We present a kinematic source model obtained from a retrospective real-time analysis of a heterogeneous data set for the 2011 Mw 9.0 Tohoku-oki, Japan earthquake. Our model is consistent with conceptual models of subduction zones, exhibiting depth dependent behavior that is quantified through frequency domain analysis of slip rate functions. The stress drop distribution is found to be significantly more correlated with aftershock locations and mechanism types when off-shore data are included. The kinematic model parameters are then used as initial conditions in a fully non-linear tsunami propagation analysis. Notably, we include the horizontal advection of steeply slopingbathymetric features. Comparison with post-event on-land survey measurements demonstrates that the tsunami's inundation and runup are predicted with considerable accuracy, only limited in scale by the resolution of available topography and bathymetry. We conclude that it is possible to produce credible and rapid, kinematic source models and tsunami predictions within minutes of earthquake onset time for near-source coastal regions most susceptible to loss of life and damage to critical infrastructure, regardless of earthquake magnitude.
Article
Gravity survey was carried out to clarify the basement topography of the Kathmandu Valley which is filled with the Quaternary lacustrine sediments. A gravity anomaly map is produced from 112 gravity measurements within a major part of the Kathmandu Valley and a basement contour map is presented based on depth calculations along two sections.The maximum depth of the basement is estimated to be a little more than 650m from the present surface. Two distinct troughs of basement are detected in the central part of the Valley. The troughs may be a part of fossil valleys of what Hagen (1968) proposed as the Proto-Bagmati River which supposedly had drained south-south-westerly across the Kathmandu Valley during the Plio-Pleistocene time.