ArticlePDF Available

Abstract and Figures

It is commonly thought that the longer the time since last earthquake, the larger the next earthquake's slip will be. But this logical predictor of earthquake size, unsuccessful for large earthquakes on a strike-slip fault, fails also with the giant 1960 Chile earthquake of magnitude 9.5 (ref. 3). Although the time since the preceding earthquake spanned 123 years (refs 4, 5), the estimated slip in 1960, which occurred on a fault between the Nazca and South American tectonic plates, equalled 250-350 years' worth of the plate motion. Thus the average interval between such giant earthquakes on this fault should span several centuries. Here we present evidence that such long intervals were indeed typical of the last two millennia. We use buried soils and sand layers as records of tectonic subsidence and tsunami inundation at an estuary midway along the 1960 rupture. In these records, the 1960 earthquake ended a recurrence interval that had begun almost four centuries before, with an earthquake documented by Spanish conquistadors in 1575. Two later earthquakes, in 1737 and 1837, produced little if any subsidence or tsunami at the estuary and they therefore probably left the fault partly loaded with accumulated plate motion that the 1960 earthquake then expended.
Content may be subject to copyright.
NATURE|Vol 437|15 September 2005 NEWS & VIEWS
329
the associated tsunamis and seen in layers of
intercalated sand and soil extending horizon-
tally for nearly a kilometre close to the mouth
of the Río Maullín. This region subsided
approximately 2 m as a consequence of the
1960 event, but the stratigraphy has been
preserved because of long-term net tectonic
uplift. Overall, the authors were able to identify
a sequence of eight large earthquakes that
have occurred over the past 2,000 years,
with an average recurrence interval of about
300 years.
But what about the apparent recurrence
sequence of 1575, 1737, 1837 and 1960, which
does not fit this pattern? The consequences of
the 1575 event are evident both from carbon
dating and in documents written by the Span-
ish conquistadors, which tell of effects that
were similar to those seen in 1960. Documen-
tation of the earthquakes of 1737 and 1837 is
less clear, but they are frequently cited in the
seismological histories of the region as being
large (estimated magnitudes of 7.5 and 8.0,
respectively)
8
.
Cisternas et al., however, find that although
the 1575 earthquake appears clearly in their
tsunami stratigraphic record, those of 1737
and 1837 do not. A tsunami was associated
with the event of 1837, reaching Hawaii with
an amplitude of 6 m. But Cisternas et al. sug-
gest that it originated outside their study area,
to the south, and they conclude that the earth-
quake of 1737 was too small to generate a
sizeable tsunami. All in all, they propose that
much of the fault dislocation that occurred in
1960 stemmed from a release of energy that
had remained ‘locked in’ since 1575 — that is,
the earthquakes that occurred between those
times had expended comparatively little of the
accumulated stress.
The relevance of these studies is underlined
by the more recent occurrence of a huge earth-
quake elsewhere in the world. In December
2004, and again in March 2005, the eastern
Indian Ocean was seriously affected by tec-
tonic movement at a subduction zone. The
giant December event stemmed from a dis-
placement of up to 30 m along a rupture 1,100
km in length
9
, which produced the devastating
tsunami that swept across the Indian Ocean —
a startling reminder of the need to learn more
about the behaviour of giant earthquakes.
Sergio Barrientos is at the Preparatory
Commission for the Comprehensive
Nuclear-Test-Ban Treaty Organization,
PO Box 1200, A-1400 Vienna, Austria.
e-mail: sergio.barrientos@ctbto.org
1. Cisternas, M. et al. Nature 437, 404–407 (2005).
2. Kanamori, H. J. Geophys. Res. 82, 2981–2987 (1977).
3. Aki, K. Bull. Earthq. Res. Inst. Tokyo Univ. 44, 23–88 (1966).
4. Benioff, H., Press, F. & Smith, S. J. Geophys. Res. 66, 605–619
(1961).
5. Plafker, G. & Savage, J. C. Geol. Soc. Am. Bull. 81, 1001–1030
(1970).
6. Barrientos, S. & Ward, S. N. Geophys. J. Int. 103, 589–598
(1990).
7. Klotz, J. et al. Earth Planet. Sci. Lett. 193, 437–446 (2001).
8. Lomnitz, C. Seism. Res. Lett. 75, 368–378 (2004).
9. Vigny, C. et al. Nature 436, 201–206 (2005).
EARTHQUAKES
Giant returns in time
Sergio Barrientos
The behaviour of a seismic fault in Chile seemed to confound predictions
of how often giant earthquakes should recur. Examination of a 2,000-year
record of tsunami deposits in the region clarifies matters.
In May 1960, south-central Chile experienced
a huge earthquake, the largest since instru-
ment records began. The consequences were
felt not only in Chile, but also in Hawaii, the
Philippines, Japan and other locations across
the Pacific, which were all hit by the ensuing
tsunami. But this giant seismic event has long
puzzled seismologists, because the energy
released by the earthquake should have taken
several centuries to build up, through the
accumulation of stress on the fault concerned,
as the Nazca and South American tectonic
plates converge at a rate of 8–9 centimetres per
year. Yet the immediately preceding events are
historically documented as occurring not only
in 1575, which would fit expectations, but also
in 1737 and 1837.
As they describe on page 404 of this issue
1
,
Cisternas et al. have revisited this puzzle by
examining records of land movement and
tsunamis left in buried layers of sand and soil in
an estuary that runs across the central part of
the fault zone (see Fig. 1 of the paper, page 404).
With these records, and earlier ones (Fig. 1), the
history of activity on this fault can be rewritten.
Since the inception of instrumental record-
ing in the late 1800s, the largest earthquakes
have all occurred in subduction zones
2
, where
one tectonic plate is being driven beneath
another. These earthquakes involve ruptures
on the order of a thousand kilometres long by
a couple of hundred kilometres wide, with
fault displacements of tens of metres. Hun-
dreds of years are necessary to accumulate the
stresses released in these giant events, which
leave traces of their consequences that can
later be identified.
Thanks to the work of Cisternas et al.
1
, we
now have more insight into the seismic cycle in
south-central Chile. The earthquake of May
1960 resulted from a rupture, about 1,000 km
long and 150 km wide, along the north–south-
trending fault where the Nazca plate dives
beneath South America. Earthquakes of this
size saturate the recorders of the seismic waves
that are usually used to estimate earthquake
magnitude and which are reported in terms of
the familiar Richter scale. Instead, a measure
known as seismic moment is used
3
. This is now
more commonly applied to seismic events in
general, and is calculated by multiplying the
area of the rupture zone by the fault displace-
ment and a quantity reflecting the rigidity of
the volume in which the rupture takes place.
The moment magnitude, derived from this
quantity, reflects the real size in terms of the
elastic energy release of an earthquake.
The 1960 earthquake measured 9.5 on
the moment-magnitude scale. It has a special
place in seismological history, because it pro-
vided experimental confirmation of the idea
that earthquakes can cause free oscillations of
the Earth
4
— that is, set the Earth ringing. The
observed changes in land elevation ranged
from 6 m of uplift to 2 m of subsidence; these
displacements have been modelled
5,6
as the
elastic response of the Earth to an average dis-
location of 20 m along the fault, with localized
peaks of more than 30 m. Even now, in 2005,
post-seismic readjustments continue to be
observed in the area
7
.
Cisternas et al.
1
undertook a detailed exam-
ination of a local stratigraphic record of earth-
quakes in south-central Chile, as produced by
Figure 1 | Spot the
surveyor. In this
print, produced in
1874, the figure of
Francisco Vidal
Gormaz can be
seen at work in the
coastal village of
Carelmapu in
southern Chile.
Here he surveys part
of the region that
will be hit by the
1960 earthquake; his
records have helped
in interpreting
the earthquake.
15.9 News & Views 325 NEW 12/9/05 2:40 PM Page 329
Nature
Publishing
Group
© 2005
Predecessors of the giant 1960 Chile earthquake
Marco Cisternas
1
, Brian F. Atwater
2
, Fernando Torrejo
´
n
3
, Yuki Sawai
5
, Gonzalo Machuca
4
, Marcelo Lagos
6
,
Annaliese Eipert
7
, Cristia
´
n Youlton
1
, Ignacio Salgado
1
, Takanobu Kamataki
5
, Masanobu Shishikura
5
,
C. P. Rajendran
8
, Javed K. Malik
9
, Yan Rizal
10
& Muhammad Husni
11
It is commonly thought that the longer the time since last earth-
quake, the larger the next earthquake’s slip will be. But this logical
predictor of earthquake size
1
, unsuccessful for large earthquakes
on a strike-slip fault
2
, fails also with the giant 1960 Chile earth-
quake of magnitude 9.5 (ref. 3). Although the time since the
preceding earthquake spanned 123 years (refs 4, 5), the estimated
slip in 1960, which occurred on a fault between the Nazca and
South American tectonic plates, equalled 250–350 years’ worth of
the plate motion
3,6–10
. Thus the average interval between such
giant earthquakes on this fault should span several centuries
3,9,10
.
Here we present evidence that such long intervals were indeed
typical of the last two millennia. We use buried soils and sand
layers as records of tectonic subsidence and tsunami inundation at
an estuary midway along the 1960 rupture. In these records, the
1960 earthquake ended a recurrence interval that had begun
almost four centuries before, with an earthquake documented
by Spanish conquistadors in 1575. Two later earthquakes, in 1737
and 1837, produced little if any subsidence or tsunami at the
estuary and they therefore probably left the fault partly loaded
with accumulated plate motion that the 1960 earthquake then
expended.
The 1960 Chile mainshock resulted from a rupture nearly
1,000 km long on a north–south trending fault that conveys the
subducting Nazca plate beneath South America at rates averaging 8 m
per century
3
. Lurching westward above the rupture, the South
America plate rose in a mostly offshore area while subsiding 1–2 m
in a coastal downwarp
6
(Fig. 1b). The ensuing tsunami, with crests
10–15 m high in Chile
11
, reached maximum heights of 10 m in
Hawaii
12
and 6 m in Japan
13
.
The 1960 earthquake was preceded historically by earthquakes in
1575, 1737 and 1837 (Fig. 1b; Supplementary Table S1). The reported
effects from 1575 most nearly resemble those from 1960 (ref. 4).
Conquistadors, at forts limited to the northern half of the 1960
rupture area, wrote of persistent marine inundation near Imperial,
Valdivia and Castro that implies widespread tectonic subsidence.
They also described a devastating tsunami near Valdivia (Sup-
plementary Table S1, record 1). The 1737 earthquake, known only
from secondary sources, damaged the few Spanish settlements then
remaining south of Concepcio
´
n. It lacks a reported tsunami, even
though tsunamis from central Chile in 1730 and 1751 were noted
locally
14
and in Japan
13,15
. The 1837 earthquake damaged towns along
the central third of the 1960 rupture area and changed land levels
along the southern half of that area. Its associated tsunami, by
reportedly cresting 6 m high in Hawaii
12
, provides evidence that
the 1837 earthquake released almost half the seismic moment of the
LETTERS
Figure 1 | Index maps. a, Plate-tectonic setting of south-central Chile.
Paired arrows indicate plate convergence at 8.4 cm yr
21
. b, Documented
effects of the 1960 earthquake and its historical predecessors. Compiled
from refs 4, 5, 13 and 14, and from Supplementary Table S1. c, Study area
along the
´
o Maullı
´
n. Barbed lines in a and b show seaward edges of
subduction zones; teeth point down the plate boundary.
1
Facultad de Agronomı
´
a, Pontificia Universidad Cato
´
lica de Valparaı
´
so, Casilla 4-D, Quillota, Chile.
2
US Geological Survey at University of Washington, Seattle, Washington
98195-1310 USA.
3
Centro EULA-Chile,
4
Departamento de Ciencias Histo
´
ricas y Sociales, Universidad de Concepcio
´
n, Casilla 160-C, Concepcio
´
n, Chile.
5
Active Fault Research
Center, Geological Survey of Japan, National Institute of Advanced Industrial Science and Technology, Tsukuba 305-8567 Japan.
6
Instituto de Geografı
´
a, Pontificia Universidad
Cato
´
lica de Chile, Casilla 306, Santiago, Chile.
7
National Oceanic and Atmospheric Administration, Seattle, Washington 98115-6349, USA.
8
Centre for Earth Science Studies,
Thiruvananthapuram 69603, India.
9
Civil Engineering, Indian Institute of Technology, Kanpur 208016, India.
10
Geological Department, Institute of Technology, Bandung 40132,
Indonesia.
11
Potential Geophysics Division, Meteorological and Geophysical Agency of Indonesia, Jakarta 10720, Indonesia.
Vol 437|15 September 2005|doi:10.1038/nature03943
404
© 2005 Nature Publishing Group
1960 mainshock
16
. However, according to primary sources in Sup-
plementary Table S1, the same tsunami caused little if any flooding at
Valdivia and no reported damage anywhere in Chile or Ancud.
To further compare the 1960 earthquake with these historical
earthquakes, and to gain perspective from earlier earthquakes as well,
we reconstructed a 2,000-yr history of repeated subsidence and
tsunamis at the
´
o Maullı
´
n estuary (Fig. 1b, c). Because of the
estuary’s central location, this history probably includes earthquakes
from full-length breaks of the 1960 rupture area, while perhaps
excluding earthquakes from partial ruptures to the north or south.
Our stratigraphic records are tied to modern analogues from 1960
along a nearly marine reach of the
´
o Maullı
´
n. There, 8 km inland
from the sea (Fig. 1c, purple dot), markers of the 1960 earthquake
extend across faint terraces and beach ridges stranded by net late
Holocene emergence
17
. Eyewitnesses recall that the 1960 tsunami
coated upper terraces with sand
18
. We traced the sand, up to 15 cm
thick, more than 1 km inland across the buried 1960 soil in areas
covered only by the highest post-earthquake tides (Figs 2a, b). In this
same area, the sandy record of post-1960 storms extends just a few
metres inland from the shore. On lower terraces, now covered
routinely by tides, a 1960 pasture soil has been eroded and biotur-
bated on post-earthquake tidal flats. Waves and currents are now
Figure 2 | Stratigraphic evidence for 1960 earthquake and its ancestors in
area shown by purple dot in Fig. 1c.
Supporting data in Supplementar y
Figs S1–S4 and Supplementary Tables S2–S4. a, Records of the 1960
earthquake that serve as modern analogues for inferring past occurrence of a
tsunami and of coseismic subsidence. b, Sequences of such records
correlated among trenches. Tides measured 1989, 2003, and 2004.
c, Chronology of the inferred events compared with the historical sequence
in Fig. 1b. Field ev idence for subsidence and tsunami (solid blue circles and
triangles at left) comes from all transects (Supplementary Figs S2 and S3).
The average of the historical recurrence intervals, 128 yr, contrasts with the
longer average intervals between the events recorded stratigraphically.
NATURE|Vol 437|15 September 2005 LETTERS
405
© 2005 Nature Publishing Group
burying the remains of this soil with sand as much as 1 m thick
(Fig. 2a), and with mud in sheltered areas.
Additional sand sheets mantle buried marsh and meadow soils
beneath the 1960 soil. Using criteria from the 1960 examples, we
interpret some of these sand sheets as tsunami deposits (blue dots in
Fig. 2) and others as indicators of subsided, post-earthquake tidal
flats (blue triangles). We traced these event records, which probably
represent eight earthquakes in all (events A–H), among 60 trenches
scattered along 2 km of transects (example, Fig. 2b). Like the 1960
earthquake (event A), four earlier events (B–D, G) produced tsunami
deposits on meadows that post-earthquake tides rarely reached and
correlative tidal-flat deposits on lower ground. Such evidence,
assembled from all transects, is summarized by solid blue symbols
in Fig. 2c. Some events are recorded less widely than the 1960
earthquake. The D sand sheet tapers landward without crossing a
former beach ridge. The E sand sheet, found entirely inland from that
ridge, may have been removed by erosion on the seaward side.
Diatom assemblages from soils that shortly predate and postdate
tsunami deposition provide further evidence for subsidence during
events A, B and D. In all three cases the assemblages above the
tsunami sand are more nearly marine than those in the soil below
(summary, Fig. 2c). The difference is clearest for the 1960 event. An
attempted comparison for event C failed because the upper part of
the buried soil is probably missing from erosion on a post-C tidal flat,
and because the remnant soil is contaminated with burrow-filling
tidal-flat sand.
In sum, our stratigraphy and paleoecology provide evidence for
seven inferred pre-1960 earthquakes from the past 2,000 years
(Fig. 2c). The youngest three (B–D), each marked by evidence for
both subsidence and tsunami, occurred within the past 1,000 years.
Event D dates to the two-sigma range
AD 1020–1180
the age of
growth-position stem bases of a rush (Juncus procerus) that tsunami
sand surrounded. The event C tsunami similarly left sand around
Juncus balticus and Scirpus americanus culms in a swale along a spur
transect (Supplementary Fig. S3b); below-ground stems (rhizomes)
that probably belonged to such plants yielded three statistically
indistinguishable ages pooled as
AD 1280–1390.
The tsunami deposit from event B probably exceeds the one from
1960 in thickness and landward extent. Because the 1837 tsunami was
large in Hawaii
12,16
, we expected this penultimate sand sheet to date
from the early nineteenth century. Instead, a burned horizon mostly
2 cm below the sand dates to
AD 1450–1510 or 1590–1620, as judged
from four statistically equivalent ages on charred twigs. Because it
followed the fire, probably by a century at most, we correlate event B
with the extensive subsidence and devastating tsunami of 1575
(Fig. 1b).
We checked additional estuarine records in a further, futile search
for signs of the 1837 earthquake. These records include trees that the
1960 earthquake lowered into tidal freshwater farther up the
´
o
Maullı
´
n (red triangle, Fig. 1c). Residents on hand for the 1960
earthquake testify that a forest, green and emergent before the
earthquake, lost its foliage from routine tidal submergence in the
first few years thereafter. Several decades later, defoliated trunks
dominated an area of 10 km
2
. But several decades after the 1837
earthquake, a nautical chart
19
depicted all trees in this area as leafy
(Fig. 3a). In an accompanying report
20
, the expedition botanist does
not mention dead or dying trees among the forest’s riparian plants
and animals, which he studied for four days. We cut slabs of 15 dead
standing trees in 2003 to estimate their lifespans by counting annual
rings. We assume these trees died in 1960. In that case, ten of them
were alive in 1837 and two in 1737 (Fig. 3b). This finding suggests
that the forest failed to subside in 1837 as much as it did in 1960, in
agreement with the nautical survey and the botanist’s report.
Shoreline changes provide additional evidence that the 1837
earthquake did not produce 1960-size subsidence along the
´
o
Maullı
´
n. Some of the islands and pastures that subsided in 1960
into the middle or lower part of the intertidal zone are barren
intertidal or subtidal flats (Fig. 1c, green triangles). At a similar
time after the 1837 earthquake, these areas were charted
19
as
emergent and vegetated (Supplementary Fig. S1b).
Earthquakes evident in these various estuarine records thus
recurred less often than did earthquakes in the historical sequence:
1575, 1737, 1837, 1960 (Fig. 2c). The best-defined of the earthquake
intervals recorded geologically, which together span most of the past
millenium, average nearly 300 yr
more than double the historical
average of 128 years. The 1960 earthquake ended a 385-year interval
that includes the years 1737 and 1837. The poorly understood
earthquakes of 1737 and 1837 probably released too little seismic
moment midway along the 1960 rupture to leave tsunami deposits or
subsidence stratigraphy at the
´
o Maullı
´
n.
Where size varies markedly among successive earthquakes on the
same part of a fault, much of the fault slip during the largest
earthquakes may have thus accumulated before earlier earthquakes
of smaller size. Such storage through multiple recurrence intervals
probably helps to explain the enormity of the 2004 Sumatra–Anda-
man earthquake. The fault slip in 2004 near the Nicobar Islands
amounted to 10 m (ref. 21) in an area where the fault had last
ruptured in 1881 during an earthquake of estimated magnitude 7.9
(ref. 22). By contrast, the fault loading between 1881 and 2004
amounted to less than 4 m at plate-convergence rates recently
estimated from satellite geodesy
22
and less than 7 m at rates inferred
from long-term plate motions
3
. As in the 1960 Chilean case, the 2004
earthquake may thus have used accumulated plate motion that a
previous earthquake left unspent.
Received 9 March; accepted 15 June 2005.
1. Shimazaki, K. & Nakata, T. Time-predictable recurrence model for large
earthquakes. Geophys. Res. Lett. 7, 279–-282 (1980).
2. Weldon, R., Scharer, K., Fumal, T. & Biasi, G. Wrightwood and the earthquake
cycle: what a long recurrence record tells us about how faults work. GSA Today
14(9), 4–-10 (2004).
3. DeMets, C., Gordon, R. G., Argus, D. F. & Stein, S. Current plate motions.
Geophys. J. Int. 101, 425–-478 (1990).
4. Lomnitz, C. Major earthquakes and tsunamis in Chile during the period 1535 to
1955. Geol. Rundsch. 59, 938–-960 (1970).
5. Urrutia, R. & Lanza, C. Cata
´
strofes en Chile 1541–-1992 90 (Editorial La Noria,
Santiago, Chile, 1993).
Figure 3 | Arboreal evidence for difference between the 1960 and 1837
earthquakes, in area shown by red triangle in Fig. 1c.
(See Supplementary
Fig. S1d, e). a, Views of riparian forest several decades after each earthquake
(1874 image from ref. 19). b, Counts of annual r ings in trees probably killed
in 1960 (species, Supplementary Table S4).
LETTERS NATURE|Vol 437|15 September 2005
406
© 2005 Nature Publishing Group
6. Plafker, G. & Savage, J. C. Mechanism of the Chilean earthquakes of May 21
and 22, 1960. Geol. Soc. Am. Bull. 81, 1001–-1030 (1970).
7. Plafker, G. Alaskan earthquake of 1964 and Chilean earthquake of 1960:
implications for arc tectonics. J. Geophys. Res. 77, 901–-925 (1972).
8. Kanamori, H. & Cipar, J. J. Focal process of the great Chilean earthquake May
22, 1960. Phys. Earth Planet. Int. 9, 128–-136 (1974).
9. Cifuentes, I. The 1960 Chilean earthquakes. J. Geophys. Res. 94, 665–-680
(1989).
10. Barrientos, S. E. & Ward, S. N. The 1960 Chile earthquake: inversion for
slip distribution from surface deformation. Geophys. J. Int. 103, 589–-598
(1990).
11. El Maremoto del 22 de Mayo de 1960 en las Costas de Chile (Servicio
Hidrogra
´
fico y Oceanogra
´
fico de la Armada de Chile, Valparaı
´
so, Chile,
1961).
12. Lander, J.F. & Lockridge, P.A. United States tsunamis 1690–-1988. (National
Geophysical Data Center Publ. 41–-2, Boulder, Colorado, 1989).
13. Watanabe, H. Comprehensive List of Destructive Tsunamis to Hit the Japanese
Islands [in Japanese] (Univ. Tokyo Press, Tokyo, 1998).
14. Lockridge, P. A. Tsunamis in Peru-Chile (Report SE-39, World Data Center A for
Solid Earth Geophysics, Boulder, Colorado, 1985).
15. Ninomiya, S. Tsunami in Tohoku coast induced by earthquake in Chile; a
chronological review. Tohoku Kenkyu [in Japanese with English summary] 10,
19–-23 (1960).
16. Abe, K. Size of great earthquak es of 1837–-1974 inferred from tsunami data.
J. Geophys. Res. 84, 1561–-1568 (1979).
17. Atwater, B. F., Jime
´
nez, N. H. & Vita-Finzi, C. Net late Holocene emergence
despite earthquake-induced submergence, south-central Chile. Quat. Int. 15/16,
77–-85 (1992).
18. Atwater, B. F. et al. Surviving a tsunami
lessons from Chile, Hawaii, and Japan.
US Geol. Surv. Circ., 1187 (1999).
19. Vidal Gormaz, F. Esploraciones hidrogra
´
ficas practicadas en las costas de
Chile, por la marina militar de la Repu
´
blica (nautical chart). Anuario
Hidrogra
´
fico de la Marina de Chile, An
˜
o1(Imprenta Nacional, Santiago, Chile,
1875).
20. Juliet, C. Informe del ayudante de la comisio
´
n esploradora de Chiloe
´
i
Llanquihue. Anuario Hidrogra
´
fico de la Marina de Chile, An
˜
o1263–-339
(Imprenta Nacional, Santiago, Chile, 1875).
21. Lay, T. et al. The great Sumatra-Andaman earthquake of 26 December 2004.
Science 308, 1127–-1133 (2005).
22. Ortiz, M. & Bilham, R. Source area rupture parameters of the 31 December
1881 Mw ¼ 7.9 Car Nicobar earthquake estimated from tsunamis recorded in
Bay of Bengal. J. Geophys. Res. 108( B4), 2215, doi:10.1029/2002JB001941
(2003).
Supplementary Information is linked to the online version of the paper at
www.nature.com/nature.
Acknowledgements This work was supported by Chile’s Fondo Nacional de
Desarrollo Cientı
´
fico y Tecnolo
´
gico (Fondecyt) and by the US Geological Survey.
Logistical help came from the Municipality of Maullı
´
n and its people (M. I. Silva,
P. Soto, R. Vergara, J. Gallardo, G. Andrade, J. Soarzo and C. Ruiz), and from the
Servicio Hidrolo
´
gico y Oceanogra
´
fico of the Armada de Chile. The manuscript
incorporates suggestions from S. Barrientos, S. Bondevik, C. Lomnitz, A. Nelson,
K. Wang, J. Clague and E. Geist.
Author Contributions M.C. and B.A. led the fieldwork and writing. F.T. studied
documents; Y.S. studied diatoms; G.M. studied tree slabs. M.L. and I.S.
contributed to three seasons of fieldwork, G.M. and C.Y. to two, and A.E., M.H.,
T.K., J.K.M., C.P.R., Y.R. and M.S. to one.
Author Information Reprints and permissions information is available at
npg.nature.com/reprintsandpermissions. The authors declare no competing
financial interests. Correspondence and requests for materials should be
addressed to M.C. (marco.cisternas@ucv.cl).
NATURE|Vol 437|15 September 2005 LETTERS
407
© 2005 Nature Publishing Group
... A similar story had occurred almost four centuries before, with more tragic consequences. The historical earthquake of 1575, recognized as the predecessor of the 1960 earthquake (Cisternas et al. 2005;Moernaut et al. 2014;Ruiz and Madariaga 2018;Matos-Llavona et al. 2022), also triggered landslides that dammed the SPR (see Sect. 3). However, on that occasion, after months of water accumulation and over ~ 50 m of water-level increase, the dam collapsed and killed hundreds of inhabitants living downstream (Montessus de Ballore 1912). ...
... Orange lines indicate the along-dip location of moderate-size, deeper interplate earthquakes (1737 (Cisternas et al. 2017a), May 21, 1960(Ruiz and Madariaga 2018, and 2016 (Moreno et al. 2018)). Red lines indicate the along-dip location of great earthquakes rupturing the entire seismogenic width (1575( , 1960( (Cisternas et al. 2005Wils et al. 2020) and 2010 (Moreno et al. 2012)). The green line indicates the approximate location of crustal earthquakes, such as the strike-slip rupture of the 2007 earthquake along the Liquiñe Ofqui Fault Zone (LOFZ) (Agurto et al. 2012 Content courtesy of Springer Nature, terms of use apply. ...
... According to the historical record, the 1,000 km long southern segment of this subduction zone has generated at least four major megathrust earthquakes (M8+) in the past 500 years (e.g., Lomnitz 2004). The sequence includes the 1960 event (M9.2-9.6), the largest instrumentally recorded worldwide (Kanamori and Cipar 1974), and other great earthquakes in 1575, 1737, 1837 of unknown magnitudes but most likely greater than 8 (Lomnitz 1970;Cisternas et al. 2005;Moernaut et al. 2014;Ruiz and Madariaga 2018;Matos-Llavona et al. 2022). Historical documents (Cisternas et al. 2017a) and sedimentary records at the coast (Cisternas et al. 2017b) and within inland lakes (Moernaut et al. 2014;Wils et al. 2020) indicate that the 1575 earthquake likely rivaled the 1960 earthquake in terms of magnitude and rupture extent. ...
Article
Full-text available
Damming rivers by landslides and ensuing outburst flooding is a common and potentially hazardous phenomenon worldwide, especially in tectonically active regions. Remarkable examples are the damming of the upper course of the San Pedro River (SPR) in south Chile during the 1960 Chile earthquake (M9.5) and its predecessor in 1575. Outburst floods following both events had tragic consequences for downstream communities. Here, we study both events from multiple sources of information, including previously published and newly found historical records, satellite imagery, LiDAR topography, and sedimentological and geomorphological field observations. We present the first detailed geomorphic map of the region. Morphological similarities between ancient deposits at the SPR and those associated with the 1960 earthquake suggest that the SPR has been dammed repeatedly in the past. The steep incision of the SPR and the sediments of glacio-lacustrine origin in the surrounding slopes facilitate the initiation of large landslides. The knowledge gained from studying these past events provides important implications for future risk assessments. We propose that besides large earthquakes, smaller and more frequent earthquakes as well as changes in land use, can also result in river-damming events.
... The dissipation of wave energy by mangrove forests contributes to the limitation of deposit areas (Sánchez-Núñez et al., 2019) and has been observed in relation to coastal erosion processes (Thampanya et al., 2006). In contrast, marshes and coastal lagoons in the OZ have been more successful in preserving tsunamis, as seen in studies by Atwater and Moore (1992) and Cisternas et al. (2005), where mangrove ecosystems are absent. Rhodes et al. (2011), studying the 2004 Indian Ocean tsunami in Thailand, observed a significant reduction in the sand layer within a mangrove environment, attributed to the presence of the forest, tidal currents, and bioturbation. ...
Article
Full-text available
Tsunami deposits provide evidence of historical and prehistorical events. However, their preservation in tropical regions is generally poor. The reasons behind this poor preservation are often linked to a number of environmental and anthropogenic factors. This study is focused on analyzing the environmental factors that impact the preservation and availability of tsunami deposits specifically in tropical regions. These factors predominantly encompass climate-related elements such as consistently high temperatures, rainfall, humidity, as well as specific soil processes, oceanic conditions, and vegetation. We compiled a comprehensive database of scientific publications on tsunami deposits, identifying the geomorphic environments where such deposits are typically preserved , as well as the commonly utilized proxies in studying tsunami deposits across different climatic zones. We propose a model that outlines the environmental factors, processes, and their interrelationships that contribute to the preservation and availability of tsunami deposits in tropical regions. This model may prove valuable in the future identification of tsunami deposits in tropical areas.
... It was observed and forced evacuation of coastal communities all around the Pacific. The 2010 Maule earthquake is one of a long history of similar events occurring offshore of Chile either in the same area in central-southern Chile (1928 Mw 8.0 event: Beck et al., 1998), or more north (1906Mw 8.4 event: Lomnitz, 19701985Mw 7.8 event: Nakamura, 1992, or more south with the famous 1960 Mw 9.5 event (Cisternas et al., 2005). All these events generated tsunamis observed throughout the Pacific Ocean with significant source. ...
Article
Full-text available
Tracking changes in sea-surface height with ship-based GNSS can be used to detect tsunamis. One year of navigation data from ships in the Pacific is examined to investigate how well-distributed a cargo-ship network would be for tsunami detection. There is excellent coverage of the most active tsunamigenic zones, with multiple ships predicted within 30-minutes travel time of notable tsunamis. Tsunamigenic regions with low ship density, such as the Southwest Pacific, require a greater percentage of ships participating to ensure sufficient data. The global nature of GNSS and ship routes make this a promising, low-cost approach, to augment tsunami detection.
... When historical events are rare, geological records are often deemed as the only approach to tracing the past (e.g., Costa and Andrade, 2020;Nanayama et al., 2003). For this purpose, extensive research on paleo-tsunami deposits has been conducted in many tsunami-prone regions, including the subduction zones of the Sumatra-Andaman (Jankaew et al., 2008;Maselli et al., 2020;Meltzner et al., 2010), Cascadia in North America Kelsey et al., 2005;Satake et al., 2003), Kuril Trench (Nanayama et al., 2003(Nanayama et al., , 2007Satake et al., 2005) and Chilean Trench in south America (Cisternas et al., 2005), etc. ...
... Chile, located within the Pacific Ring of Fire, finds itself in constant exposure to earthquakes across its territory [3,4]. These occurrences bring not only damage to homes and infrastructure but also high economic and social costs linked with reconstruction efforts [5][6][7]). ...
Chapter
Full-text available
In 2010, a severe earthquake occurred in Chile that brought high-impact destruction effects in the affected areas. It has studied the damage to homes, the determiners of the cost of reconstruction, and how households finance reconstruction using the Casen post-earthquake 2010 panel survey. A probit and an OLS model were used. The results show that houses closest to the epicentre were the most affected, where damage increased when roofs and walls were in worse condition or with more vulnerable materials. The reconstruction costs are related to the degree of destruction, the distance to the epicentre, the condition of the walls before the event and the house’s value. Provinces with more bank branches are associated with a lower cost. Bank credit is more likely to be used to rebuild in urban areas, when the head of the household has more years of education and when the repair cost is higher. Own savings will be used when there is no insurance, the higher the income of the head of household and the lower repair costs. Finally, subsidies will be an option when there is no insurance, the repair cost is higher, and for lower income, age, and education.
... Background seismicity (Figure 2o) could be considered for this scenario, since Madella and Ehlers (2021) show a possible correlation with long-term uplift rates in northern Japan and central South America. However, high-frequency changes in the inferred signal of background seismicity and the relatively short time interval of ∼50 years analyzed in comparison with the duration of a seismic cycle (∼100-300 years; Cisternas et al., 2005;Comte & Pardo, 1991) prevent further interpretations regarding the role of this mechanism. ...
Article
Full-text available
The tectonically active South American margin is characterized by the accumulation of deformation contributing to uplift of the Andean forearc at millennial time scales. However, the mechanisms responsible for permanent coastal uplift are debated, mainly because methodologically consistent, continental‐scale analyses of uplifted terraces have not yet been carried out for South America. Uplifted marine terraces are generally used to infer permanent coastal deformation and uplift; we used almost 2,000 measurements of last‐interglacial marine terraces to calculate an uplift‐rate signal on which we performed a wavelength analysis. The same spectral analysis was applied to tectonic and subduction parameters associated with accumulation of permanent deformation to detect possible links with the uplift‐rate signal. The uplift‐rate signal displays a constant background‐uplift rate along the margin, perturbed by changes at variable wavelengths. Similarities between its wavelength spectrum and the spectra of tectonic parameters suggest potential correlations pointing toward underlying processes. For example, crustal faulting is mainly responsible for short‐wavelength deformation; intermediate‐wavelength to long‐wavelength tectonic features indicate various extents of locked areas on the megathrust that relate to its long‐term seismotectonic segmentation. We suggest that moderate, long‐term background uplift is caused by major, deep earthquakes near the Moho, although records of such events are sparse. Due to their disparate occurrence, we infer accumulation of permanent deformation over millennial time scales through multiple, distinct uplift phases that are spatially and temporally distributed. Our study highlights the application and utility of a signal‐analysis approach to elucidate the mechanisms driving surface deformation in subduction zones at a continental scale.
Article
Full-text available
Globally, the largest tsunamigenic earthquakes have occurred along subduction zones. Devastating events exceeding magnitude 9, such as those in Chile, Sumatra, and Japan, struck in regions lacking instrumental records of similar events. Despite the absence of such events along the 1000-kilometer-long Mexican subduction zone, historical and geologic evidence suggests the occurrence of a magnitude 8.6 tsunamigenic earthquake. However, the Guerrero seismic gap has not experienced a high-magnitude earthquake in over 100 years. Here we present results on analyses of sediment grain size, geochemistry, microfossils, magnetic properties, and radiometric and optical stimulated luminescence dating conducted along the Guerrero coast. We provide evidence of a 2000-year history of large tsunamis triggered by potentially large earthquakes. Numerical modeling supports our findings, indicating a magnitude >8 event around the year 1300 in the Guerrero seismic gap. This evidence underscores the importance of assessing earthquake and tsunami potential using long-term evidence and instrumental observations along subduction zones globally.
Article
Full-text available
This book contains true stories that illustrate how to survive-and how not to survive-a tsunami. It is meant for people who live, work, or play along coasts that tsunamis may strike. Such coasts surround most of the Pacific Ocean but also include other areas, such as the shores of the Caribbean, eastern Canada, and the Mediterranean. Although many people call tsunamis 'tidal waves,' they are not related to tides but are rather a series of waves, or 'wave trains,' usually caused by earthquakes. Tsunamis have also been caused by the eruption of some coastal and island volcanoes, submarine landslides, and oceanic impacts of large meteorites. Tsunami waves can become more than 30 feet high as they come into shore and can rush miles inland across low-lying areas. The stories in this book were selected from interviews with people who survived a Pacific Ocean tsunami in 1960. Many of these people, including the nurse at right, contended with the waves near their source, along the coast of Chile. Others faced the tsunami many hours later in Hawaii and Japan. Most of the interviews were done decades later in the 1980's and 1990's. The stories provide a mixed bag of lessons about tsunami survival. Some illustrate actions that reliably saved lives-heeding natural warnings, abandoning belongings, and going promptly to high ground and staying there until the tsunami is really over. Others describe taking refuge in buildings or trees or floating on debris-tactics that had mixed results and can be recommended only as desperate acts.
Article
Full-text available
We present historical and geomorphological evidence of a regularity in earthquake recurrence at three different sites of plate convergence around the Japan arcs. The regularity shows that the larger an earthquake is, the longer is the following quiet period. In other words, the time interval between two successive large earthquakes is approximately proportional to the amount of coseismic displacement of the preceding earthquake and not of the following earthquake. The regularity eanbles us, in principle, to predict the approximate occurrence time of earthquakes. The data set includes 1) a historical document describing repeated measurements of water depth at Murotsu near the focal region of Nankaido earthquakes, 2) precise levelling and 14C dating of Holocene uplifted terraces in the southern boso peninsula facing the Sagami trough, and 3) similar geomorphological data on exposed Holocene coral reefs in Kikai Island along the Ryukyu arc.
Article
Full-text available
The concept of the earthquake cycle is so well established that one often hears statements in the popular media like, "the Big One is overdue" and "the longer it waits, the bigger it will be." Surprisingly, data to critically test the variability in recurrence intervals, rupture displacements, and relationships between the two are almost nonexistent. To generate a long series of earthquake intervals and offsets, we have conducted paleoseismic investigations across the San Andreas fault near the town of Wrightwood, California, excavating 45 trenches over 18 years, and can now provide some answers to basic questions about recurrence behavior of large earthquakes. To date, we have characterized at least 30 prehistoric earthquakes in a 6000-yr-long record, complete for the past 1500 yr and for the interval 3000-1500 B.C. For the past 1500 yr, the mean recurrence interval is 105 yr (31-165 yr for individual intervals) and the mean slip is 3.2 m (0.7-7 m per event). The series is slightly more ordered than random and has a notable cluster of events, during which strain was released at 3 times the long-term average rate. Slip associated with an earthquake is not well predicted by the interval preceding it, and only the largest two earthquakes appear to affect the time interval to the next earthquake. Generally, short intervals tend to coincide with large displacements and long intervals with small displacements. The most significant correlation we find is that earthquakes are more frequent following periods of net strain accumulation spanning multiple seismic cycles. The extent of paleoearthquake ruptures may be inferred by correlating event ages between different sites along the San Andreas fault. Wrightwood and other nearby sites experience rupture that could be attributed to overlap of relatively independent segments that each behave in a more regular manner. However, the data are equally consistent with a model in which the irregular behavior seen at Wrightwood typifies the entire southern San Andreas fault; more long event series will be required to definitively outline prehistoric rupture extents.
Article
An unusual foreshock sequence that began with an earthquake of moment 2.0 × 1021 N m and rupture length of about 150 km preceded the great 1960 Chilean earthquake. Earthquake relocations of the sequence show a progression of seismic activity toward the initiation of the main shock. The rupture length of the great 1960 Chilean earthquake is estimated from the distribution of aftershocks and crustal deformation to be 920 ± 100 km. The source mechanisms of a foreshock and two aftershocks are determined and used to constrain the source mechanism of the main shock. The first event in the sequence ruptured the segment of the Peru-Chile Trench between 37.03°S and 38.74°S, whereas the great 1960 Chilean main shock ruptured the adjoining segment to the south. The great 1960 Chilean main shock initiated within the rupture area of the first event and terminated near the intersection of the Chile Ridge with the Peru-Chile Trench (46.5°S). -Author
Article
The 1964 Alaskan earthquake (Ms ≈ 8.4) involved a segment of the eastern Aleutian arc 800 gm long; the 1960 Chilean earthquake sequence (Ms ≈ 8.5) affected roughly 100 km of the southern Peru-Chile arc. These two major events are strikingly similar in that (1) seismicity was shallow (<70 km), the earthquake focal regions and most of the associated tectonic deformation being between the oceanic trenches and volcanic chains of the two arcs; (2) regional vertical displacements were characterized by broad asymmetric downwarps elongate parallel to the arcs with flanking zones of marked uplift on the seaward sides and minor, possibly local, uplift on the landward sides; and (3) horizontal displacements, where determined by retriangulation, involved systematic shifts in a generally seaward direction and transverse tensile strains across the zones of subsidence. Surface displacements and seismicity for both events are compatible with dislocation models involving predominantly dip-slip movement of 20 meters or more on major complex thrust faults (megathrusts) inclined at average angles of about 9° beneath the eastern Aleutian arc and perhaps 20° beneath the Peru-Chile arc. The thrust-fault mechanism deduced for both the Alaskan and Chilean earthquakes is broadly consistent with the concept that the sectors of the Pacific rim in which they occurred are major zones of convergence along which the oceanic plates progressively underthrust the less mobile America plate. Directions of convergence between lithospheric plates at these arcs as deduced primarily from paleomagnetic data are in reasonably good agreement with the observed earthquake-related deformation; the deduced rates of convergence, however, appear to be too high in the eastern Aleutian arc and too low in the southern Peru-Chile arc. Despite gross similarities in tectonic setting and the present style of earthquake-related deformation, the geologies of the continental margins in the eastern Aleutian arc and southern Peru-Chile arc differ significantly. This difference suggests that Mesozoic and Cenozoic sediments and volcanic rocks conveyed into the eastern Aleutian trench have progressively accreted to the Alaskan continental margin, whereas most or all of the material carried into the southern Peru-Chile trench has disappeared beneath the Chilean continental margin.
Article
Eine Aufstellung der größeren Erdbeben Chiles (angenommene Stärke größer als 7,5) wird vorgelegt. Dieser Aufstellung liegt eine Lochkartenkartei chilenischer Erdbeben mit mehr als 15 000 Eintragungen zugrunde. Für jedes Beben werden die Auswirkungen einschließlich der Tsunami-Beobachtungen beschrieben und Schätzungen der Lage der Epizentren und der Stärke angegeben. Größere Erdbeben treten in Chile in nur wenigen Bebengebieten auf. Diese sind linear im Meer und entlang der Verwerfungen zwischen der Küstenkette und dem Zentraltal angeordnet. In Mittelchile zwischen Valparaiso und Concepción treten größere Erdbeben hauptsächlich im Innern des Landes auf. Südlich von Concepción liegen die größeren Epizentren im Meer. Jedes Herdgebiet liefert voraussagbare seismische und Tsunami-Effekte.
Article
An unusual foreshock sequence that began with an earthquake of moment 2.0 × 1021 N m and rupture length of about 150 km preceded the great 1960 Chilean earthquake. Earthquake relocations of the sequence show a progression of seismic activity toward the initiation of the main shock. The rupture length of the great 1960 Chilean earthquake is estimated from the distribution of aftershocks and crustal deformation to be 920 ± 100 km. The source mechanisms of a foreshock and two aftershocks are determined and used to constrain the source mechanism of the main shock. The first event in the sequence ruptured the segment of the Peru-Chile Trench between 37.03°S and 38.74°S, whereas the great 1960 Chilean main shock ruptured the adjoining segment to the south. The great 1960 Chilean main shock initiated within the rupture area of the first event and terminated near the intersection of the Chile Ridge with the Peru-Chile Trench (46.5°S).
Article
Long-period strain seismogram recorded at Pasadena is used to determine the focal process of the 1960 Chilean earthquake. Synthetic seismograms computed for various fault models are matched with the observed strain seismogram to determine the fault parameters. A low-angle (~ 10°) thrust model with rupture length of 800 km and rupture velocity of 3.5 km/sec is consistent with the observed Rayleigh/Love wave ratio and the radiation asymmetry. A seismic moment of 2.7 . 1030 dyn . cm is obtained for the main shock. This value, together with the estimated fault area of 1.6 . 105 km2, gives an average dislocation of 24 m. The strain seismogram clearly shows unusually long-period (300-600 sec) wave arriving at the P time of a large foreshock which occurred about 15 minutes before the main shock, suggesting a large slow deformation in the epicentral area prior to the major failure. A simple dislocation model shows that a dislocation of 30 m, having a time constant of 300-600 sec, over a fault plane of 800 × 200 km2 is required to explain this precursory displacement. The entire focal process may be envisaged in terms of a large-scale deformation which started rather gradually and eventually triggered the foreshocks and the ``main'' shock. This mechanism may explain the large premonitory deformations documented, but not recorded instrumentally, for several Japanese earthquakes. The moments of the main shock and the precursor add to 6 . 1030 dyn . cm which is large enough to affect the earth's polar motion.
Article
A new magnitude scale Mt is defined by using the logarithm of the maximum amplitude of far-field tsunami waves measured by tide gauges or their substitutes. The Mt scale is experimentally adjusted to the Mw scale introduced by Kanamori (1977), so that the Mt scale measures the seismic moment of a tsunamigenic earthquake as well as the overall size of tsunami at the source. Mt and the conventional tsunami magnitude m are distinct scales. By using many amplitude data of tsunami waves now available the values of Mt are assigned to 65 tsunamigenic earthquakes that occurred in the Pacific area during the period form 1837 to 1974. The 1960 Chilean shock has the largest Mt. 9.4. The 1946 Aleutian (Mt=9.3), the 1837 Chilean (Mt=91/4), and the 1964 Alaskan (Mt=9.1) events follow. Nine great events having Mt=9 or over occurred during this period, and their occurrence is clustered in the years around 1840, 1870, and 1960. Of all the 65 events listed, at least six unusual earthquakes having tsunamis with an amplitude disproportionately large for their surface-wave magnitude Ms are identified from the Mt-Ms relation.