Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Articles

https://doi.org/10.1038/s41561-020-00679-9

Postseismic geodetic signature of cold forearc


mantle in subduction zones
Haipeng Luo   1 ✉ and Kelin Wang   1,2 ✉

A sharp thermal contrast between the cold forearc and the hot arc and backarc is considered fundamental to various
subduction-zone processes. However, direct observational evidence for this contrast is rather limited. If this contrast is pres-
ent, it must cause a rheological contrast in the mantle wedge: elastic in the forearc and viscoelastic in the arc and backarc for
the timescale of earthquake cycles. Here we demonstrate that postseismic deformation following large subduction earthquakes
provides independent evidence for the thermally controlled rheological contrast. Specifically, we show that seaward postseis-
mic motion is deflected upward at the edge of the cold forearc mantle wedge, causing diagnostic uplift just seaward of the
volcanic arc. From numerical simulations of postseismic deformation following the 2011 moment magnitude (Mw) 9 Tohoku-oki,
2010 Mw 8.8 Maule, 2007 Mw 8.4 Bengkulu and 1960 Mw 9.5 Chile earthquakes, together with a global synthesis of postseismic
uplift measurements, we find that cold forearc mantle is present irrespective of the diversity in tectonic settings. Our findings
also indicate that field surveys eight years after the 1960 Chile earthquake provided some of the earliest evidence for visco-
elastic postseismic deformation. We suggest that the established link between long-term thermal processes and short-term
earthquake cycle deformation is important to understanding subduction-zone dynamics.

A
thermal contrast between the cold forearc and the hot arc rheological contrast affecting postseismic deformation in a visco-
and backarc is considered fundamental to subduction-zone elastic Earth (Fig. 1b). The cold nose should have a very high viscos-
dynamics1,2. The contrast reflects a downdip transition of ity and behave elastically in megathrust earthquake cycles. The hot
kinematic coupling along the subduction interface1. Shallower than arc and backarc should have a much lower viscosity and feature fast
70–80 km depths, mantle-wedge material does not travel with the viscoelastic relaxation.
subducting slab (decoupled). Farther downdip, mantle material
travels with the slab at the subduction rate (coupled) (Fig. 1a). The Postseismic uplift due to viscoelastic stress relaxation
shallow decoupling results in a stagnant and cold mantle-wedge Before we lay out geodetic evidence for this process, we first explain
corner referred to as the ‘cold nose’ that provides a condition for the mechanism using a simple two-dimensional (2D) finite-element
the formation of serpentine and other hydrous minerals by incor- model of postseismic deformation with a rheological structure
porating aqueous fluids from the dehydrating subducting slab3,4. similar to that in Fig. 1b. The modelling strategy and the choice of
The deep coupling results in viscous mantle-wedge flow to supply parameters are similar to previous subduction-zone postseismic
heat for melt generation and arc volcanism5 (Fig. 1a). The decou- deformation models17–19, and details are given in Methods. Two of
pling–coupling transition and the associated thermal contrast is the three models summarized in Fig. 2a do not have an elastic cold
also inferred to govern the evolution of the subduction interface in nose, but the other one does as shown in Fig. 2b. The three models
the dip direction6,7. The coldest shallow part of the decoupled inter- all show that postseismic stress relaxation gives rise to a seaward vis-
face hosts great megathrust earthquakes. Along the coupled inter- cous flow in the mantle wedge (Fig. 2b) and a landward flow in the
face beneath the hot arc and backarc, the interface becomes a broad oceanic mantle (not displayed). Note that this postseismic viscous
zone of viscous shear and does not host earthquakes1. flow is a perturbation field added to, and not to be confused with,
However, despite the geodynamic importance of the thermal the long-term background mantle flow (Fig. 1a). The postseismic
contrast, observational evidence is limited. The most direct evi- mantle flow leads to an opposing motion of the near-trench area and
dence is forearc heat-flow measurements8–10, but very few subduc- the area farther landward (Fig. 2a,b), which has been emphasized
tion zones have adequate heat-flow coverage1. Even for one of the in previous models18–20. But the horizontal deformation is insensi-
best-studied subduction zones, the Japan Trench, uncertainties in tive to the rheological contrast (Fig. 2a), although the absence of
heat-flow observations can allow arguments against the presence the cold nose would necessitate the use of a higher mantle-wedge
of the cold nose11. Indirect evidence for the cold nose includes low viscosity to produce similar horizontal velocities.
seismic attenuation (high seismic quality factor (Q))12–14 and geo- By contrast, the short-term vertical deformation is strongly
physical imaging from a few relatively warm subduction zones sup- affected by the rheological contrast (Fig. 2a). Without the cold nose,
porting the presence of serpentinization15,16, but the coverage and the model predicts no or little uplift landward of the rupture zone.
resolution of the geophysical data are both limited. In this work, With the cold nose, there is uplift around its landward edge with
we present unique evidence using what may seem to be unre- a rate observable with geodetic measurements. The reason for the
lated observations: geodetic measurements of vertical postseismic strong signature of the rheological contrast in the uplift pattern is
deformation following great subduction earthquakes. The key that the postseismic seaward motion of the mantle wedge and upper
point is that the thermal contrast, if present, must lead to a plate is deflected upwards upon encountering the elastic cold nose

School of Earth and Ocean Sciences, University of Victoria, Victoria, British Columbia, Canada. 2Pacific Geoscience Centre, Geological Survey of Canada,
1

Natural Resources Canada, Sidney, British Columbia, Canada. ✉e-mail: hpluo@uvic.ca; kelin.wang@canada.ca

104 Nature Geoscience | VOL 14 | February 2021 | 104–109 | www.nature.com/naturegeoscience


NATurE GEOSCIEnCE Articles
a Low heat flow High heat flow rheological contrast is irrespective of the thermal state of the
subduction zone, which depends mainly on the age of the subduct-
Ocean Cold and
ing plate (Extended Data Figs. 1 and 8). In the absence of other com-
ic crus Hot and
t H2O stagnant flowing plications, the forearc uplift in the postseismic period is expected
Serpentinization Melting
to be offset by later interseismic subsidence25,26. Here, however, we
Subd
ucting High Q Low Q focus only on the postseismic uplift.
plate
We choose four of the events summarized in Extended Data
Ba
The megathrust c
ma kgrou Fig. 8 to construct 3D postseismic deformation models for a more
ntle nd
flow quantitative analysis. The four events represent a wide range of data
type, data quality, event size, and the age and hence thermal state of
Slab–mantle
full coupling the subducting plate. We use a spherical-Earth finite-element mod-
(≥˜70–80 km) elling code that employs bi-viscous Burgers rheology (Fig. 1b; for
example, refs. 27,28), and we invoke the actual slab geometry29 and
coseismic slip distribution previously inferred from geodetic, seis-
b Postseismic horizontal and vertical displacements mic and tsunami observations (details in Methods and Extended
Data Fig. 9). In real Earth, the rigidity of the cold nose may be influ-
CNE Arc enced by the temperature field, and the definition of the CNE is
Trench
not as ideal as shown in Figs. 1b and 2b, although fine details may
Viscoelastic
mantle wedge
not be resolvable by present geodetic observations. For simplicity, in
Elastic upper plate
(lower viscosity) modelling the four earthquakes, we assume that the CNE is located
Rup where the depth of the subduction interface is 70 km, consistent
ture
Elastic
cold nose with the thermal argument of ref. 1.
Po
Viscoelastic
ssi
ble
Post
seis
mic
The main results along a margin-normal corridor are sum-
afte
oceanic mantle Ela
stic rsli
p Vis
cou
marized in Fig. 3 compared with observations. A map view of the
sub
du
ctin
s flo
w observations and model results for the largest of the four events,
Burgers rheology
µK µ
gp
late the 1960 Mw 9.5 Chile earthquake, is shown in Fig. 4. Map views
η
M M
of model-predicted postseismic deformation of the other three
ηK events compared with GNSS observations are shown in Extended
Data Figs. 2, 4 and 6. Postseismic displacement time series from our
Fig. 1 | Schematic illustration of thermal and rheological structure in models are compared with GNSS observations in Extended Data
subduction zones. a, Thermal structure. The sharp contrast between cold Figs. 3, 5 and 7. Details for the four events and their models are
forearc and hot arc–backarc is inferred from limited heat-flow data and described as follows.
seismic attenuation (high or low Q). b, Rheological structure. The cold nose Among all the events examined (Extended Data Fig. 8), the 2011
exhibits elastic behaviour in earthquake cycles. η and μ are viscosity Mw 9 Tohoku-oki earthquake yielded the highest-quality coseismic
and rigidity, respectively, and subscripts M and K denote Maxwell and and postseismic deformation data. Japan Trench that hosted this
Kelvin, respectively. The postseismic viscous flow in b is induced by event is about the coldest subduction zone examined, for its old sub-
megathrust rupture, but the background mantle flow in a is driven by the ducting plate (~130 Ma) and fast subduction rate (>8 cm yr–1). Over
subducting plate. 40 coseismic slip distribution models have been obtained for this
earthquake by inverting various coseismic observations30. For our
modelling, we use a slip distribution that is an average of 19 of these
(Fig. 2b). An enlarged view in Fig. 2c shows the absence of this models that included seafloor GNSS measurements in the inversion
deflection in models that do not include the cold nose. As shown by (Extended Data Fig. 9). The postseismic deformation following this
a suite of sensitivity tests (Fig. 2d), the wavelength of the uplift zone earthquake was recorded by a dense land-based GNSS network and
and the location of the peak uplift are controlled by the size of the some seafloor sites31,32 (Extended Data Figs. 2 and 3). The landward
cold nose. If the cold-nose edge (CNE) extends farther landward, motion of the near-trench area that opposes the seaward motion
the zone of uplift is also farther landward and is broader. of the rest of the network and its geodynamic implications were
studied by Sun et al.18. Important to the theme of this paper is the
Modelling modern geodetic postseismic observations pronounced uplift just seawards of the volcanic arc extending to
We have conducted a global synthesis of observations of short-term the outer coast (Fig. 3). Five years after the earthquake, the peak
vertical deformation in the forearc region following great or large uplift is ~0.4 m and is located around the coastline where the plate
(moment magnitude (Mw) ≥ 7.8) subduction earthquakes (Extended interface depth is ~60–70 km (Extended Data Fig. 2). Our model
Data Figs 1–8) with the results summarized in Extended Data Fig. that includes the cold nose readily explains this uplift (Fig. 3). The
8. For most events that occurred in this century, continuous Global model-predicted postseismic deformation near the volcanic arc is
Navigation Satellite System (GNSS) observations were available, but not sensitive to details of the coseismic slip distribution because the
very few had dense spatial coverage just seaward of the volcanic arc. area is quite far from the main rupture zone.
For some earlier events such as the 1946 Nankai, 1957 and 1965 Postseismic deformation following the 2010 Mw 8.8 Maule earth-
Aleutian, and 1960 Chile earthquakes, vertical postseismic defor- quake was also recorded by a large number of GNSS sites (Extended
mation was constrained by limited levelling, tide gauge or coastal Data Figs. 4 and 5). Compared with the Japan Trench, the cen-
field observations21–23. The quantity and quality of the observa- tral Chile subduction zone has a much younger subducting plate
tions vary tremendously between events, but postseismic uplift (~33 Ma), slightly slower subduction rate (6.6 cm yr–1) and thus a
just seawards of the volcanic arc is observed or inferred wherever warmer thermal regime. The coseismic slip had been determined
relevant data are available. Even for an event that occurred in the in a number of studies by inverting GNSS and other data. From the
seventeenth century (southern Kuril), available palaeo-postseismic many published slip models, we use the average of nine of them that
data show pronounced postseismic uplift in the region where included GNSS data in the inversion (Extended Data Fig. 9). Five
the plate interface depth is 60–80 km (ref. 24). It is important to years after the earthquake, GNSS measurements show broad uplift
note that the postseismic uplift supporting the presence of the where the plate interface depth is around 80 km (ref. 33; Extended

Nature Geoscience | VOL 14 | February 2021 | 104–109 | www.nature.com/naturegeoscience 105


Articles NATurE GEOSCIEnCE

a
2

Normalized 5 yr displacement (%) 0

–2 Vertical Horizontal
Cold nose
No cold nose
–4
No cold nose
but higher viscosity
–6
0 50 100 150 200 250 300

b CNE Arc
0

20

Depth 70 km
40 Normalized 5 yr Panel c
Depth (km)

displacement

60 3%

80
Slip

100

0 50 100 150 200 250 300


Distance from trench (km)

c 0
d 30
1.5

Normalized 5 yr
uplift (%)
50
Interace depth at
Depth (km)

CNE (km)

20
0

70

40
90

160 180 200 220 100 150 200 250 300 350
Distance from trench (km) Distance from trench (km)

Fig. 2 | 2D models of postseismic deformation to illustrate the physical process discussed in this paper. a, Vertical and horizontal postseismic
displacements (normalized by peak coseismic slip value) five years after an earthquake for three models. Only the model including an elastic cold nose
features pronounced uplift around the CNE. b, Cross-section view of postseismic displacement field of the cold-nose model. The postseismic motion is
deflected upwards by the cold nose. Inset shows coseismic slip distribution along the megathrust for all the models in a. c, Enlarged view of the rectangular
area in b comparing displacement directions in the three colour-coded models in a. d, Sensitivity tests showing that the size of the elastic cold nose
controls the postseismic forearc uplift.

Data Fig. 4). Our model results show that this uplift is consistent sites just seaward of the volcanic arc35. Again, our model readily
with the presence of the cold nose (Fig. 3 and Extended Data Figs. explains this uplift by invoking the cold nose (Fig. 3 and Extended
4 and 5). The presence of the cold nose was also inferred by Weiss Data Figs. 6 and 7). The more-complex uplift and subsidence pat-
et al.34, who directly inverted the Maule postseismic geodetic data to tern closer to the trench and away from our region of interest can be
estimate mantle rheology. explained by having local afterslip (Extended Data Fig. 6).
The North Sumatra subduction zone has a slow margin-normal
convergence rate (75% of ~5.9 cm yr–1 oblique convergence) and a Postseismic deformation of the 1960 giant Chile
moderate slab age (~69 Ma). The 2007 Mw 8.4 Bengkulu earthquake earthquake
is an example of a smaller event recorded by sparse measurements. The most remarkable example is the 1960 Mw 9.5 Chile earth-
Despite the limited data coverage, the smaller size of the earth- quake, by far the largest instrumentally recorded earthquake,
quake and the many tectonic differences from the other examples in which involved the warmest subduction-zone segment studied here
Fig. 3, postseismic uplift was systematically observed at GNSS (Extended Data Fig. 8). G. Plafker21 conducted a geological field

106 Nature Geoscience | VOL 14 | February 2021 | 104–109 | www.nature.com/naturegeoscience


NATurE GEOSCIEnCE Articles
4). Therefore, these measurements provided some of the earliest
0.5 2011 Mw 9 Tohoku-oki (5 yr) evidence for postseismic deformation and the presence of the cold
Uplift (m)

0 nose. Independent evidence for the postseismic uplift here came


from tide gauge measurements at Puerto Montt39, the location of
–0.5 Outer coast
Coseismic slip CNE Arc which is shown in Fig. 4. Modern GNSS measurements indicate that
the uplift is continuing today but at a diminished rate40,41.
100 150 200 250 300

0.5 Other contributing factors


2010 Mw 8.8 Maule (5 yr)
Some of the uplift signals summarized in Extended Data Fig. 8 were
Uplift (m)

0
Coseismic
Outer coast
previously noticed by other researchers. In the past, the uplift was
slip CNE Arc
–0.5
attributed to frictional slip42,43, or almost equivalently postseismic
100 150 200 250 shear of a thin zone of extremely low viscosity44, at depths as large as
70 to >100 km. The very deep frictional or viscous slip of the sub-
0.5 duction interface is less compatible with the full coupling of the slab
2007 Mw 8.4 Bengkulu (5 yr)
Uplift (m)

0
and mantle wedge at such depths required by the presence of the hot
Outer sub-arc mantle wedge1,2,15 (Fig. 1).
Coseismic slip coast CNE Arc
–0.5 The mechanism discussed in this paper is based on geologically
100 150 200 250 reasonable structural conditions. Similar postseismic uplift may be
produced in some models without invoking the cold nose of the
1.0 forearc mantle wedge if less-reasonable conditions are introduced.
1960 Mw 9.5 Chile (8 yr)
For example, if one assumes a very thick elastic upper plate in the
0.5 Modelled
arc and backarc area by neglecting its very warm thermal condition,
Uplift (m)

5 yr
0 minor uplift occurs near the volcanic front in a model without the
Outer coast
cold nose (Extended Data Fig. 10a) unless the coseismic slip of the
–0.5
Coseismic
megathrust is assumed to be deeper (Extended Data Fig. 10b).
–1.0 slip CNE Arc
Some recent postseismic deformation models used
temperature-dependent rock rheology45–47. By invoking
100 150 200 250 subduction-zone thermal structures similar to those in refs. 1,2,15,
Distance from trench (km) these models all included the mantle-wedge cold nose (Fig. 1), and
they all predicted postseismic uplift near the volcanic area qualita-
Fig. 3 | Model postseismic vertical deformation for four subduction tively similar to our models. Although these previous studies did
earthquakes compared with observations. See Extended Data Fig. 8 for not recognize the importance of the cold nose but speculated on
data type and sources and Extended Data Fig. 9 for model parameters. For other mechanisms to cause the uplift45 or even assumed it to be
the Chile event, only 8 yr postseismic data are available, but model uplift coseismic47, their results in effect corroborate our findings. More
for 5 years after earthquake is also shown for comparison. Map views of important, these results also demonstrate the robustness of the
model results, postseismic observations, coseismic slip distribution and role of the elastic cold nose in controlling postseismic deformation
location of the display profile are shown in, from top to bottom, Extended regardless of parametric details of fault and rock rheology.
Data Figs. 2, 4 and 6 and Fig. 4. Data (symbols) shown here are from
within 200 km of display profile for the 2007 Bengkulu earthquake but Online content
100 km for other events. Circles for the Tohoku-oki event are seafloor GNSS Any methods, additional references, Nature Research report-
measurements. Model-predicted displacement time series are compared ing summaries, source data, extended data, supplementary infor-
with GNSS observations in Extended Data Figs. 3, 5 and 7. mation, acknowledgements, peer review information; details of
author contributions and competing interests; and statements of
data and code availability are available at https://doi.org/10.1038/
survey of the rupture area in 1968, eight years after the earthquake s41561-020-00679-9.
and two decades before the birth of GNSS. In terms of constrain-
ing earthquake deformation, these geological data are equivalent to Received: 12 May 2020; Accepted: 3 December 2020;
campaign-style geodetic measurements. The uplift and subsidence Published online: 18 January 2021
pattern mapped by Plafker led to the recognition of the subduc-
tion megathrust being the source of the rupture and hence became References
a landmark in the advancement of the plate tectonics theory. 1. Wada, I. & Wang, K. Common depth of slab–mantle decoupling: reconciling
However, the belt of uplift sandwiched between the main zone of diversity and uniformity of subduction zones. Geochem. Geophys. Geosyst. 10,
coseismic subsidence and the volcanic arc (Fig. 4) has been enig- Q10009 (2009).
2. Syracuse, E. M., van Keken, P. E. & Abers, G. A. The global range of
matic for the past five decades. subduction zone thermal models. Phys. Earth Planet. Inter. 183,
The uplift in this belt peaks to ~1 m where the slab interface 73–90 (2010).
depth is ~70–80 km (Fig. 4b). Plafker’s observations were meant to 3. Hyndman, R. D. & Peacock, S. M. Serpentinization of the forearc mantle.
delineate coseismic deformation, but no rupture models can explain Earth Planet. Sci. Lett. 212, 417–432 (2003).
such a large uplift so far away from the rupture zone21,36,37, unless 4. Krawczyk, C. M. et al. in The Andes: Active Subduction Orogeny
(eds Oncken, O. et al.) 171–192 (Springer, 2006).
physically implausible, exceedingly deep slip is invoked38. The 5. Agard, P., Plunder, A., Angiboust, S., Bonnet, G. & Ruh, J. The subduction
most recent rupture model predicts coseismic subsidence in this plate interface: rock record and mechanical coupling (from long to short
area (Fig. 4b). However, we should recall that Plafker’s measure- timescales). Lithos 320, 537–566 (2018).
ments were made eight years after the earthquake and thus must 6. Gao, X. & Wang, K. Strength of stick-slip and creeping subduction
contain substantial postseismic deformation signal. On the basis of megathrusts from heat flow observations. Science 345, 1038–1041 (2014).
7. Barbot, S. Frictional and structural controls of seismic super-cycles at the
the mechanism of postseismic deformation illustrated in Fig. 2, we Japan trench. Earth Planets Space 72, 63 (2020).
can firmly conclude that this uplift is primarily viscoelastic post- 8. Furukawa, Y. Depth of the decoupling plate interface and thermal structure
seismic deformation in the presence of the cold nose (Figs. 3 and under arcs. J. Geophys. Res. 98, 20005–20013 (1993).

Nature Geoscience | VOL 14 | February 2021 | 104–109 | www.nature.com/naturegeoscience 107


Articles NATurE GEOSCIEnCE

a 1960 Mw 9.5 earthquake b Coseismic model c 8 yr postseismic model


41° S
4

vertical (m)
Modelled
2
0
–2
40° S
–4
Profile in
Fig. 3
Data, 1 m
100 km 42° S
Model, 1 m
42° S
6.6 –1
r
cm y
Nazca
plate

44° S Panels
b and c 43° S

100 km
100 km

70

90
90

50
70
50
South
America
46° S plate

76° W 74° W 72° W 75° W 74° W 73° W 75° W 74° W 73° W

Fig. 4 | Observed and modelled deformation associated with the 1960 Mw 9.5 Chile earthquake. a, Distribution of the Plafker measurements (purple
squares)21 and coseismic slip distribution contoured at 10 m intervals37. Red triangles represent locations of volcanoes. The purple circle marks the location of
the tide gauge at Puerto Montt. b, Comparison of Plafker measurements (orange arrows) with coseismic deformation (green arrows and background colour)
modelled using the shown coseismic slip distribution (same contours as in a). c, Similar to b, but the model deformation is for eight years after the earthquake.

9. Lewis, T. J. et al. Subduction of the Juan de Fuca plate: thermal consequences. 26. Li, S., Fukuda, J. I. & Oncken, O. Geodetic evidence of time-dependent
J. Geophys. Res. 93, 15207–15225 (1988). viscoelastic interseismic deformation driven by megathrust locking in the
10. von Herzen, R. et al. A constraint on the shear stress at the Pacific–Australian southwest Japan subduction zone. Geophys. Res. Lett. 47, e2019GL085551
plate boundary from heat flow and seismicity at the Kermadec forearc. (2020).
J. Geophys. Res. 106, 6817–6833 (2001). 27. Bürgmann, R. & Dresen, G. Rheology of the lower crust and upper mantle:
11. Horiuchi, S. S. & Iwamori, H. A consistent model for fluid distribution, evidence from rock mechanics, geodesy, and field observations. Annu. Rev.
viscosity distribution, and flow-thermal structure in subduction zone. Earth Planet. Sci. 36, 531–567 (2008).
J. Geophys. Res. 121, 3238–3260 (2016). 28. Pollitz, F. F. Lithosphere and shallow asthenosphere rheology from
12. Tsumura, N., Matsumoto, S., Horiuchi, S. & Hasegawa, A. Three-dimensional observations of post-earthquake relaxation. Phys. Earth Planet Inter. 293,
attenuation structure beneath the northeastern Japan arc estimated from 106271 (2019).
spectra of small earthquakes. Tectonophysics 319, 241–260 (2000). 29. Hayes, G. P., Wald, D. J. & Johnson, R. L. Slab1.0: a three-dimensional
13. Wang, Z., Zhao, D., Liu, X., Chen, C. & Li, X. P and S wave attenuation model of global subduction zone geometries. J. Geophys. Res. 117,
tomography of the Japan subduction zone. Geochem. Geophys. Geosyst. 18, B01302 (2012).
1688–1710 (2017). 30. Wang, K. et al. Learning from crustal deformation associated with the M9
14. Wei, S. S. & Wiens, D. A. P-wave attenuation structure of the Lau back-arc 2011 Tohoku-oki earthquake. Geosphere 14, 552–571 (2018).
basin and implications for mantle wedge processes. Earth Planet. Sci. Lett. 31. Ozawa, S. et al. Preceding, coseismic, and postseismic slips of the 2011
502, 187–199 (2018). Tohoku earthquake, Japan. J. Geophys. Res. 117, B07404 (2012).
15. Abers, G. A., Van Keken, P. E. & Hacker, B. R. The cold and relatively 32. Yokota, Y., Ishikawa, T. & Watanabe, S. I. Seafloor crustal deformation data
dry nature of mantle forearcs in subduction zones. Nat. Geosci. 10, along the subduction zones around Japan obtained by GNSS-A observations.
333–337 (2017). Sci. Data 5, 180182 (2018).
16. Bostock, M. G., Hyndman, R. D., Rondenay, S. & Peacock, S. M. An inverted 33. Li, S. et al. Spatiotemporal variation of mantle viscosity and the presence of
continental Moho and serpentinization of the forearc mantle. Nature 417, cratonic mantle inferred from 8 years of postseismic deformation following
536–538 (2002). the 2010 Maule, Chile, earthquake. Geochem. Geophys. Geosyst. 19,
17. Wang, K., Hu, Y. & He, J. Deformation cycles of subduction earthquakes in a 3272–3285 (2018).
viscoelastic Earth. Nature 484, 327–332 (2012). 34. Weiss, J. R. et al. Illuminating subduction zone rheological properties in the
18. Sun, T. et al. Prevalence of viscoelastic relaxation after the 2011 Tohoku-oki wake of a giant earthquake. Sci. Adv. 5, eaax6720 (2019).
earthquake. Nature 514, 84–87 (2014). 35. Feng, L. et al. A unified GPS-based earthquake catalog for the Sumatran plate
19. Sun, T., Wang, K. & He, J. Crustal deformation following great subduction boundary between 2002 and 2013. J. Geophys. Res. 120, 3566–3598 (2015).
earthquakes controlled by earthquake size and mantle rheology. J. Geophys. 36. Moreno, M. S., Bolte, J., Klotz, J. & Melnick, D. Impact of megathrust
Res. 123, 5323–5345 (2018). geometry on inversion of coseismic slip from geodetic data: application to the
20. Freed, A. M. et al. Resolving depth-dependent subduction zone viscosity and 1960 Chile earthquake. Geophys. Res. Lett. 36, L16310 (2009).
afterslip from postseismic displacements following the 2011 Tohoku-oki, 37. Ho, T. C., Satake, K., Watada, S. & Fujii, Y. Source estimate for the 1960 Chile
Japan earthquake. Earth Planet. Sci. Lett. 459, 279–290 (2017). earthquake from joint inversion of geodetic and transoceanic tsunami data.
21. Plafker, G. & Savage, J. C. Mechanism of the Chilean earthquakes of May 21 J. Geophys. Res. 124, 2812–2828 (2019).
and 22, 1960. Geol. Soc. Am. Bull. 81, 1001–1030 (1970). 38. Barrientos, S. E. & Ward, S. N. The 1960 Chile earthquake: inversion for slip
22. Thatcher, W. The earthquake deformation cycle at the Nankai Trough, distribution from surface deformation. Geophys. J. Int. 103, 589–598 (1990).
southwest Japan. J. Geophys. Res. 89, 3087–3101 (1984). 39. Barrientos, S. E., Plafker, G. & Lorca, E. Postseismic coastal uplift in southern
23. Wahr, J. & Wyss, M. Interpretation of postseismic deformation with a Chile. Geophys. Res. Lett. 19, 701–704 (1992).
viscoelastic relaxation model. J. Geophys. Res. 85, 6471–6477 (1980). 40. Wang, K. et al. Crustal motion in the zone of the 1960 Chile earthquake:
24. Sawai, Y. et al. Transient uplift after a 17th-century earthquake along the detangling earthquake-cycle deformation and forearc-sliver translation.
Kuril subduction zone. Science 306, 1918–1920 (2004). Geochem. Geophys. Geosyst. 8, Q10010 (2007).
25. Nishimura, T. Pre-, co-, and post-seismic deformation of the 2011 41. Luo, H. et al. A recent increase in megathrust locking in the southernmost
Tohoku-oki earthquake and its implication to a paradox in short-term and rupture area of the giant 1960 Chile earthquake. Earth Planet. Sci. Lett. 537,
long-term deformation. J. Disaster Res. 9, 294–302 (2004). 116200 (2020).

108 Nature Geoscience | VOL 14 | February 2021 | 104–109 | www.nature.com/naturegeoscience


NATurE GEOSCIEnCE Articles
42. Linde, A. T. & Silver, P. G. Elevation changes and the great 1960 Chilean 46. Peña, C. et al. Impact of power-law rheology on the viscoelastic relaxation
earthquake: support for aseismic slip. Geophys. Res. Lett. 16, pattern and afterslip distribution following the 2010 Mw 8.8 Maule
1305–1308 (1989). earthquake. Earth Planet. Sci. Lett. 542, 116292 (2020).
43. Muto, J. et al. Coupled afterslip and transient mantle flow after the 2011 47. van Dinther, Y., Preiswerk, L. E. & Gerya, T. V. A secondary zone
Tohoku earthquake. Sci. Adv. 5, eaaw1164 (2019). of uplift due to megathrust earthquakes. Pure Appl. Geophys. 176,
44. Klein, E., Fleitout, L., Vigny, C. & Garaud, J. D. Afterslip and viscoelastic 4043–4068 (2019).
relaxation model inferred from the large-scale post-seismic deformation
following the 2010 Mw 8.8 Maule earthquake (Chile). Geophys. J. Int. 205,
1455–1472 (2016).
45. Peña, C. et al. Role of lower crust in the postseismic deformation of the 2010 Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
Maule earthquake: insights from a model with power-law rheology. Pure published maps and institutional affiliations.
Appl. Geophys. 176, 3913–3928 (2019). © Crown 2021

Nature Geoscience | VOL 14 | February 2021 | 104–109 | www.nature.com/naturegeoscience 109


Articles NATurE GEOSCIEnCE

Methods 57. Ioki, K. & Tanioka, Y. Re-estimated fault model of the 17th century great
To model postseismic deformation, we use the spherical-Earth finite-element earthquake off Hokkaido using tsunami deposit data. Earth Planet. Sci. Lett.
code PGCviscl-3D developed by J. He at the Pacific Geoscience Centre, Geological 433, 133–138 (2016).
Survey of Canada. The model includes elastic plates and viscoelastic mantle with the 58. Kogan, M. G. et al. The mechanism of postseismic deformation triggered by
bi-viscous Burgers rheology. The Burgers body is composed of a Kelvin solid (rigidity the 2006–2007 great Kuril earthquakes. Geophys. Res. Lett. 38 (2011).
μK and viscosity ηK) in series with a Maxwell fluid (rigidity μM and viscosity ηM), 59. Kogan, M. G. et al. Rapid postseismic relaxation after the great 2006–2007
representing the transient and steady-state rheology, respectively (Fig. 1b). We assume Kuril earthquakes from GPS observations in 2007–2011. J. Geophys. Res. 118,
a uniform rock density and Poisson’s ratio of 3,300 kg m–3 and 0.25, respectively, and 3691–3706 (2013).
rigidity values of the elastic plates and viscoelastic mantle of 48 GPa and 64 GPa, 60. Nicolsky, D. J., Freymueller, J. T., Witter, R. C., Suleimani, E. N. & Koehler, R.
respectively. The mantle-wedge viscosity is assumed to be an order of magnitude lower D. Evidence for shallow megathrust slip across the Unalaska seismic gap
than the oceanic mantle, to account for the higher temperature and the presence of during the great 1957 Andreanof Islands earthquake, eastern Aleutian Islands,
fluids from the subducting plate and melts17, similar to many other subduction-zone Alaska. Geophys. Res. Lett. 43, 10–328 (2016).
earthquake cycle deformation models19,20,48,49. Coseismic slip and afterslip are assigned 61. Plafker, G. Tectonics of the March 27, 1964 Alaska Earthquake Professional
using the split-node method50. The effect of gravity is approximated using the Paper 543–I (USGS, 1969); https://pubs.usgs.gov/pp/0543i/
stress-advection approach51,52. 62. Hergert, T. & Heidbach, O. New insights into the mechanism of postseismic
The geometry of the slab in the 2D models shown in Fig. 2 is designed to stress relaxation exemplified by the 23 June 2001 Mw = 8.4 earthquake in
represent a typical subduction zone. In these models, the thicknesses of the southern Peru. Geophys. Res. Lett. 33 (2006).
subducting and upper plates are 40 and 30 km, respectively. The Maxwell (ηM) and 63. Bevis, M., Bedford, J. & Caccamise, D. J. II in Geodetic Time Series Analysis in
Kelvin (ηK) viscosities are 5 × 1019 and 5 × 1018 Pa s, respectively, for the oceanic Earth Sciences (eds Montillet, J. P. & Bos, M. S.) 1–27 (Springer, 2020).
mantle, and 5 × 1018 (ηM) and 5 × 1017 (ηK) for the mantle wedge. In the cold-nose 64. Müller, R. D., Sdrolias, M., Gaina, C. & Roest, W. R. Age, spreading rates, and
model, the cold nose extends to where the depth of the subduction interface is spreading asymmetry of the world’s ocean crust. Geochem. Geophys. Geosyst.
70 km except for those shown in Fig. 2d for sensitivity tests. In the model without 9, Q04006 (2008).
the cold nose but with a higher mantle-wedge viscosity (Fig. 2a), ηM = 1 × 1019 Pa s. 65. Delouis, B., Nocquet, J. M. & Vallée, M. Slip distribution of the February 27,
The models of the four great earthquakes summarized in Fig. 3I each require 2010 Mw = 8.8 Maule earthquake, central Chile, from static and high-rate GPS,
site-specific 3D geometry of the subducting plate and parameter values. We InSAR, and broadband teleseismic data. Geophys. Res. Lett. 37, L17305 (2010).
employ the slab geometry of slab 1.0 (ref. 29) except for the 2011 Mw 9 Tohoku-oki 66. Lin, Y. N. N. et al. Coseismic and postseismic slip associated with the 2010
earthquake, in which the slab geometry is smoothed from ref. 18. The viscosity Maule earthquake, Chile: characterizing the Arauco Peninsula barrier effect.
values vary slightly between models to fit surface observations (Extended Data Fig. J. Geophys. Res. 118, 3142–3159 (2013).
9). Unlike the 2D models shown in Fig. 2, where we purposely leave out afterslip to 67. Lorito, S. et al. Limited overlap between the seismic gap and coseismic slip of
focus on the main physical process, we include afterslip for the 3D real-earthquake the great 2010 Chile earthquake. Nat. Geosci. 4, 173–177 (2011).
models wherever observational constraints are adequate. However, our results 68. Luttrell, K. M., Tong, X., Sandwell, D. T., Brooks, B. A. & Bevis, M. G.
indicate that its effect is minor. We incorporate a thin layer of ~3–5 km between the Estimates of stress drop and crustal tectonic stress from the 27 February 2010
slab and the cold nose in the depth range of ~55–70 km, with the same viscosity as Maule, Chile, earthquake: implications for fault strength. J. Geophys. Res. 116,
the rest of the viscoelastic mantle wedge, to simulate the viscous or semi-viscous B11401 (2011).
shear of materials in postseismic deformation18,53. We use a thinner upper plate 69. Moreno, M. et al. Toward understanding tectonic control on the Mw 8.8 2010
at the volcanic arc to represent lower stiffness in this area54,55. The location and Maule Chile earthquake. Earth Planet. Sci. Lett. 321, 152–165 (2012).
width of the thinner part of the upper plate are based mainly on the distribution 70. Pollitz, F. F. et al. Coseismic slip distribution of the February 27, 2010 Mw 8.8
of arc volcanoes. Postseismic observations for these earthquakes and the other Maule, Chile earthquake. Geophys. Res. Lett. 38, L09309 (2011).
earthquakes that are not modelled but are shown in Extended Data Fig. 1 are 71. Tong, X. et al. The 2010 Maule, Chile earthquake: downdip rupture limit
summarized in Extended Data Fig. 8 (refs. 56–73). revealed by space geodesy. Geophys. Res. Lett. 37, L24311 (2010).
72. Vigny, C. et al. The 2010 Mw 8.8 Maule megathrust earthquake of central
Chile, monitored by GPS. Science 332, 1417–1421 (2011).
Data availability 73. Yue, H. et al. Localized fault slip to the trench in the 2010 Maule, Chile Mw =
All the data used in this study are from published literature as referenced in 8.8 earthquake from joint inversion of high-rate GPS, teleseismic body waves,
Extended Data Fig. 8. The data used in Extended Data Fig. 1 are presented in the InSAR, campaign GPS, and tsunami observations. J. Geophys. Res. 119,
Source data. Source data are provided with this paper. 7786–7804 (2014).
74. Bird, P. An updated digital model of plate boundaries. Geochem. Geophys.
Code availability Geosyst. 4, 1027 (2003).
The computer code used in this study is available from the authors upon reasonable 75. Nakagawa, H. Development and validation of GEONET new analysis strategy
request. (version 4) (in Japanese). J. Geogr. Surv. Inst. 118, 1–8 (2009).
76. Tsang, L. L. et al. Afterslip following the 2007 Mw 8.4 Bengkulu earthquake in
Sumatra loaded the 2010 Mw 7.8 Mentawai tsunami earthquake rupture zone.
References J. Geophys. Res. 121, 9034–9049 (2016).
48. Moreno, M. et al. Heterogeneous plate locking in the South–Central Chile
subduction zone: building up the next great earthquake. Earth Planet. Sci.
Lett. 305, 413–424 (2011). Acknowledgements
49. Suito, H. Importance of rheological heterogeneity for interpreting viscoelastic We thank X. Zhou for discussion. H.L. was supported by a James A. & Laurette Agnew
relaxation caused by the 2011 Tohoku-oki earthquake. Earth Planets Space 69, Memorial Award, a University of Victoria Graduate Award, Graduate Support from
21 (2017). Ocean Networks Canada, and Discovery Grant RGPIN-2016-03738 to K.W. from the
50. Melosh, H. J. & Raefsky, A. A simple and efficient method for introducing faults Natural Sciences and Engineering Research Council of Canada. This is Geological Survey
into finite element computations. Bull. Seismol. Soc. Am. 71, 1391–1400 (1981). of Canada contribution 20200524.
51. Peltier, W. R. The impulse response of a Maxwell Earth. Rev. Geophys. 12,
649–668 (1974). Author contributions
52. Wang, K., He, J., Dragert, H. & James, T. S. Three-dimensional viscoelastic H.L. and K.W. together designed the research and did the writing. H.L. carried out data
interseismic deformation model for the Cascadia subduction zone. Earth synthesis and analyses and numerical modelling.
Planets Space 53, 295–306 (2001).
53. Hu, Y., Bürgmann, R., Uchida, N., Banerjee, P. & Freymueller, J. T. Competing interests
Stress-driven relaxation of heterogeneous upper mantle and time-dependent The authors declare no competing interests.
afterslip following the 2011 Tohoku earthquake. J. Geophys. Res. 121,
385–411 (2016).
54. Hu, Y., Bürgmann, R., Freymueller, J. T., Banerjee, P. & Wang, K. Additional information
Contributions of poroelastic rebound and a weak volcanic arc to the Extended data is available for this paper at https://doi.org/10.1038/s41561-020-00679-9.
postseismic deformation of the 2011 Tohoku earthquake. Earth Planets Space Supplementary information The online version contains supplementary material
66, 106 (2014). available at https://doi.org/10.1038/s41561-020-00679-9.
55. Itoh, Y., Wang, K., Nishimura, T. & He, J. Compliant volcanic arc and backarc
Correspondence and requests for materials should be addressed to H.L. or K.W.
crust in southern Kurile suggested by interseismic geodetic deformation.
Geophys. Res. Lett. 46, 11790–11798 (2019). Peer review information Nature Geoscience thanks Sylvain Barbot, Onno Oncken and
56. Itoh, Y., Nishimura, T., Ariyoshi, K. & Matsumoto, H. Interplate slip following Fred Pollitz for their contribution to the peer review of this work. Primary Handling
the 2003 Tokachi-oki earthquake from ocean bottom pressure gauge and land Editors: Tamara Goldin; Stefan Lachowycz.
GNSS data. J. Geophys. Res. 124, 4205–4230 (2019). Reprints and permissions information is available at www.nature.com/reprints.

Nature Geoscience | www.nature.com/naturegeoscience


NATurE GEOSCIEnCE Articles

Extended Data Fig. 1 | Global synthesis of postseismic uplift following subduction earthquakes observed in the forearc near the volcanic arc. Events are
labelled with index numbers (circled) as listed in Extended Data Fig. 8, year of occurrence, moment magnitude, and age of the subducting plate at trench.
Black-solid line boxes approximately outline the map areas of Fig. 4c and Extended Data Figs. 2, 4, and 6. Plate boundaries (grey lines) are from ref. 74.
Locations of volcanoes are from Aster Volcano Archive (NASA, METI; https://gbank.gsj.jp/vsidb/image/).

Nature Geoscience | www.nature.com/naturegeoscience


Articles NATurE GEOSCIEnCE

Extended Data Fig. 2 | Map view of observed and modelled postseismic deformation 5 years after the 2011 Mw 9 Tohoku-oki earthquake. a, Horizontal
component. Coseismic slip distribution (Extended Data Fig. 9) is contoured at 10 m interval (coppery lines). Cumulative afterslip, contoured at 1 m interval
(grey lines), is slightly modified from ref. 30. b, Vertical component. The black solid line shows the location of the corridor in Fig. 3. Land GNSS data are
based on GNSS daily solutions from the Geospatial Information Authority of Japan (GSI)31,75, and seafloor GNSS data are from Japan Coast Guard32. Time
series of the labelled land and seafloor GNSS sites are shown in Extended Data Fig. 3. In these and other map view figures in Fig. 4 and Extended Data, plate
interface depth is contoured using dashed lines, and locations of volcanoes from Aster Volcano Archive (NASA, METI; https://gbank.gsj.jp/vsidb/image/)
are indicated using red triangles.

Nature Geoscience | www.nature.com/naturegeoscience


NATurE GEOSCIEnCE Articles

Extended Data Fig. 3 | Observed (colour symbols) and modelled (black curves) time series of GNSS sites after the 2011 Mw 9 Tohoku-oki earthquake.
From left to right: east, north, and vertical components. Site locations are shown in Extended Data Fig. 2. a, Land sites. b, Seafloor sites.

Nature Geoscience | www.nature.com/naturegeoscience


Articles NATurE GEOSCIEnCE

Extended Data Fig. 4 | Map view of observed and modelled postseismic deformation 5 years after the 2010 Mw 8.8 Maule earthquake. a, Horizontal
component. Coseismic slip distribution (Extended Data Fig. 9) is contoured at 4 m interval (coppery lines). Cumulative afterslip, contoured at 0.3 m
interval (grey lines), is modelled using the method of ref. 19. b, Vertical component. The black line shows the location of the corridor in Fig. 3. GNSS data are
from ref. 33. Time series of the labelled GNSS sites are shown in Extended Data Fig. 5.

Nature Geoscience | www.nature.com/naturegeoscience


NATurE GEOSCIEnCE Articles

Extended Data Fig. 5 | Observed (colour symbols) and modelled (black curves) time series of GNSS sites after the 2010 Mw 8.8 Maule earthquake.
From left to right: east, north, and vertical components. Site locations are shown in Extended Data Fig. 4.

Nature Geoscience | www.nature.com/naturegeoscience


Articles NATurE GEOSCIEnCE

Extended Data Fig. 6 | Map view of observed and modelled postseismic deformation 5 years after the 2007 Mw 8.4 Bengkulu earthquake. a, Horizontal
component. Coseismic slip distribution is contoured at 1 m interval (coppery lines; ref. 76). Cumulative afterslip, contoured at 0.4 m interval (grey lines),
is modelled using the method of ref. 19. b, Vertical component. The black line shows the location of the corridor in Fig. 3. GNSS data are from ref. 35. Time
series of the labelled GNSS sites are shown in Extended Data Fig. 7.

Nature Geoscience | www.nature.com/naturegeoscience


NATurE GEOSCIEnCE Articles

Extended Data Fig. 7 | Observed (colour symbols) and modelled (black curves) time series of GNSS sites after the 2007 Mw 8.4 Bengkulu earthquake.
From left to right: east, north, and vertical components. Site locations are shown in Extended Data Fig. 6.

Nature Geoscience | www.nature.com/naturegeoscience


Articles NATurE GEOSCIEnCE

Extended Data Fig. 8 | Summary of postseismic forearc uplift observations following subduction earthquakes. Locations of the earthquakes are shown in
Extended Data Fig. 1.

Nature Geoscience | www.nature.com/naturegeoscience


NATurE GEOSCIEnCE Articles

Extended Data Fig. 9 | Parameters for the postseismic deformation models of the four earthquakes shown in Fig. 3. Map views of the model results for
the Tohoku-oki, Maule, and Bengkulu earthquakes are shown in Extended Data Figs. 2, 4 and 6, respectively. Those for the Chile earthquake are shown in
Fig. 4.

Nature Geoscience | www.nature.com/naturegeoscience


Articles NATurE GEOSCIEnCE

Extended Data Fig. 10 | 2-D Models of postseismic deformation to test the effects of upper plate thickness. The model setup is identical to that in
Fig. 2 (Methods), except for the use of different upper plate thicknesses and/or coseismic slip depths in different simulations. Shown cumulative
postseismic displacements are 5 years after an earthquake, normalized by peak coseismic slip. a, Model tests showing how the upper plate thickness
affects postseismic forearc uplift with or without the elastic cold nose. Minor uplift in the area of interest may occur without the cold nose if the upper
plate in the warm arc and backarc area is unreasonably thick (that is, 50 km). b, Same as a except for deeper coseismic slip. c, Model structure and
parameters. Inset shows coseismic slip distribution along the megathrust for the models in a (Reference) and b (Deeper).

Nature Geoscience | www.nature.com/naturegeoscience

You might also like