ArticlePDF Available

Depth Extent of the Lau Back-Arc Spreading Center and Its Relation to Subduction Processes

Authors:

Abstract

Seismic tomography and wave form inversion revealed that very slow velocity anomalies (5 to 7 percent) beneath the active Lau spreading center extend to 100-kilometer depth and are connected to moderately slow anomalies (2 to 4 percent) in the mantle wedge to 400-kilometer depth. These results indicate that geodynamic systems associated with back-arc spreading are related to deep processes, such as the convective circulation in the mantle wedge and deep dehydration reactions in the subducting slab. The slow regions associated with the Tonga arc and the Lau back arc are separated at shallow levels but merge at depths greater than 100 kilometers, suggesting that slab components of back-arc magmas occur through mixing at these depths.
Physica C 191, 485 (1992).
20. C. J. Muller et al.,Phys. Rev. B 53, 1022 (1996), and
references therein.
21. J. M. Tour et al.,J. Am. Chem. Soc. 117, 9529
(1995).
22. C. Zhou, C. J. Muller, M. R. Deshpande, J. W.
Sleight, M. A. Reed, Appl. Phys. Lett. 67, 1160
(1995).
23. V. Mujica, M. Kemp, A. Roitberg, M. Ratner, J.
Chem. Phys. 104, 7296 (1996).
24. M. P. Samanta et al.,Phys. Rev.53B, R7626 (1996).
25. S. Datta et al.,Phys. Rev. Lett., in press; W. Tian et
al.,Physica E, in press.
26. We thank M. R. Deshpande for fruitful discussions
and the Defense Advanced Research Projects
Agengy for support.
25 April 1997; accepted 22 August 1997
Depth Extent of the Lau Back-Arc Spreading
Center and Its Relation to Subduction Processes
Dapeng Zhao,* Yingbiao Xu, Douglas A. Wiens, LeRoy Dorman,
John Hildebrand, Spahr Webb
Seismic tomography and wave form inversion revealed that very slow velocity anomalies
(5 to 7 percent) beneath the active Lau spreading center extend to 100-kilometer depth
and are connected to moderately slow anomalies (2 to 4 percent) in the mantle wedge
to 400-kilometer depth. These results indicate that geodynamic systems associated with
back-arc spreading are related to deep processes, such as the convective circulation in
the mantle wedge and deep dehydration reactions in the subducting slab. The slow
regions associated with the Tonga arc and the Lau back arc are separated at shallow
levels but merge at depths greater than 100 kilometers, suggesting that slab components
of back-arc magmas occur through mixing at these depths.
Knowledge of the seismic structure be-
neath back-arc spreading centers is impor-
tant because the width and depth of the
slow-velocity regions below spreading cen-
ters provide constraints on the origin of
back-arc spreading (1,2), the geochemical
source of arc and back-arc magmas (3), the
interaction between subduction and back-
arc spreading (1), whether the mantle up-
welling beneath spreading centers is passive
or active, and to what depth the upwelling
persists (2). A subduction zone with an
associated back-arc spreading center and
the existence of deep earthquakes immedi-
ately beneath the center provide an ideal
geometry to image and understand back-arc
spreading processes. The Tonga-Fiji region,
which contains two-thirds of all deep earth-
quakes in the world, represents an optimal
region for exploring these questions. Previ-
ous studies have discussed the seismic ve-
locity anomalies due to the Tonga slab (4,
5), but this work has been hampered by the
poor distribution of seismic stations.
The installation (6) of 12 broadband
stations in the Tonga and Fiji islands from
November 1993 through December 1995
and a related 3-month deployment of 25
ocean bottom seismometers (OBS) (7)in
the Lau back arc and the Tonga forearc
provided a unique opportunity to determine
high-resolution three-dimensional (3D)
structure in this region (Fig. 1A). We used
41,471 arrival times from 926 earthquakes
that occurred in the Tonga-Fiji region dur-
ing the seismic experiment (Fig. 1B). Most
of the events were associated with the sub-
duction of the Tonga slab; they had a rela-
tively uniform distribution in the entire
upper mantle. This uniform distribution is
an advantageous feature over other subduc-
tion zones, such as Japan and Alaska, where
most of the seismicity is concentrated at
depths shallower than 250 km (8,9). We
picked about 8200 arrival times at the 12
land stations from the 926 earthquakes and
about 2900 arrivals at the 25 OBS stations
from 250 earthquakes that occurred during
the OBS deployment. The picking accuracy
is estimated to be 0.05 to 0.3 s. The remain-
ing arrival times were recorded by stations
reporting to the Preliminary Determination
of Epicenters (PDE) with epicentral dis-
tances up to 90°. The PDE arrival times
have lower quality (picking accuracy of 0.2
to 0.5 s), so they were assigned less than
half the weight of the local data. All of the
926 earthquakes were recorded by more
than 20 stations, and their hypocentral lo-
cations have a statistical accuracy of 63to
9 km. We also picked 450 arrival times at
the 12 land and 25 OBS stations from 45
large (magnitude of 6.0 to 8.0) teleseismic
events with epicentral distances from 30° to
90°, which were assigned the same weight
as the local data in the inversion.
We used a tomography method (9)to
determine the 3D Pwave velocity structure
in the Tonga-Fiji region (9,10) (Figs. 2 and
3). To confirm that the major velocity fea-
tures were adequately resolved by the inver-
sion, we conducted checkerboard resolution
tests (11) (Fig. 4). The checkerboard test
with a grid spacing of 50 km indicates good
resolution for the area in and around the
subducting Tonga slab and along the main
line of OBSs (Fig. 4, A and B). For the test
with a grid spacing of 70 km, the resolution
is good for all the areas discussed (Fig. 4, C
and D). We also conducted a number of
inversions and resolution tests by changing
the grid spacing, the grid configuration, and
the initial model (10). The results show
that the velocity structure in the study area
(Fig. 3) can be resolved with a resolution of
50 to 70 km. This resolution scale is better
than the 100-to 200-km resolution ob-
D. Zhao, Southern California Earthquake Center and De-
partment of Earth Sciences, University of Southern Cali-
fornia, Los Angeles, CA 90089, USA.
Y. Xu and D. A. Wiens, Department of Earth and Plane-
tary Sciences, Washington University, St. Louis, MO
63130, USA.
L. Dorman, J. Hildebrand, S. Webb, Scripps Institution of
Oceanography, University of California, San Diego, La
Jolla, CA 92093, USA.
*To whom correspondence should be addressed. E-mail:
dzhao@usc.edu
Fig. 1. (A) Map showing the seismometer deploy-
ments in the Fiji-Tonga region. Twelve broadband
instrument island sites, 25 OBS sites, and two
PDE sites (at Samoa and Fiji) recorded the data
used in this study. A 2-year sample of deep earth-
quakes (depths of 300 to 680 km and m
b
.4.8)
(dots) delineates the deep Tonga slab. PASSCAL,
Program for Array Seismic Studies of the Conti-
nental Lithosphere; IRIS, Incorporated Research
Institutions in Seismology; GSN, Global Seismo-
graphic Network. (B) Hypocentral distribution of
the 926 earthquakes used in this study.
SCIENCE zVOL. 278 z10 OCTOBER 1997 zwww.sciencemag.org254
tained in previous studies (5).
The subducting Tonga slab was imaged
as a 100-km-thick zone with a Pwave
velocity that is 4 to 6% higher than the
surrounding mantle (Fig. 2). Beneath the
Tonga arc and the Lau back arc, low-
velocity anomalies of up to 7% are visible
(Figs. 2 and 3). The slow-velocity anomaly
beneath the Tonga arc represents a dip-
ping region about 30 to 50 km above the
slab, extending from the surface to about
140-km depth (Fig. 2). This feature is
similar to the low-velocity features found
beneath the Japan and Alaska volcanic
fronts (8,9). This slow anomaly probably
represents the source zone for island arc
magmas. Volatiles released from the sub-
ducting slab may reduce the melting point
of the rock above and allow partial melt-
ing to produce arc magmas (3,12). Slow
anomalies beneath areas of the active
Central Lau Spreading Center (CLSC)
and the Eastern Lau Spreading Center
(ELSC) extend to depths of about 100 km.
These depths correspond to regions where
the primary magma genesis is expected to
take place beneath an oceanic spreading
center (13,14). The maximum heteroge-
neity of Pwave velocity between the Lau
back-arc basin and the Pacific Plate is
about 13% at these depths. Slow anoma-
lies are located to the west of the CLSC
and ELSC (Figs. 2 and 3). Beneath 100-
km depth, the amplitude of the back-arc
anomalies is reduced, but a moderately
slow anomaly (22to24%) exists down to
a depth of at least 400 km. To investigate
the depth extent of slow anomalies in the
Lau back arc with a different methodolo-
gy, we inverted 16 wave forms from seven
regional earthquakes recorded at the land
stations to determine the 1D Swave struc-
ture beneath the Lau Basin (15). The
inversion results (Fig. 5) show a similar
level of velocity heterogeneity and depth
distribution of the back-arc anomalies to
that found in the Pwave tomography. The
level of Swave velocity heterogeneity
reaches a maximum of about 18% between
the Lau Basin and the old Pacific litho-
sphere at depths of 40 to 90 km (Fig. 5).
The velocity difference decreases to about
2% at 180-km depth, but a small, poorly
resolved difference persists to greater
depths (15). There has been disagreement
concerning the depth extent of slow-ve-
locity anomalies at mid-ocean ridges
(MORs) (13,16). Our results show that,
at least for back-arc spreading centers,
moderately slow velocity anomalies ex-
tend to depths of at least 400 km. These
anomalies may reflect either the depth
extent of oceanic spreading centers due to
the depth of the associated upwelling pat-
terns or processes endemic to back-arc
spreading centers, perhaps due to interac-
tions between the slab and the back arc.
The slow-velocity anomalies at depths
of 300 to 400 km (Fig. 2) could be caused
by upwelling flow patterns in the back-arc
region or by volatiles resulting from the
deep dehydration reactions occurring in
the subducting Tonga slab. Volatiles
would have the effect of lowering the
melting temperature and the seismic ve-
locity and may produce small amounts of
partial melt (17). Temperatures in fast
subducting slabs like Tonga are low
enough for water to reach the stability
depths of dense hydrous magnesian silicate
phases (18), which may allow water pen-
etration down to depths of 660 km (18,
19). The phase diagrams of important hy-
drous phases, the associated reaction ki-
netics, and the relevant mantle conditions
(slab temperature and composition) are
not known sufficiently well enough to pre-
dict the depth at which dehydration would
occur. Partial melting of the back-arc re-
gion by volatiles from the deep slab may
be important in localizing low seismic ve-
locities in the back arc; the slow anomalies
we observed at depths of 300 to 400 km
may represent this process.
The slowest anomaly in the back-arc
region is not found beneath the spreading
center, but rather to the west. This is sim-
0
100
200
300
400
500
600
700
Depth (km)
-6% -3% 0% 6%3%
Fiji CLSC ELSC
Lau
ridge Tonga arc
volcanoes
Fig. 2. East-west vertical
cross section of a Pwave ve-
locity image from 0- to 700-
km depth along the line AB
(1220-km length) in Fig. 3A.
Red and blue colors denote
slow and fast velocities, re-
spectively. Solid triangles de-
note active volcanoes. CLSC
denotes the location of the
Central Lau Spreading Center
and ELSC denotes the loca-
tion of the Eastern Lau
Spreading Center. Earth-
quakes within a 40-km width
from the cross section are
shown as white circles. The
velocity perturbation scale is
shown at the bottom.
177˚E 179˚W 175˚W 171˚W
A
C
EF
B
A
B
D
177˚E 179˚W 175˚W 171˚W
15˚S
19˚S
23˚S
15˚S
19˚S
23˚S
15˚S
19˚S
23˚S
Fig. 3. Pwave velocity
images at (A) 25-, (B) 60-,
(C) 100-, (D) 140-, (E)
230-, and (F) 430-km
depths. Earthquakes
within a 20-km depth
range of the slice are
shown as white circles.
The long contour line to
the right shows the Tonga
trench. The lines in the
middle show the back-arc
spreading centers. The
short contour lines to the
left and in the upper right-
hand corner show islands.
Line AB in Fig. 3A shows
the location of the cross
section in Fig. 2. All other
labeling is the same as in
Fig. 2.
REPORTS
www.sciencemag.org zSCIENCE zVOL. 278 z10 OCTOBER 1997 255
ilar to observations from several recent ex-
periments along MORs, which showed
smaller delay times for arrivals at stations
near the MORs than for stations on the
flanks (20,21). The faster arrivals near the
ridge axis have been attributed to the align-
ment of anisotropic minerals in the mantle,
with fast propagation for vertically traveling
Pwaves caused by focused vertical flow
beneath the spreading center. This effect
may cause the arrivals at OBS stations near
the spreading center to be faster than those
off the ridge, causing the slowest anomalies
to be displaced off the spreading center.
The actual magma chamber beneath the
spreading center would be expected to be
less than 10 km in width (22), too small to
image in this study.
Anisotropy may explain why the slow-
est velocities are not found immediately
beneath the ridge, but it cannot explain
why the western flank of the spreading
center is slower than the eastern flank.
This observation may be related to the
ongoing tectonic processes in the Lau
Basin. The CLSC, toward the west, is
lengthening southward by rift propagation
at the expense of the older ELSC, trans-
ferring the spreading activity westward
(23). Our results suggest that this transfer
may be favored by the proximity of the
western ridge to upper mantle with the
slowest velocities, which is also presum-
ably the hottest mantle and the best
source region for magma. Thus, the ridge
propagation may be an attempt by the
tectonic system to maintain the spreading
center at or near the upper mantle magma
source region.
The slow-velocity regions beneath the
Tonga arc and the Lau back arc seem to be
separated at shallow levels but merge at
deeper levels (compare Fig. 3, A and C).
This behavior suggests that although the
arc and back-arc magma systems are sepa-
rated at shallow levels, where most of the
magma is generated, there may be some
interchange between the magma systems at
depths greater than 100 km. Interchange
with slab-derived volatiles at depths greater
than 100 km may help to explain some of
the unique features in the petrology of
back-arc magmas relative to typical MOR
basalts, including excess volatiles and large
ion lithophile enrichment (24).
REFERENCES AND NOTES
___________________________
1. S. Uyeda, in Island Arcs, Deep Sea Trenches and
Back-Arc Basins, M. Talwani and W. C. Pitman III,
Eds. (American Geophysical Union, Washington,
DC, 1977), pp. 1–14; M. N. Toksoz and P. Bird, in
Island Arcs, Deep Sea Trenches and Back-Arc Ba-
sins, M. Talwani and W. C. Pitman III, Eds. (American
Geophysical Union, Washington, DC, 1977), pp.
379–393; N. H. Sleep and M. N. Toksoz, Nature 33,
548 (1971).
2. D. W. Forsyth, in Mantle Flow and Melt Generation at
Mid-Ocean Ridges, J. Morgan, D. Blackman, J. Sin-
ton, Eds. (American Geophysical Union, Washing-
ton, DC, 1993), pp. 1–65; D. L. Turcotte and J. P.
Morgan, in ibid., pp. 155–182.
3. S. M. Peacock, Science 248, 329 (1990).
4. L. N. Huppert and C. Frohlich, J. Geophys. Res. 86,
3771 (1981); G. Bock, ibid. 92, 13863 (1987); K. M.
Fischer, K. C. Creager, J. H. Jordan, ibid. 96, 14403
(1991).
5. H. Zhou, Phys. Earth Planet. Inter. 61, 199 (1990); J.
Geophys. Res. 101, 27791 (1996); R. van der Hilst,
Nature 374, 154 (1995).
6. D. Wiens et al., IRIS Newsl. 14, 1 (1995).
7. R. Jacobson, L. Dorman, G. Purdy, A. Schultz, S.
Solomon, Eos 72, 506 (1991).
8. D. Zhao, D. Christensen, H. Pulpan, J. Geophys.
Res. 100, 6487 (1995).
9. D. Zhao, A. Hasegawa, S. Horiuchi, ibid.. 97,
19909 (1992); D. Zhao, A. Hasegawa, H. Kana-
mori, ibid. 99, 22313 (1994). A 3D net of nodes was
set up in the Tonga-Fiji region with a horizontal grid
spacing of 50 km and a vertical spacing of 25 to 50
km. Hypocentral locations of earthquakes and ve-
locities at the grid nodes were taken as unknown
parameters. The velocity at any point was calculat-
ed by linearly interpolating the velocities at the eight
grid nodes surrounding that point. An efficient 3D
ray tracing technique was used to calculate ray
paths and travel times. A conjugate gradient algo-
rithm [C. Paige and M. Saunders, ACM Trans.
Math. Software 8, 43 (1982)] was used to solve the
large sparse system of observational equations by
regularizing the solution in a damped least-squares
fashion [K. Aki and W. Lee, J. Geophys. Res. 81,
4381 (1976)]. The nonlinear tomographic problem
was solved by iteratively conducting linear inver-
sions. Hypocenters were relocated in the inversion
process. For the local data from the 926 events
(Fig. 1), raw arrival times were used in the inversion.
For the teleseismic data with epicentral distances
greater than 30°, we used the relative travel time
residuals by removing the average of the travel time
residuals for each station, to remove the effect of
the structures outside the study area.
10. D. Zhao et al.,Eos (Fall Suppl.) 77, 498 (1996). The
starting 1D model for the inversion was the IASP91
Earth model (25), but the crustal thickness has
lateral variations according to a seismic refraction
survey along the line of OBS stations (26) and the
inversion of land station wave forms (15). The crust-
al thickness varies from 5 to 24 km, with the thick-
est crust beneath the Lau Ridge, Tonga arc, and Fiji
Islands. Numerous previous studies (4,5) have
demonstrated the existence of the Tonga slab,
which is at least 100 km thick with a Pwave velocity
of up to 7% faster. The slab substantially perturbs
ray trajectories, such that the ray path perturbation
can be over 100 km from that in the 1D Earth model
(4,9). Our primary concern in this study is the
precise structure below the Lau Basin and in the
mantle wedge. For this purpose, it was necessary
to use a starting model incorporating the a priori
slab information (8,9) because we hoped to resolve
the anomalies in the Lau Basin with a resolution of
50 to 70 km. We introduced the Tonga slab into the
model using the slab geometry determined by Bill-
ington (27). The introduction of the slab into the
model reduced the nonlinearity of the problem be-
cause seismic rays are traced realistically from the
beginning of the inversion (8,9). We conducted a
number of inversions by changing the initial slab
thickness from 50 to 150 km and the initial slab
velocity from 1 to 9% faster than the normal mantle.
We also conducted inversions without the slab in
the starting model. All of the inversions imaged the
very slow velocity anomalies beneath the Lau
spreading center to about 100-km depth and the
moderately slow anomalies to about 400-km
depth, but there are 1 to 2.5% changes in the
amplitudes of the anomalies and up to 10% chang-
es in the final travel time residuals. The inversion
with the starting model that included a 100-km-
thick slab and that was 6% faster than the normal
mantle resulted in the best fit of model to data. The
results of this inversion are shown in Figs. 2 and 3.
11. To make a checkerboard in the resolution tests, we
assigned positive and negative velocity anomalies
Fig. 4. Results of checker-
board resolution tests for P
wave velocity structure at
25-km (Aand C) and 480-
km depths (Band D). The
grid spacing is 50 km in (A)
and (B) and 70 km in (C) and
(D). Open and solid circles
denote low and high veloci-
ties, respectively. The per-
turbation scales are shown
at the bottom.
Fig. 5. Swave velocity models determined by
inversions of entire regional vertical and radial
wave forms recorded at island broadband seismic
stations for various tectonic regions of the south-
west Pacific.
SCIENCE zVOL. 278 z10 OCTOBER 1997 zwww.sciencemag.org256
with magnitudes of 5% to the 3D grid nodes. Syn-
thetic data were calculated for the checkerboard
model. Then we added random errors to the synthet-
ic data and inverted them with the same algorithm
that we used for the observed data. The inverted
image of the checkerboard suggests where the res-
olution is good and where it is poor. The checker-
board resolution tests and other synthetic tests we
conducted showed that both the high-velocity
Tonga slab and the low-velocity back arc and mantle
wedge were reliably resolved and that there was no
trade-off between them.
12. Y. Tatsumi, J. Geophys. Res. 94, 4697 (1989); J. H.
Davies and D. J. Stevenson, ibid. 97, 2037 (1992).
13. Y. Zhang and T. Tanimoto, Nature 355, 45 (1992); T.
Tanimoto and D. J. Stevenson, J. Geophys. Res. 99,
4549 (1994).
14. Y. Shen and D. W. Forsyth, J. Geophys. Res. 100,
2211 (1995).
15. Y. Xu and D. Wiens, ibid., in press. To invert region-
al wave forms we used a nonlinear inversion meth-
od that adopts a reflectivity formalism (28) to com-
pute the partial derivatives. This method allows the
entire regional distance (400- to 1500-km range)
wave form to be inverted from Pwave arrival to
surface waves at frequencies between 0.01 and
0.055 Hz. Broadband seismograms from earth-
quakes of 10 to 240 km deep that propagate al-
most entirely within one of the tectonic regions of
the southwest Pacific were used in the wave form
inversion. Parameter variances and resolution tests
suggest that the results are well constrained to
depths of about 200 km. The Swave velocities
from the wave form inversion and the Pwave ve-
locities from the tomography would not necessarily
show the same structure. The larger total hetero-
geneity from the wave form inversion may result
from a greater effect of partial melt beneath the Lau
back arc on Swave velocity than on Pwave veloc-
ity (9,29).
16. W. Su, R. L. Woodward, A. M. Dziewonski, Nature
360, 149 (1992).
17. G. Nolet and A. Zielhuis, J. Geophys. Res. 99, 15813
(1994); G. Nolet, in Processes of Deep Earth and
Planetary Volatiles, K. Farley, Ed. (American Institute
of Physics, New York, 1995), pp. 22–32.
18. A. B. Thompson, Nature 358, 295 (1992); H. Stau-
digel and S. D. King, Earth Planet. Sci. Lett. 109, 517
(1992).
19. A. Navrotsky and K. Bose, in Processes of Deep
Earth and Planetary Volatiles, K. Farley, Ed. (Ameri-
can Institute of Physics, New York, 1995), pp. 221–
228.
20. D. Toomey et al.,Eos (Fall Suppl.) 77, 652 (1996).
21. D. Blackman et al.,Geophys. J. Int. 127, 415 (1996).
22. J. S. Collier and M. C. Sinha, J. Geophys. Res. 97,
14031 (1992).
23. L. M. Parson and I. C. Wright, Tectonophysics 263,1
(1996); B. Taylor, K. Zellmer, F. Martinez, A.
Goodliffe, Earth Planet. Sci. Lett. 144, 35 (1996).
24. J. A. Pearce et al.,inVolcanism Associated with
Extension at Consuming Plate Margins, J. L. Smellie,
Ed. (Geological Society, London, 1992), pp. 53–75;
J. W. Hawkins, in Active Margins and Marginal Ba-
sins of the Western Pacific, B. Taylor and J. Natland,
Eds. (American Geophysical Union, Washington,
DC, 1995), pp. 125–173.
25. B. L. N. Kennett and E. R. Engdahl, Geophys. J. Int.
105, 429 (1991).
26. W. Crawford, S. Webb, J. Hildebrand, Eos (Fall
Suppl.) 77, 478 (1996).
27. S. Billington, thesis, Cornell University, Ithaca, NY,
(1980).
28. B. L. N. Kennett, Seismic Wave Propagation in Strat-
ified Media (Cambridge Univ. Press, Cambridge,
1983); G. E. Randall, Geophys. J. Int. 118, 245
(1994).
29. U. H. Faul, D. R. Toomey, H. S. Waff, Geophys. Res.
Lett. 21, 29 (1994).
30. We thank M. Bevis, W. Crawford, K. Draunidalo, S.
Escher, T. Fatai, H. Gilbert, S. Helu, K. Koper, M.
McDonald, J. McGuire, B. Park-Li, G. Prasad, E.
Roth, A. Sauter, P. Shore, and L. Vuetibau for their
assistance during the seismic experiment and at the
data-processing stage and G. Abers and an anony-
mous referee for thoughtful reviews, which improved
the manuscript. Broadband seismographs were ob-
tained from the PASSCAL program of the Incorpo-
rated Research Institutions in Seismology (IRIS).
Supported by the NSF under grants EAR-9219675,
OCE-9314446, and EAR-9614502. This paper is
Southern California Earthquake Center publication
386.
8 May 1997; accepted 18 August 1997
Microscopic Molecular Diffusion Enhanced by
Adsorbate Interactions
B. G. Briner,* M. Doering, H.-P. Rust, A. M. Bradshaw
The diffusion of carbon monoxide molecules on the (110) surface of copper was inves-
tigated in the temperature range between 42 and 53 kelvin. The activation energy for
thermal motion was determined directly by imaging individual molecular displacements
with a scanning tunneling microscope. An attractive interaction between carbon mon-
oxide molecules gave rise to the formation of dimers and longer chains. Carbon mon-
oxide chains diffused substantially faster than isolated molecules although the chains
moved by a sequence of single-molecule jumps. A higher preexponential factor in the
Arrhenius law was found to be responsible for the observed efficiency of chain hopping.
Adsorbate diffusion is of fundamental im-
portance for surface chemistry (1). It is
often the rate-limiting step in catalysis be-
cause adsorbed atoms or molecules first
have to reach a reaction partner or an ac-
tive site (2) on the surface before a reaction
can take place. Efforts to study diffusion on
a microscopic scale are needed to under-
stand how interactions with the surface and
with neighboring adsorbates influence the
way a particle diffuses. This information
forms an indispensable basis to model diffu-
sion on a macroscopic scale under the con-
ditions that prevail in catalysis. This report
focuses on the microscopic diffusion of CO
molecules on Cu(110). Carbon monoxide is
only weakly chemisorbed on Cu(110) (3),
and helium scattering experiments have
suggested very low diffusion barriers (4).
Therefore, CO can serve as a test case to
assess whether diffusion of the often weakly
bound molecules that are of interest in sur-
face chemistry is accessible to microscopic
observation.
All experiments were performed with an
Eigler-type, variable temperature scanning
tunneling microscope (STM), which oper-
ates in an ultrahigh vacuum and can be
cooled down to4K(5). We applied exper-
imental techniques similar to those used in
earlier STM-based studies on diffusion (6),
but the present results differ in two ways
from earlier findings. First, we observed that
the activation energy for CO diffusion was
substantially lower than the barrier heights
that had been determined before with mi-
croscopic imaging techniques. This result in-
dicates that the STM can indeed be used to
probe the motion of weakly bound species
and that the artifacts of STM-induced adsor-
bate motion that were reported in (7) can be
avoided. Second, our study went beyond the
observation of single-particle diffusion. It
was found that CO forms chains on
Cu(110). These chains experienced consid-
erable thermal mobility in the same temper-
ature range in which the diffusion of isolated
molecules was observed. By comparing the
diffusion of single molecules with that of CO
chains, we could investigate the influence of
molecular interactions on the adsorbate mo-
bility. Cluster diffusion was first investigated
by field ion microscopy (FIM) (8). Although
this technique is limited to the study of
strongly bound transition metal adatoms, it
could provide detailed information on the
characteristics of cluster diffusion. In gener-
al, the rule that cluster mobility decreases
strongly with increasing cluster size was con-
firmed, but FIM experiments have also dem-
onstrated that there are exceptions to this
rule. Iridium tetramers have been found to
diffuse faster than trimers (9), and for rheni-
um on tungsten(211), dimers have been
shown to be faster than single adatoms (10).
The reason for this enhanced mobility is a
reduction of the activation energy; adding an
atom to a cluster can strengthen the cluster
bonds at the expense of weakening the
bonds to the substrate (11–13). We observed
that CO chains also experienced an en-
hanced mobility, but in contrast to the metal
clusters described above, no reduced activa-
tion energy for chain diffusion was found.
Samples were prepared by adsorbing CO
onto the clean Cu(110) substrate at a tem-
perature of about 60 K. We found that under
these adsorption conditions, CO still has
substantial mobility. This mobility is inferred
Fritz-Haber-Institut der Max-Planck-Gesellschaft, Fara-
dayweg 4-6, 14195 Berlin, Germany.
*To whom correspondence should be addressed. E-mail:
briner@fhi-berlin.mpg.de
REPORTS
www.sciencemag.org zSCIENCE zVOL. 278 z10 OCTOBER 1997 257
... HFZ = Hunter fracture zone; LETZ = Lau extensional transform zone; CLSC = Central Lau Spreading Center; ELSC = Eastern Lau Spreading Center; VFR = Valu Fa Ridge. (For interpretation of the colors in the figure(s), the reader is referred to the web version of this article.) of the Tonga-Lau-Fiji region obtained by different methods and from several seismic experiments (Conder and Wiens, 2006;Roth et al., 1999;Wei et al., 2016Wei et al., , 2015Wei and Wiens, 2018;Zha et al., 2014;Zhao et al., 1997), however, either their low resolutions (>50 km) or their shallow depth ranges (<200 km depth) restrict our understanding of the deep structure beneath the whole area. A detailed 3-D model of depth-varying anisotropy is needed to investigate the 3-D mantle flow field at depth. ...
... /ds /nodes /dmc/). We also used the arrival-time data of 926 local earthquakes recorded during 1993-1995 in the study region (Zhao et al., 1997). As a result, our data set includes 23,781 P-arrivals and 5878 S-arrivals from a total of 1955 local earthquakes (Table S1). ...
... Large-scale low-velocity (low-V) anomalies (Figs. 4 and 5) are clearly revealed beneath the spreading center of the Lau Basin and volcanic arc in the mantle wedge. The low-V anomalies extend to ∼400 km depth (Figs. 4 and 5), which is similar to the previous tomographic result (Zhao et al., 1997). This low-V zone is slightly broader beneath the CLSC and ELSC (200-400 km wide) than that beneath the VFR (∼150 km wide) (Figs. 4 and 5). ...
Article
Full-text available
The Tonga-Lau-Fiji region is important to study plate-plume and subduction-ridge interactions, but its deep mantle structure is still not very clear. Here we present high-resolution tomography of 3-D P-wave azimuthal anisotropy down to 400 km depth of the Tonga subduction zone derived from arrival-time data of local earthquakes recorded at seafloor and land seismometers. The subducting Tong slab is imaged as high-velocity anomalies at depths of 100-400 km, whereas large-scale low-velocity anomalies down to 400 km depth are revealed in the mantle wedge beneath the backarc basin and volcanic arc. Trench-parallel anisotropy beneath the Lau Basin extends southwards to ∼140 km depth at ∼20.5◦S, representing the extent of both southward flow of the Samoan plume and toroidal flow by the slab rollback. At depths of 140-400 km, the Lau Basin and Fiji Plateau mainly exhibit plate-parallel fast-velocity directions (FVDs) north of ∼20.5◦S, indicating strong corner flow in the mantle wedge driven by the slab subduction and dehydration. The Tonga slab exhibits trench-parallel FVDs at depths of <200 km, reflecting fossil fabric formed during the plate spreading stage, whereas, at greater depths, the slab mainly exhibits trench-normal FVDs, which may reflect complicated deformations within the slab. These results suggest that the Samoan plume has a significant impact on the Tonga-Lau-Fiji region, leading to variations in the scale and depth extent of mantle flows.
... The seismic tomography method of Zhao et al., (1992Zhao et al., ( , 1994) is used in this study to determine the detailed velocity Poisson's ratio structures beneath the Red Sea rift zone and its flanks. This method has been used worldwide with different aspects (e.g., Zhao and Kanamori, 1995;Zhao et al., 1996Zhao et al., , 1997Zhao et al., , 2001Serrano et al., 2002;Kayal et al., 2002, Salah et al., 2007Abdelwahed and Zhao, 2007;Ş ahin et al., 2019;Abdelwahed et al., 2016Abdelwahed et al., , 2023. The method adopts the 3-D initial model, the modified pseudo-bending 3-D ray tracing technique (Zhao et al., 1992), and the conjugate gradients algorithm (Paige and Saunders, 1982) for travel time inversion. ...
Article
The role of the Red Sea rift on the development and the discrepancy of the seismo-volcanic activity along its flanks is still a debate. Here we tried to resolve this debate by a high-resolution tomographic imaging of the northern Red Sea subsurface structures. For the first time, a large number of arrival time data from the Saudi and Egyptian seismic networks were used. The area comprises the Lunayyir volcanic field in the Saudi Arabia, the Abu-Dabbab seismogenic zone in the Egyptian Red Sea coast, and the Zabargad Shear Zone in the Red Sea. This study revealed clear images of the Red Sea rift-related structures along its flanks. The subsurface extension of the different shear and suture zones existing in the Arabian Nubian shield are well imaged. It is found that the Lunayyir seismo-volcanic activity is possibly controlled by the reactivation of the Yanbu suture zone that is associated with steeply northwestward dipping structure. The suture detachments were observed and identified across the Red Sea as low-V channels. The Egyptian Eastern Desert is found to be highly deformed with crustal-scaled low-V structures which were inherited from the early period of Gondwana collision. The NE-SW strike slip faults along the Red Sea were found to be a part of this deformation that are extended deeply in the crust. The Abu-Dabbab and Marsa-Alam areas were strongly influenced by this deformation with possible magma intrusions. This study provides new insights on the role of the Red Sea in the seismo-volcanic activity along its flanks.
... Low-velocity anomalies in the mantle wedge of the subduction zone have been confirmed by seismological tomography (Abers, 2005;Zhang et al., 2004;Zhao et al., 1997Zhao et al., , 2001. Previous studies have generally attributed these anomalies to the presence of melt/fluid (Nakajima et al., 2005), anomalous temperature (Tamura et al., 2002), or serpentinization (Cai et al., 2018;Kawakatsu & Watada, 2007;Wang et al., 2019Wang et al., , 2022, among which serpentinization may be the main cause. ...
Article
Full-text available
Talc is widely distributed in metamorphic and hydrothermally altered rocks, and its velocity is useful for constraining the velocity of mantle wedge. In this study, we used ultrasonic interferometry to study the elastic properties of talc polycrystalline samples at high pressure (up to 5.6 GPa). VP increases monotonically with increasing pressure, and VS increases with pressure up to ∼3.2 GPa, but negative pressure dependence occurs above ∼3.2 GPa. Talc has a VP/VS ratio of 1.63–2.14 and Poisson's ratio of 0.20–0.36 at the pressure of 1.2–5.6 GPa, which can be used to explain some anomalous seismic observations in subduction zone. Combining the elastic properties of talc with those of olivine, orthopyroxene, and antigorite, the wave velocity of mantle wedge was calculated as a function of water content. The results suggest that lower water content is needed to match the observed values of seismic tomography when talc is present, compared with the antigorite scenario.
... km/s have been observed by tomography. Seismologists attribute the high-velocity anomalies in the oceanic crust to the formation of eclogite in hydrated oceanic crust during the subduction process (Kita et al., 2010;Nakajima et al., 2001Nakajima et al., , 2009aNakajima et al., , 2009bNakajima et al., , 2011Tsuji et al., 2008;Zhao et al., 1997). The velocities of eclogite under high-temperature and high-pressure conditions based on the results for Omp80Grt20 and Omp50Grt50 obtained in this study are V P = 8.3-8.7 km/s and V S = 4.6-4.8 ...
Article
The increased velocity associated with decaying seismicity in the subducting oceanic crust usually has been attributed to its eclogitization. However, the degree of the eclogitization of subducting oceanic crust at depth remains unclear due to the lack of velocity data for eclogite at pressures of >3 GPa. In this study, the P- and S-wave velocities of eclogite aggregates composed of omphacite and garnet (Omp100, Omp80Grt20, and Omp50Grt50) were measured simultaneously at pressures of up to 8 GPa using ultrasonic interferometry. The velocities of the eclogite aggregates increased with rising pressure and garnet content. By determining the velocities at different pressures using the finite strain equation, the elastic moduli of the eclogites and their pressure derivatives were determined to be KS0 = 119−134 GPa, KS0′ = 5.2−5.3, G0 = 73−80 GPa, and G0′ = 1.5−1.6. The high-wave velocity and low VP/VS ratio of the eclogites, combined with the seismic tomography and seismicity distribution, jointly constrain the depth and composition of eclogitization in the subducted oceanic crust of Northeast Japan, and a new 1-D velocity structure is proposed. We also compiled and calculated the depth of the eclogitization and the garnet content of the subducting oceanic crusts in 27 typical locations around the Pacific Ocean. The depth of the eclogitization was positively correlated with the age of the subduction zone, and the garnet content was estimated to be 12%−32% in our model.
... Seismic imaging provides critical information on the structure of subduction zones. While seismic velocities images based on travel times are used routinely to constrain the geometry and structure of subducting plates and mantle wedges (e.g., Cai et al., 2018;Guo et al., 2021;van der Hilst et al., 1997;Zhao et al., 1997), tomography models show discrepancies in part due to differences in inversion methodology or parameters. In principle, seis mic amplitudes should provide a powerful complementary data set that can enhance travel-time images, often revealing sharper features that are blurred by tomography regularization. ...
Article
Full-text available
Seismic wave amplitudes have tremendous sensitivity to subduction structure; however, they are affected by attenuation, scattering and focusing, and have therefore been sparsely used compared with traveltimes. We measure and model teleseismic body wave amplitudes recorded at a dense broadband array in the Washington Cascades. These data show anomalous amplitude variations with complex azimuthal dependence at frequencies as low as 0.05 Hz, accompanied by significant multipathing. We demonstrate using spectral‐element numerical simulations that focusing of the teleseismic wavefield by the Juan de Fuca slab is responsible for some of the amplitude anomalies. The focusing effects can contaminate the apparent differential attenuation measurements and produce at least 20% of the inferred attenuation signal. Our results indicate that the amplitudes are sensitive to the subducting slab geometry and subduction structure, and can be used to refine seismic images. Ubiquitous and consistent amplitude anomalies are observed along the arc, suggesting that the Juan de Fuca slab may be continuous from Canada to northern California.
... These values are deduced from seismic velocity profiles converted using a water velocity of 1,480 m/s (Hayes and Lewis, 1985;Armada et al., 2020). While the observed magmatic growth rates showed concurrence with the age of the subducting slabs, this work does not discount the role of other factors which affect the composition, structure, and thickness of different arc crusts, such as crustal anatexis (Izu-Bonin-Marianas Arc- Tatsumi et al., 2008); delamination processes (Kuril Arc- Tsumura et al., 1999); degree of back-arc spreading (Tonga Arc- Zhao et al., 1997), and the level of maturity in crustal evolution (Aleutian Arc- Holbrook et al., 1999). This work presents the latest estimates of the magmatic productivity of oceanic island arcs. ...
Article
Full-text available
Island arcs are postulated as the juvenile components that contribute to the growth of continental crust. Growth rates of arc crusts were previously computed using crustal thicknesses derived from seismic data. Consequently, crustal growth rates of oceanic island arcs are also constrained by the limited seismic data availability. This work presents the first comparison of gravity-derived magmatic growth rates of Western Pacific oceanic island arcs. We used the statistical correlation between Bouguer anomalies and seismic-derived crustal thicknesses to generate an empirical formula. The new empirical formula was utilized to estimate the crustal thicknesses of oceanic island arcs using Bouguer anomalies from the EGM2008 global gravity model. The resulting crustal thicknesses were employed to compute the magmatic growth rates of western Pacific island arcs and the Philippine island arc system. The latest magmatic growth rate estimates show that the magmatic productivity of Western Pacific island arcs, which are directly associated with Pacific Plate subduction, is significantly higher (28–60 km³/km/m.y). The growth rate of the Pacific island arcs is higher compared to the magmatic growth rate computed for the other oceanic island arcs (12–25 km³/km/m.y), which are derived from the subduction of other oceanic lithospheres (i.e., the Philippine Sea Plate; Caribbean Sea Plate; and Eurasia-South China Sea slabs). This is attributed to the variation in the ages of the subducting plates. The Pacific Plate, being older, is associated with higher degrees of serpentinization and sediment cover, which introduce more volatiles inducing more robust partial melting of the mantle wedge.
... In this study, the seismic tomography method of Zhao et al. (1992Zhao et al. ( , 1994) is used to determine the detailed 3-D Vp, Vs, and Poisson's ratio images beneath Harrat Lunayyir and the adjacent Red Sea offshore zone. This method has been used in different tectonic settings (e.g., Zhao and Kanamori, 1995;Zhao et al., 1996Zhao et al., , 1997Zhao et al., , 2001Serrano et al., 1998Serrano et al., , 2002Kayal et al., 2002, Salah et al., 2007Abdelwahed and Zhao, 2007;Abdelwahed et al., 2016;Ş ahin et al., 2019). It is capable of dealing with the complex geometry of the Moho and Conrad discontinuities. ...
Article
Harrat Lunayyir is one of the youngest Cenozoic volcanic fields in western Saudi Arabia, which in 2009 experienced a seismic swarm of about 34,000 events with a maximum magnitude of M5.4. Many studies debated about the relationship of a depicted shallow-level magma intrusion with the Red Sea rifting and the channelized northward flow from the Afar mantle plume. In this study, a detailed seismic tomography inversion for the P- and S-wave velocities, and the Poisson’s ratio structures beneath Harrat Lunayyir and the adjacent Red Sea offshore area is presented. The tomographic images showed evidence of magma intrusions beneath the Lunayyir Swarm Area (LSA) characterized by low P- and S-wave velocity and high Poisson’s ratio anomalies down to 21 km depth. These anomalies are in juxtaposition with some high-velocity anomalies attributed to remnants of solidified magma intrusions relevant to previous episodes of volcanicity. Beneath the Red Sea, the lithospheric mantle structure is clearly observed as a high P-wave velocity high-Poisson’s ratio anomaly at a depth of 20-40 km. Traces of high-Vp anomalies with anomalous Poisson’s ratio are observed in the area between the Red Sea and the LSA suggesting that a channel of hot lithospheric mantle material may link the Red Sea rift structure with the LSA. We suggest that the Red Sea drifting and its related processes in the Zabargad Shear Zone might have significantly contributed to the early formation of the Lunayyir volcanic field and could possibly have in an impact on any magmatic activity in the future.
... By imaging seismic velocities, local earthquake travel-time tomography (LET) is a geophysical tool that can shed light on slab hydration, dehydration, and mantle wedge dynamics. Previous LET studies have primarily centred on fast Pacific subduction zones, such as in Tonga (Zhao et al., 1997), the southern Hikurangi (Eberhart-Phillips et al., 2005), Alaska (Rossi et al., 2006), central America (Syracuse et al., 2008), South America (Hicks et al., 2014), and Mariana (Barklage et al., 2015). They have successfully imaged key domains, such as the subducting slab and mantle wedge, allowing inferences about volatile recycling at various depths. ...
Article
Full-text available
Volatiles play a pivotal role in subduction zone evolution, yet their pathways remain poorly constrained. Studying the Lesser Antilles subduction zone can yield new constraints, where old oceanic lithosphere formed by slow-spreading subducts slowly. Here we use local earthquakes recorded by the temporary VoiLA (Volatile recycling in the Lesser Antilles) deployment of ocean-bottom seismometers in the fore- and back-arc to characterize the 3-D seismic structure of the north-central Lesser Antilles subduction zone. Along the slab top, mapped based on seismicity, we find low Vp extending to 130–150 km depth, deeper than expected for magmatic oceanic crust. The slab's most prominent, elevated Vp/Vs anomalies are beneath the fore- and back-arc offshore Guadeloupe and Dominica, where two subducted fracture zones lie with the obliquely subducting boundary between Proto-Caribbean and Equatorial Atlantic lithosphere. These structures, therefore, enhance hydration of the oceanic lithosphere as it forms and evolves and the subsequent dehydration of mantle serpentinite when subducted. Above the slab, we image the asthenosphere wedge as a high Vp/Vs and moderate Vp feature, indicating slab-dehydrated fluids rising through the overlying cold boundary layer that might induce melting further to the west. Our results provide new evidence for the impact of spatially-variable oceanic plate formation processes on slab dehydration and mantle wedge volatile transfer that ultimately impact volcanic processes at the surface, such as the relatively high magmatic output observed on the north-central islands in the Lesser Antilles.
... Seismic reflection data were collected in the western half of the Lau Basin as part of the site survey data for ODP Leg 135 , and several attempts were made to image the magma chamber of the Valu Fa Ridge and ELSC (Collier and Sinha, 1992;Turner et al., 1999;Jacobs et al., 2007). Zhao et al. (1997) and Dunn et al. (2013) conducted active-source seismic tomography experiments on the ELSC, revealing persistent seismic low-velocity volumes beneath all segments. Prior to 2019, only two refraction lines crossed the arcto-backarc transition. ...
Article
A 1:1,000,000-scale lithostratigraphic assemblage map of the Lau Basin (southwestern Pacific Ocean) has been created using remote predictive mapping (RPM) techniques developed by geological surveys on land. Formation-level geological units were identified in training sets at scales of 1:100,000–1:200,000 in different parts of the basin and then extrapolated to the areas where geological data are sparse. The final compilation is presented together with a quantitative analysis of assemblage-level crustal growth based on area-age relationships of the assigned units. The data sets used to develop mapping criteria and an internally consistent legend for the compilation included high-resolution ship-based multibeam, satellite- and ship-based gravity, magnetics, seafloor imaging, and sampling data. The correlation of units was informed by published geochronological information and kinematic models of basin opening. The map covers >1,000,000 km2 of the Lau-Tonga arc-backarc system, subdivided into nine assemblage types: forearc crust (9% by area), crust of the active volcanic arc (7%), backarc rifts and spreading centers (20%), transitional arc-backarc crust (13%), relict arc crust (38%), relict backarc crust (8%), and undivided arc-backarc assemblages (<5%), plus oceanic assemblages, intraplate volcanoes, and carbonate platforms. Major differences in the proportions of assemblage types compared to other intraoceanic subduction systems (e.g., Mariana backarc, North Fiji Basin) underscore the complex geological makeup of the Lau Basin. Backarc crust formed and is forming simultaneously at 12 different locations in the basin in response to widely distributed extension, and this is considered to be a dominant pattern of crustal accretion in large arc-backarc systems. Accelerated basin opening and a microplate breakout north of the Peggy Ridge has been accommodated by seven different spreading centers. The result is an intricate mosaic of small intact assemblages in the north of the basin, compared to fewer and larger assemblages in the south. Although the oldest rocks are Eocene (~40 m.y. old basement of the Lau and Tonga Ridges), half of the backarc crust in the map area formed within the last 3 m.y. and therefore represents some of the fastest growing crust on Earth, associated with prolific magmatic and hydro-thermal activity. These observations provide important clues to the geological evolution and makeup of ancient backarc basins and to processes of crustal growth that ultimately lead to the emergence of continents.
Article
Dehydration and hydration reactions in both the downgoing lithosphere and the overlying mantle wedge have been examined in order to understand the role of H2O in the production of magmas at convergent plate boundaries. The subduction of oceanic lithosphere, occurring with increasing pressures and rising temperatures, causes liberation of H2O from the slab. Amphibole, which can be stable to the highest PT conditions among hydrous phases in the slab, decomposes at around 90 km depth. It follows that the subducted lithosphere is essentially anhydrous beneath volcanic arcs lying more than 110 km above the slab and that the supply of slab-derived H2O is not a direct trigger for the production of arc magmas. Instead, the H2O released from downgoing lithosphere reacts with the forearc mantle wedge to crystallize hydrous minerals (serpentine, talc, amphibole, chlorite, and phlogopite). This hydrated peridotite is dragged downward on the slab toward higher PT regions and releases H2O to shallower potential magma source regions in the mantle wedge. Combining experimental data on the stability of serpentine and talc with the thermal structure in the mantle wedge, it is concluded that those minerals decompose beneath the forearc region. On the other hand, high PT experimental and thermodynamic data suggest that dehydration of amphibole and chlorite in the downdragged hydrated peridotite can take place just beneath a volcanic front. Phlogopite in the peridotite decomposes to release H2O at a deeper level (about 200 km). H2O liberated from the hydrated peridotite causes partial melting of overlying mantle wedge peridotites. Along with the migration of H2O through the above processes, subduction components, especially large ion lithophile elements, can be overprinted on the magma source region, which governs the geochemical characteristics of arc magmas.
Article
Subduction at the Tonga Trench gives rise to arc volcanism and plays an active role in forming the Lau (backarc) Basin. Ocean Drilling Program drilling, together with GLORIA imagery, and intensive studies of backarc spreading centers have given us new insight into the evolution of the Lau Basin. Many of the features and processes appear to be common to other western Pacific backarc basins. Among the more significant new findings are recognition that the basin opened in a two-stage process. Initially, crustal extension and rifting formed small sub-basins that were partly filled with basaltic flows and sediments. The second stage of opening involves seafloor spreading from propagating rifts. Although the magma source is in a supra-subduction zone setting, the petrology of the crust indicates that the predominant source of melt has been mantle similar to the MORB-source. Evidence for mixing with melts, on a small scale, shows a contribution from a subduction component (i.e., similar to melts of arc-source mantle). As the basin opened, arc-composition, intrabasin edifices formed alongside the main production of MORB-like crust. A highly heterogeneous source is implied. Isotope data show that two distinct MORB-source mantle components were involved. Older units came from a "Pacific" source, whereas modern basalts come from mantle with an "Indian" source character. Parts of the northern basin carry the isotope signature of OIB source and helium isotope data suggest an influx like that of the Samoan "plume-source".
Article
We applied residual sphere analysis to 25 intermediate- and deep-focus earthquakes in the Tonga subduction zone and mapped variations in slab strike and dip, depth extent and deep lateral advection at all latitudes along the arc. Inferred slab structure varies considerably throughout the subduction zone, and the nearly ubiquitous match between the observed seismicity and detailed undulations in slab strike and dip is evidence that residual spheres do reflect near-source heterogeneity. Seismicity, strain rate and travel time data are consistent with (although they do ot uniquely require) a slab structure in northern Tonga which curves sharply to the west at the base of the upper mantle but which strikes to the north at shallower depths and in the lower mantle. -from Authors
Article
The Office of Naval Research, together with Scripps Institution of Oceanography, University of Washington, Massachusetts Institute of Technology, and Woods Hole Oceanographic Institution, is pleased to announce the formation of two national Ocean Bottom Seismometer (OBS) facilities. Recent advances in marine seismic and acoustic research, including whole Earth tomography, seismic refraction tomography, detailed passive seismology, high‐resolution seismic refraction, and marine ambient noise studies, require a suite of identical calibrated seafloor instruments for analysis of array data collected by OBS capable of sustained deployment periods. Such instruments require a recording capability that is substantially improved in terms of bandwidth, recording capability, fidelity, and deployment duration over that possible just a few years ago. Recognizing a deficiency in existing instrumentation, in 1987 ONR embarked on an effort to fund the design and construction of a new generation of OBS. Thirty‐one instruments are now available for general use, and we encourage investigators to use the national OBS facilities as an effective means of acquiring state‐of‐the‐art ocean floor seismic data. The two OBS facilities will be managed and operated on a joint institutional basis by WHOI and MIT, and SIO and UW, respectively. While the instruments will be managed and operated by the OBS facilities, ownership of the OBS will be retained by
Article
The thermal structure of a generic subduction zone is investigated to elucidate the source region of subduction zone volcanics. The steady state thermal field is evaluated for a model subduction zone where the plates are prescribed by kinematic boundary conditions, such that the subducting slab induces a flow in the mantle wedge. The resulting model suggests that the oceanic crust of the downgoing slab is not melted extensively, if at all, and hence is not the source of subduction zone magmatism. Provided the subducting oceanic crust enters the asthenosphere at a velocity > 6(± 2) cm/yr, the mantle wedge will be hot enough at the limit of the lateral water transport mechanism to generate melting at the amphibole-buffered solidus. Best estimates of the buoyancy sources and appropriate viscosity in the wedge suggest that there is likely to be only a weak modulation of the slab-induced flow unless the slab and wedge are locally decoupled. -from Authors
Article
Three months after its launch, Japan's Advanced Earth Observing Satellite (ADEOS) is set to begin day-to-day science operations at the end of November. The global change research satellite, launched August 16 from the Tanegashima Space Center by Japan's National Space Development Agency (NASDA), has settled into a circular polar orbit at 800 km altitude. ADEOS includes instruments from Japan, the United States, and France that will observe ocean chlorophyll production, ocean temperature, atmospheric gases, polarization and direction of solar energy reflected by the Earth, and the distribution of vegetation.