Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

T~~fo~op~ysies, 8 1 (1982) 133- 159 133

Elsevier Scientific ~blishing Company, Amsterdam-Punted in The Netherlands

SUBDUCTION ZONES AN INTRODUCTION TO COMPARATIVE


SUBDUCTOLOGY **

SEIYA UYEDA

Depurtment oj Geophysics, Texus A&M tJnioersi<y,College Stution, TX 77843 (U.S.A.)*


(Received July 9, 1981)

ABSTRACT

Uyeda, S., 1982. Subduction zones: an introduction to comparative subductology. In: A.L. Hales (Editor),
Geodynamics Final symposium. Tectonophysics, 8 1: 133- 159.

The subduction model for trench-arc-back arc systems was deduced more or less as a logical
consequence of sea-floor spreading. Many features in these systems, such as thrust-type inter-plate
earthquakes and deep earthquakes along Wadati-Benioff zones, are readily explained by the subduction
model. But, other features such as extensional spreading and high heat flow in back arc regions and arc
volcanism are difficult to explain. An approach to solve these problems may be to recognize the existence
of the two basically different modes of subduction: one Chilean-type - causing a compressional stress
regime in the arc and back arc regions; the other - Mariana-type - causing a tensional stress regime.
Implications of the two types of the mode of subduction for some tectonic problems, such as vertical
movement and sediment accretion at trenches are discussed, with special reference to the results of recent
DSDP active margin drilling. In discussing the possible causes of the two modes, it is suggested that the
large-scale thermal and stress regimes in the back arc areas may not be the direct products of simple
subduction, but of some additional factors involving much greater energy, such as mantle flow associated
with the motion of major plates. Finally, the possible great importance of various types of collision,
accretion and erosion at subduction zones is emphasized.

It has long been suspected that various tectonic phenomena associated with the
trench- arc- back arc systems have a common basic cause, namely the down-thrusting
of mantle convection currents (Fisher, 188 I ; Holmes, 1929; Griggs, 1939; and many
others). Later, with the advent of plate tectonics, this idea became more definite and
the process is now called subduction.
The concept of subduction in the framework of plate tectonics was introduced
more or less as a logical consequence of sea-floor spreading to keep the surface area

* On leave from the Earthquake Research Institute, University of Tokyo, Japan I 13.
** Texas A&M Geodynamics Research Program Cont~bution No. 23.

OectO-1951/82/oooO-oooO/$O2.75 @I982 Elsevier Scientific Publishing Company


134

of the earth constant. Although deep sea trenches are obviously the most plausible
candidates for the sites of plate consumption, whether or not the subduction model
can explain the various geological and geophysical features of the trench-arc-back
arc systems is not a priori obvious. In the present paper I review some of the
important problems related to the tectonics of trench-arc-back arc systems or active
margins and present a possible approach to solve these problems. It will be
suggested that both of the major back arc features, namely back arc extension and
high heat flow, and the major mountain belts may not be the simple results of
subduction itself. In addition to subduction, the former may require sea-floor
spreading in the back arc induced by the motions of major plates and the latter may
require the process of collision/accretion of buoyant features.

GENERAL FEATURES OF SUBDUCTION PLATE BOUNDARIES

In order to review the general characteristics of the trench-arc-back arc systems,


an example is taken from the east-west cross section at 40’N of northeast Honshu,
Japan in Fig. 1. As seen in Fig. lB, the crustal structure of the Sea of Japan, a back
arc basin, is oceanic. It is now a widely held view that the Sea of Japan was formed
by the seaward migration of the Japanese islands relative to the Asiatic continent
during some relatively young but disputed geological period. Another notable
feature in Fig. IB is the ~om~ously low Pn velocity (7.5 km/set) of the uppermost
mantle under the main arc of Japan. It is suspected that the asthenosphere or the
low-velocity layer is raised under the Japanese arc almost up to the Moho-
discontinuity. This is in harmony with yet another equally remarkable observation
shown in Fig. lE, that mantle earthquakes are distinctly absent in the uppermost
mantle wedge under the arc. Both facts appear to indicate anomalously high
temperature in the uppermost mantle. The boundary between the outer seismogenic
mantle and the aseismic mantle in Fig. 1E is called the aseismic front (Yoshii, 1975).
The aseismic front is located slightly more oceanward than the volcanic front
(Fig. 2). Heat flow distribution shown in Fig. 1D (Uyeda and Horai, 1964; Yasui et
al., 1968) is also in harmony with the above observations because heat flow is low on
the Pacific side and high on the continent side of the Japanese arc and in the back
arc basin. As can be seen in the figure, the exact position of the transition between
the low and high heat flow zones can not be defined by the presently available data
with sufficient accuracy to be compared with the locations of the volcanic and
aseismic fronts.
If the formation of such a back arc basin as the Sea of Japan is due to sea-floor
spreading as assumed in this paper, the heat flow in the Japan Basin suggests that its
age would be 30-25 m.y. (Anderson, 1980). Altough not clearly shown in Fig. lE, it
has been shown that the lithosphere under the Sea of Japan is as thin as 30 km (Abe
and Ranamori, 1970). This indicates that the upper mantle under the back arc basin
represents a really large-scale thermal anomaly.
135

*n 1 ......
._

20 B
CRUSTAL STRUCTURE t.1

C
-lOO- GRAVITY ANOMALY

Fig. 1. A cross-section of northeast Japan at 40’N. In Fig. IC, Free-air and Bouguer anomalies are shown
for sea and land areas, respectively; in Fig. IE, V and Q mean, the velocity and quality factor of seismic
waves (Adapted from Yoshii, 1979a; Utsu, 1971; Hasegawa et al., 1978).

The shallow earthquakes in the uppermost right comer of the wedge in Fig. IE
are mainly of thrust-type and include almost all the great earthquakes. These
earthquakes are interpreted as caused by the relative motion of the overlying plate
and the subducting slab of the oceanic lithospherewhich has been delineated by the
i

.b
\

Fig. 2. Positions of Quatemary volcanoes and earthquakes with depth of 40-60 km. Black dots show
epicenters of all the events of which focal depths were determined to be in 40-60 km range by the Japan
Meteorological Agency during period 1%7- 1977. Dotted and dashed curves indicate the “volcanic front”
and the “aseismic front” respectively, (Yoshii, 1979b). Open symbols are Quatemary volcanoes and
triangles are presently active ones.

anomalously high V and high Q tongue (Utsu, 1971) as shown in Fig. 1E. The source
mechanism of the shallow events right under the arc and its western extension,
including the central part of the Sea of Japan (IL Shimazaki, personal communica-
tion), is known to be ch~acte~stic~y thrust type, t~tif~ng that the crust of the arc
137

and the back arc of Japanese subduction zone is now under a horizontally compres-
sional tectonic stress. Very shallow earthquakes oceanward of the trench axis (not
shown in Fig. 1E) are of normal fault type and are commonly interpreted as caused
by the downward bending of the oceanic plate before subducting. This agrees with
the observations of numerous normal faults on the oceanward wall of the trench
(Ludwig et al., 1973; Hilde and Sharman, 1978; Honza et al., 1978). Some of the
normal fault type earthquakes, such as Sanriku earthquake, 1933, are of great
magnitude and attributed to the tensile fracture penetrating through the whole
thickness of the subducting slab (Kanamori, 1971). The origin of the tensile stress
for these great normal fault type events is suspected to be the gravitational pull of
the long-extended tongue of the slab.
One of the important recent discoveries concerning the Wadati-Benioff zone
under Japan (Wadati, 1935) is its double-layered structure as shown in Fig. 1E
(Hasegawa et al., 1978). The source mechanism in the upper layer is characteristi-
cally down-dip compression and that in the lower layer is down-dip tension;
unbending of the down-bent slab (Engdahl and Scholz, 1977), sagging of the plate
(Yoshii, 1979a) and thermal stress within it (Goto and Hamaguchi, 1978) are the
examples of the suggested causes of the double-layered Wadati-Benioff zone. Such a
double-layered Wadati-Benioff zone has been found for some other arcs also: e.g.
Central Japan (Tsumura, 1973), Kurile Arc (Veith, 1977) and Mariana Arc (Samo-
witz and Forsyth, 1981).
Now, the question is, “can we explain all these features by the subduction of a
cold slab of oceanic lithosphere ?’ Apparently it is not an easy matter (Uyeda, 1977).
There are two main difficulties, which undoubtedly are closely inter-related; namely
how to explain the extensional tectonics behind convergent plate boundaries and
how to explain the anomalously hot upper mantle under the arc and back arc
regions where a cold slab is subducted. By intuition alone, both appear paradoxical.
In order to overcome these difficulties, various thermo-mechanical models have been
proposed, calling for such mechanisms as frictional heating followed by a diapiric
rise of an enormous amount of magma from the shear zone between the slab and the
mantle wedge (e.g. Hasebe et al., 1970; Oxburgh and Turcotte, 1970; Karig, 1971a)
or a secondary convection cell induced in the m,antle wedge (e.g. Holmes, 1965;
McKenzie, 1969; Sleep and Toksoz, 1973; Hsui and Toksoz, 1979). It may be
important to note, however, that the amount of energy to produce the enormous
thermal anomaly under young back arc basins is orders of magnitude greater that
the potential energy available from subduction of high density slab (Artyushkov,
1981) ruling out, at least, the models in which the latter energy is assumed to be the
main source of the thermal anomaly.
It may, however, be neither necessary nor appropriate to explain all the features
by a single model. Although the Japanese arc is probably one of the typical
subduction zones, all the subduction zones are by no means the same. Moreover, the
features also change with time. For instance, arcs such as Mariana and Tonga have
138

the actively spreading Mariana Trough and Lau Basin behind them, while the
Peru-Chile arc has no back arc basin. The arcs of Japan and Kuriles, on the other
hand, had actively spreading back arc basins in the past but not now. Obviously,
subduction is an important factor but not the only agent that engineers the complex
situation in the trench-arc-back arc systems.

“COMPARATIVE SUBDUCTOLOGY” AND TWO MODES OF SUBDUCTION

Uyeda and Kanamori (1979) classified the subduction zones as shown in Table 1.
The criteria for this classification were related to the nature of the back arc regions.
Namely, the arcs were first grouped into continental and island arcs. The former, by
definition, have no back arc basins. The island arcs, then, were grouped into those
having inactive back arc basins and active back arc basins, each of which were
further divided into subgroups according to the possible origin of the back arc basin
formation. Among these groups, the continental arcs and island arcs with actively
spreading back arc basins are considered to be the two end-members and the other
groups are interpreted as intermediate between, or as hybrids of these end-members.
The former end-member was called the Chilean-type and the latter the Mariana-type.
Japan arc, for instance, is now a Chilean-type subduction zone, because the Sea of
Japan is not actively spreading, but it was most probably a Mariana-type when the
Sea of Japan was being formed.
Uyeda and Kanamori (1979) demonstrated that:
(1) Present-day stress in the back arc areas, deduced from source mechanisms of
intraplate earthquakes within the landward plates, is indeed compressional for the
Chilean-type arcs and tensional for the Mariana-type arcs.
(2) Although interplate thrust-type earthquakes occur at every subduction plate
boundary, truly great earthquakes with the new moment-based magnitude
(Kanamori, 1977a) significantly greater than 8 occur exclusively in the Chilean-type
arcs as can be clearly seen in Fig. 3.
These observations led these authors to conclude that the two types of subduction
zones represent two basically different modes of subduction process. The term mode,
here, represents the strength of mechanical coupling between the subducting slab
and the upper landward plate. In the Chilean-type mode, two plates are closely
coupled whereas in the Mariana-type they are virtually decoupled. A number of
trench-arc-back arc features, which vary from one arc to another, appear to be
explained in terms of the difference in the mode of subduction defined above as
indicated in Fig. 4 (see Uyeda and Kanamori, 1979; Uyeda, 1979; 1981). Although it
is not intended to re-iterate on these features in this paper, some salient ones will be
explained in the following in the light of recent information, notably that from the
active margin drilling of IPOD (von Huene and Uyeda, 1981).
TABLE I

Classification of arcs

Classification Stress regime Typical examples

AK Continental arc (without back arc basins) Compressive (or neutral) Peru-Chile
I Alaska

Island arc Back-arc inactive Back-arc inactivated Compressive (or neutral) Kuril, Japan
(with back-arc basin) Shikoku, Parece
Vela Basins

Back-arc trapped Compressive (or neutral) Bering Sea

Back-arc spreading Tensional Mariana Trough


Scotia Sea,
Lau Basin

Leaky transform Tensional with shear Andaman Sea


0 30 6Q 90 I20 150 -120 -40 0

‘, ‘---1946(8.‘?)

I1 __ I.--
,

-4o- -40

I*
1904 - 1976 :
a- -60
M, ? 8.0 Pius Alaska, I?58 :7.91 Kuriie,1969 (7.8. Azores, 1975 i7.91
Peru,1966 17.5) Peru, 1970 17.8)
rokochi-Qkl,1968 17.9i Peru, 1974 (7.6)

Fig. 3. Great earthquakes from 1904 to lY76. Conventional magnitudes arc in parentheses and new magnitudes are in brackets. Black areas are rupture
zones (Kanamori, 1978).
f
142

CASE HISTORY FOR THE NORTHEAST JAPAN ARC

The depth of trenches are systematically greater for the Mariana-type subduction
zones than the Chilean-type ones. Therefore, if there is a change in the mode of
subduction at a trench over a period of time, there will also be a corresponding
change in its depth. This possibility may be tested for a trench where a change in the
mode in the past is suspected.
Figure5 shows the time change of the water depth of the Sanriku Deep Marine
Terrace as deduced from benthic foraminifera obtained from Site 438/439 of DSDP
Leg 57 (von Huene et al., 1980; Keller, 1981). This figure indicates that the site (see
also Fig. 6) has subsided from sea-level to about 3000m depth since the early
Miocene. This agrees with the view that the subduction at northeast Honshu started
in early Neogene or late Paleogene (Uyeda and Miyashiro, 1974), whatever the
physical cause of the subsidence to.forrn a trench may be (Langseth et al., 1981). An
interesting point to be noted in Fig. 5 for our present discussion is the change in the
trend from subsidence to uplift in the Pliocene. This change, possibly starting in
Upper Miocene, may be interpreted as related to the corresponding change in the
mode of subduction from Mariana-type to Chilean-type. Altough there are still some
problems with regard to the exact timing of events, the above interpretation of the
change in subduction mode will be substantiated in the following discussion.
Stronger coupling between the two plates is expected to work favorably for
scraping off the trench sediments or even oceanic crust on subduction and weaker
coupling may allow easier subduction of sediments down into the mantle. Therefore,
if there is an accretionary prism of oceanic sediments and crustal fragments at all,
the Chilean-type margins are the more likely place to find it. Of course, the
sedimentary structures in the fore arc depends much on the amount of sediment
supply. The fore arc of Sumatra, for instance, receives an enormous amount of
material from the Bengal Fan, and being Chilean-type, develops impressive fore arc
structures (Karig et al., 1979). In spite of the name “Chilean” type, the accretionary
prism does not seem to be particularly well developed at much of the Peru-Chile
Trench, probably because of small supply of sediments from inland (Kulm et al.,
1977), although a well developed accretion complex seems to exist in northern Peru
and the southern Chile Trench where a large supply of sediment is available.
The lack of an accretionary prism in the Andean margin may be related to the
possible occurrence of tectonic erosion. Among various forms of tectonic erosion,
abrasive erosion of the landward plate may be more likely to take place when the
coupling of the plates are stronger. Hence, accretion on the slope and abrasion from
below may be competing processes at typically Chilean-type subduction zones.
Interplay of these competing processes may lead to an extremely complex situation
in a long term development of the Chilean-type margin (Hussong et al.. 1976; Kulm
et al., 1977). Moreover, the horst and graben structure of the subducting plate, that
seems better developed in Ma~ana-tie zones, will also help subduction of sedi-
ments (Hilde and Sharman, 1978).
143

\
\ Sedimenr
\
\
Bosement \
\

Fig. 5. A diagram o$ sediment thickness above the unconformity and depth of depositional environments
for site 438 and 439 (von Huene et al., 1980).

Active margin drilling of IPOD has provided much new information relevant to
the problem of sediment accretion also. In the landward wall of the Mariana Trench,
for instance, no sign of accretion of ocean derived materials was found (Hussong,
Uyeda et al., 1978). Most interesting is the finding that the thick sedimentary wedge

Fig. 6. Japan Trench drill site locations for Legs 56 and 57 of the Deep Sea Drilling Project (von Huene,
Nasu et al., 1978).
144

of the landward wall of the Japan Trench is composed exclusively of terrigenous and
hem&pelagic materials at least to the depth of hole penetration and the room
assigned for possible accretion is curiously small (von Huene, Nasu et al., 1978) as
illustrated in Fig. 6. To the present author, these results appear to be explained in the
same way as the subsidence-uplift history mentioned above: the mode of subduction
at the Japan Trench was Mariana-type during most of the Miocene period and it
became Chilean-type only recently (probably only several m.y. ago), so that the
effective time for possible sediment accretion has been very short.
This supposition that the mode of subduction changed in the Japan Trench
several m.y. ago fits quite well with the change in the tectonic stress of the Japanese
area derived from independent approach as shown in Figs. 7A and 7B (Nakamura
and Uyeda, 1980). These figures indicate the trajectories of the maximum horizontal
stress, unmax, in northeast Japan for the present-day and for the period 21-7
m.y.B.P.: the directions of uHmaxhave been deduced from various lines of evidence
including the directions of dikes, faults, folds, alignment of volcanoes and stress axes
of earthquake source mechanisms. The regional tectonic stress today (Fig. 7A) is
compressional in the direction parallel to that of the plate convergence as expected
for the Chilean-type subduction. The tectonic stress for the 21-7 m.y. period (Fig.
7B) was clearly tensional in the back arc region of northeast Japan, suggesting that
the mode of subduction at that time was Marina-type. It is, therefore, conceivable
that at least part of the Sea of Japan was actively spreading during the above period.
Vertical crustal movements during the late Cenozoic time in Japan have also been
investigated. Matsuda et al., (1967) showed that since early Miocene, the western
zone of northeast Japan has subsided several thousand meters, as shown in Fig. 8A.
This zone of subsidence is characterized by an intensive volcanism known as the
“Green-tuff volcanism”, which was essentially a submarine volcanic episode under
presumably an extensional stress regime. On the other hand, the vertical movement
in Japan during the Quaternary is characterized by an upheaval (Res. Group for
Quat. Tect. Map, 1968) ,as shown in Fig. SB. Recently, Sugi et al. (1982) have
estimated the vertical movements of northeast Japan since ca 17 m.y.B.P. in more
detail. Their results as summarized in Fig. 9 clearly indicate that a rapid subsidence
during the 17-10 m.y. period was followed by a decline of subsidence and then, an
accelerated Quaternary uplift. The general feature is remarkably similar to that of
Fig. 5 and the change in the stress regime (Fig. 7), although the exact timing is still
problematic.
Putting these lines of information together, the Neogene tectonics of northeast
Japan may be described by a dominantly Mariana-type subduction which was
changed to Chilean-type several million years ago.
__.._~_...._ -__- _-
Fig. 8.A. Vertical displacement of Japan since eariy Miocene (Matsuda et al., 1967). Unit in km.
B Vertical displacement of Japan during Quaternary (Research Group for Quat. Tect. Map, 196X). Unit in
km.
r
i

Fig. ?.A. oHmax trajectory map for the present-day northeast Japa~~. Arrows indicate slip vectors far low
angle thrust events. Dots: active volcanoes. Bars: dikes. Trajectory lines are darker where uHma* is judged
to be u, than where it is judged to be u2 (Nakamura and Uyeda, 1980).
B. 0 t+,_ trajectory map for a part of northeast Japan for Miocene (ca. 21-7 m.y.B.P.). The area covered
is the southern portion of the area shown in Fig. 7A. Dotted trajectories show that cHmax is a2 and parallel
lines show that it is not certain if aHmax is ut or 9. Short bars: dikes. {Nakamura and Uyeda, 1980),
146

WESTERN COASTS DEWA HILLS INTRA -MOUNTAIN


BASINS
m

loo0

6U BACKBONE RANGE KITAKAMI MOUNTAINS

_.,,,_p’;;:, Ma
i
Ma

__,, BB

___-•---- CC’

v/ INTRA -MOUTAIN BASI NS

Fig. 9. Vertical movement since 17 m.y.B.P. of northeast Japan (Sugi et al., 1982)

CASES OF THE MARIANA, MEXICO AND GUATEMALA TRENCHES

Figure 10 is the schematic representation of some of the drilling results obtained


from the four IPOD Trench transects, i.e. Japan, Mariana, Mexico off Oaxaca and
Guatemala Transects. The case for the Japan Trench, discussed in the last section, is
included in Fig. 1OA as a guide to other cases.
As mentioned already, no ocean derived material was recovered from the Mariana
(Fig. 10B) fore arc drilling and the vertical tectonics in the Mariana fore arc has
been that of dominant subsidence ever since the initiation of the present subduction
in probably Eocene time (Hussong, Uyeda et al., 1978).
147

This, in our view, is the typical case for the Mariana-type subduction. Detailed
history of the Philippine Sea, however, indicates that the spreading of back arc
basins has been episodic; i.e. 32-15 m.y. for the Pareee Vela Basin and 6-O m.y. for
the Mariana Trough (Kroenke et al., 1978; Hussong, Uyeda et al., 1978). Therefore,
it may be inferred that during the pause of spreading, namely in 15-6 m.y. period,
the mode of subduction could have been more of the Chilean-type. More detailed
investigation on the vertical tectonics in future may well reveal such a history as
shematically shown in Fig. 10B by a broken line.
From the fore arc of the Mexican Trench off Oaxaca (Fig. lOC), landw~d-~pp~g
reflectors observed in seismic profiling were found to be the progressively seaward
accretion of the trench sediments (Moore, Watkins et al., 1979). (Even in this case,
most or all of the truly oceanic components are considered to be consumed.) Vertical
movements across the transect have been studied by McMillen and Bachman (1982)

4381439 Japan Trench

Mariana Trench

*A* MexicoTrench

Guatemala Trench

Cross section Depth VS. time

Fig. 10. Simplified representation of the results of IPOD drilling at four active margins. Left-hand figures
are cross sections and right-hand figures are the estimated time variations of depth. (Compited from van
Huene and Uyeda, 1981; McMill~ and Bachman, 1982).
148

as summarized in Fig. 1OC. Upheaval during the last ca 10 m.y. at deeper sites (488,
492,491) started with a rapid rise and became slower at a later stage, corresponding
to the initial uplift at the trench toe and slower uplift on the slope. The former
process is a local one of imbrication and the latter may be of more regional origin as
it is approximately equal to the uplift rate on the continental crust (sites 489, 493). I
consider these latter movements to reflect, at least partially, the vertical tectonics of
interest here. The curves for the shallower sites show a very rapid (1 km/m.y.)
subsidence in the early Miocene. It may be inferred that this early subsidence
corresponds to the onset of the present phase of subduction, roughly synchronous to
the age of the trans-Mexican volcanic belt, and the later uplift corresponds to
accretion and vertical tectonics inherent to the Chilean-type subduction. Whether or
not the earliest several m.y. mark the Mariana-type subduction is not certain. Lastly,
in Fig. IOD, the case of the Guatemala Trench (von Huene, Aubouin et al., 1980)
revealed practically no sediment accretion, and the benthic foraminifera data
indicated dominant subsidence of the trench slope since early Miocene, probably
attesting to the occurrence of the Mariana-type subduction throughout the above
period.
The above reasoning also agrees with the dominantly extensional tectonics in the
Mariana and Guatemala fore arcs (Hussong, Uyeda et al., 1978; von Huene,
Aubouin et al., 1980) and contractional tectonics in the modem sediments of the
Japanese fore arc (von Huene, Nasu et al., 19’78) and in the Mexican fore arc
(Moore, Watkins et al., 1979).
It would be very pertinent to ask why the two relatively closely spaced parts of
the Middle American Trench show such different trench tectonics, or the contrasting
modes of subduction. A possible answer to this question will be given later after a
discussion of the possible causes of the two modes of subduction is presented.

POSSIBLE CAUSES OF THE TWO MODES OF SUBDUCTION AND THE ENIGMA OF THE
MIDDLE AMERICAN TRENCH

Here, one is concerned with the coupling between the buoyant landward plate,
and the sinking oceanic plate at subduction zones (Uyeda and Kanamori, 1979).
Three posibilities have been suggested:
(1) The difference in the strength of coupling is ascribed to the difference in the
nature of the contact zone between the two plates. More specifically, Kanamori
(1977b) suggested that the different modes of subduction represent different stages
of an evolutionary process: the subduction starts with low-angle thrusting of the
Chilean-type mode and as the process goes on, the coupling gradually weakens and
finally the mode becomes the Mariana-type.
(2) The age of the subducting plate controls the mode of subduction, because the
older the plate, the colder and heavier it is and, hence, the more likely it is to sink
with greater speed and less coupling with the landward plate, giving rise to the
149

Mariana-type mode. Conversely, when the subducting plate is younger, the Chilean-
type mode will result (Molar and Atwater, 1978).
(3) The motion of the landward plate relative to the trench line controls the mode
(Chase, 1978; Uyeda and Kanamori, 1979). Namely, if the landward plate tends to
go away from the trench line, as with the Philippine Sea plate, the Mariana-type
mode of subduction occurs, whereas when the landward plate advances toward the
trench, as with the South American plate, the mode is Chilean-type. Although the
trench line in general should move, when the slab of the oceanic plate sinks to a
great depth in the mantle, it may be, so to speak, anchored to the mantle for its
lateral movements. Since the motions in the deep mantle are probably much slower
than the plate motions, we may then be able to regard the position of the trench line
as virtually stationary. Then, the relative motion of the upper plate and the trench
line can be approbated by the absolute velocity of the upper plate alone. Figure 11
shows the absolute velocity vectors of the c&urn-Pacific upper plates taken from
Minster et al. (1974) and Fitch (1972). It can be seen in the figure that the possibility
(3) is well supported.
The above three possibilities all seem to contain some truth in them. Recently,
Dewey (1980) combining the possibilities (2) and (3), developed a general description
of the problem. Possibility (1) appears partly inherent to the subduction process and

+90 tlzo” +150 t180” -WI -120” -90 -80 -30” 0’


I ’ f ’ ’ ’ ’ ’ 1

'-I ‘2.6

I- L 3.5 4 !? 7 y.2.i j

-40
I- -40’

ttrill
+90 +120” T150” 1tl80’ -150” -120” -90’ -60” -30’ 0’

Fig. Il. Absolute velocity vectors of upper plates around the Pacific (after Minster et al., 1974 and Fitch,
1972). ‘l’be number attached to arrows are velocities in cm/yr. (Uyeda, 1979).
150

may explain some cyclic nature in the arc evolution (Kobayashi and Isezaki, 1976;
Niitusuma, 1978). If, however the process in possibility (1) is the sole agent, one
would have to prove that similar cycles have taken place at every subduction zone,
including the Peru-Chile zone. Possibility (2) is physically sound and well supported
by the present-day data to a first approximation. Without some other factors,
however, it may be difficult to explain the episodic nature of back arc spreading or
the time-variation of the mode of subduction from one type to another. Possibility
(3) is in a way free from this difficulty, because the major role is assigned to the
absolute velocity of the landward plate. In the case of the South American continent,
its absolute velocity has probably been steady, at least in its sense, during the
opening of the Atlantic Ocean, and if so, the mode of subduction in the western
South American coat may have been essentially of Chilean-type throughout the
period. On the other hand, the absolute velocity of the Eurasian plate at its eastern
margin (Fig. ll), which is relevant to the discussion of the western Pacific subduc-
tion zones, is very smaIl at present (Minster et al., 1974). Therefore, it seems quite
possible that the velocity had a component away from the trench lines in relatively
recent geologic past.
From these considerations, a simple solution to the enigma of the Middle
American Trench, introduced in the previous section, may be found as shown in Fig.
12. The key point is the position of the Mexican and Guatemala Transects relative to
the Caribbean plate. At the Mexican Trench, the landward plate is a part of the
westward moving North American plate, whereas at the Guatemala Trench, it is the
Caribbean plate of which present day absolute motion is not significantly different

Fig. 12 Simplified tectonic system of the Middle America. NA =North American plate; SA =South
American plate; CAR=Caribbean plate; CO=Cocos plate; NAZ=Nazca plate; M=Legg 66 area;
G=Leg 67 area; GT=Grenada Trough; MF=Motagua Fault.
I.51

from zero (Jordan, 1975). The absolute velocity vector for Middle America in Fig. 11
is that of the South American plate because the model AM-l of Minster et al. (1974)
does not include the Caribbean plate. The absolute velocity vectors of the Caribbean
plate given in Jordan (1975) and Minster and Jordan (1978) are directed more
northerly and smaller in magni,ude (- lcm/yr), providing a component away from
the Middle America Trench. Although the absolute motion of the Caribbean plate is
uncertain, its eastward motion relative to the North American plate is undeniable
from the nature of its northern transform boundaries and eastern subduction
boundary at the Antilles arc. Thus, the landward plate at the Guatemala Trench is
much more likely to give rise to the Mariana-type su~uction than at the Mexican
Trench. In fact, the presence of diffuse zone of volcanism and graben structure south
of Motagua fault system indicates that the stress regime in Guatemala is extensional
(Plafker, 1976) and, therefore, the mode of subduction is Mariana-type. Probably the
situation has been the same for some geologic past. This is consistent with the
drilling result that showed subsidence and no sediment accretion since early Miocene.
If, however, back arc spreading was taking place to generate the Grenada Through
behind the Antilles Trench at some time in the past, the mode of subduction at the
Antilles Trench then should have been Mariana-type. In that case, the motion of the
Caribbean plate would have been westward relative to the Antilles Trench and,
therefore, in turn, the subduction at the Guatemala Trench could have been
Chilean-type. Therefore, to examine if the timing of the possible back arc spreading
in the Grenada Trough far to the east is matched by the changes in the mode of
subduction at the Guatemala Trench to Chilean-type would be an interesting test of
the model presented here. It may be suspected that the Nicoya Ophiolite Complex in
Costa Rica (Kuijpers, 1980) is extending under the fore arc zone of the Guatemala
Trench (Seely, 1979). Such an abduction of ophiolite body might mark either the
event of the initiation of subduction or strongly coupled subduction (the Chilean-
type). However, the present discussion is concerned mainly with the present phase of
subduction. In order to develop a more .chomprehensiveargument on the tectonic
evolution of the region, may more factors such as the truncation of ancient geologic
structures either by tectonic erosion or lateral displacement will have to be taken
into consideration.
Aside from these possible complications during earlier tectonic phases, it seems
that at least the modem differences between the Mexican and the Guatemala
Trenches can be explained by the difference in the mode of subduction, the Mexican
Trench being more of Chilean-type and the Guatemala Trench more of Mariana-type,
and the different modes of subduction along these juxtaposed margins are probably
due to the difference in the absolute motion of the corresponding landward plates.

APPLICATION OF THE TWO MODES OF SUBDUCTION TO METALLOGENESIS

‘Ike nature of volcanic rocks is closely related to the nature of tectonic settings
(e.g. Miyashiro, 1975). From the ~nsideration of the possible role of the tectonic
152

stress in controlling the composition of erupted magma, it may be expected that the
Chilean-type subduction zones have more abundance of talc-alkaline adesites than
the Mariana-type subduction zones. Namely, it may be inferred that under a
compressive stress regime, magma is not allowed to reach the earth’s surface as freely
as under a tensional regime, so that it has more chance to become talc-alkaline
andesite through processes such as fractionation and assimilation. This view is
undoubtedly supported by observation on the end-member subduction zones, namely
andesites are much more abundant in the Andes than in the Marianas.
Somewhat related to the above is the possible role of the different modes of
subduction on metallogenesis. Figures 13A and 13B show the world distributions of
post-Mesozoic porphyry copper and massive sulphide deposits, respectively. The
genesis of these copper ores is believed to be the result of volcanic activities at
convergent plate boundaries. Figure 13A shows that the porphyry copper deposits
are abundant in both the North and South American coastal regions, but are
distinctly scarce in the western Pacific island arcs of Aleutians, I&riles, Japan,
Ryukyu, Mariana and Tonga. Instead, there are a number of massive sulphides
(mostly Kuroko-type) in Japan (Fig. 13B). The absence of porphyry copper deposits
in the western Pacific island arcs has long been an enigma in economic geology.
Uyeda and Nishiwaki (1980) pointed out that the uneven and complementary
occurrences of porphyry copper and massive sulphide deposits can be explained by
the difference in the stress field between the Chilean- and Mariana-type subduction
zones: porphyry copper mineralization is favored by compressive tectonic stress of
the Chilean-type subduction zones whereas massive sulphide mineralization is favored
by submarine hydrothermal activities in the rift-like situations of the back arc
regions of the Mariana-type subduction zones. It should be recalled that Japanese
subduction was probably Mariana-type during most of Miocene when the massive
sulphide deposits were formed in association with the submarine “Green-tuff’
volcanism in the back arc zone, which may have been an aborted back arc spreading
center. From these considerations, it can be expected that present-day active
metallogenesis of massive sulphide is taking place in the floor of the actively
spreading back arc basin such as the Mariana Trough, where active hydrothermal
circulation has recently been discovered (Hobart et al., 1979). With regard to the
incompatibility of porphyry copper and massive sulphide deposits, Sillitoe (1980)
has come to more or less the same conclusion but from a quite different standpoint.
In Fig. 13A, it may be noticed that many porphyry copper deposits occur along
the southwestern Pacific islands. from the Philippines to the Solomons. The mode of
subduction in these areas does not appear to be characterized as typically Chilean-
type. Uyeda and Nishiwaki (1980), however, pointed out that these islands have
undergone complex tectonic history, involving collisions of arcs and continental
fragments and reversals of the polarity of subduction. Since the collision process
invariably produces strong compressive stress, the occurrences of porphyry copper
deposits in these islands (including New Guinea) fit the above stated premise. This
B

Ksrmodsc-Tonqo

Antorctrc Plate
A”t.rC+iC Plots

- S”bd”c+ron lWS cf--c co:,irwJn __ ironstorm


_-- - wlc*r+oin p,a+* wundory - f+i*ps ox,s - Dtractlon 0, pm+e motian

A Port *sromz *o**wo *“+Phlde d*p*tr+* A Present r”Dmorlne txJ+ w(i+*, J”,cp

Fig. 13.A. Distribution of young porphyry copper deposits (Uyeda and Nisbiwaki, 1980). B. Distribution
of young massive sulphide deposits (Uyeda and Nishiwaki, 1980).
Fig. 14. Generalized map of Cordilleran Suspect Terranes. Dashed pattern=North American autoch-
thonous cratonic basement. Barbed Zinezeastern limit of Cordilleran Mesozoic deformation. Barbed
arrows=direction of major strike-slip movements. Abbreviated symbols such as NS. Kv, En etc. arc the
suspect terranes (after Coney et al., 1980).
final point is suggestive of the possible importance of collisions at convergent plate
boundaries.

IMPORTANCE OF COLLISION/ACCRETION TECTONICS

If the process of subduction continues, as plate tectonics asserts, and the


intervening oceanic crust is all subducted, collisions of continents or island arcs are
inevitable. Moreover, the world’s oceans have numerous uplifted features of various
sizes and origins, i.e. seamounts, islands, ridges, arcs, plateaus and micro-continents.
These features will arrive at the subduction zones and collide with the landward
plate before the final closure of oceans. In fact, numerous actual and past instances
of such collisons have been documented and their importance in tectonics em-
phasized (e.g. Dewey and Bird, 1970; Vogt et al., 1976). Recently, the interest in
collision/accretion have been remarkably heightened with the recognition that
numerous terranes on the circum-Pacific continents may be the accreted features
(e.g. Fujita, 1978; Kerr, 1980; Churkin and Trexler, 1980; Coney et al., 1980; Saito
and Hashimoto, 1982; Fig. 14), and even might have come from a single hypotheti-
cal lost continent “Pacifica” (Nur and Ben-Avraham, 1977). We still do not know
exactly what tectonic significance accretion may have, in addition to contributing to
the areal increase of continents. One of the basic questions may be posed as “Can
the Cordillera be formed by strai~t-fo~ard subduction (presumably of the Chilean
-type) alone or it can be formed only when accretion of exotic terranes is involved?’
Sometime ago we (Matsuda and Uyeda, 197 1) proposed the “Pacific-type” orogeny
as one of the basic forms of orogeny, contrasting to the collison type one. Now, the
third type that might be called “Accretion-type” orogeny is emerging. In fact, this
third type may be much more common and important than the others. The other
basic question would be the relation of collision/accretion with tectonic erosion of
various types which also has been suggested as an important process at continental
margins (e.g. Karig, 1974; Hussong et al., 1976; Kulm et al., 1977; Murauchi and
Ludwig, 1980). In the next years to come, evaluation of the relative roles of
subduction collison, accretion and tectonic erosion at convergent plate boundaries
would be one of the major objectives of geotectonics.

CONCLUSIONS

Problems related to subduction zones and back arc basins were reviewed from the
standpoint of “comparative subductology”, and the existence of two basic modes of
subduction (i.e. Chilean-type and Mariana-type) was postulated. It was shown that
the concept of the two modes of subduction zones may be viable in understanding
various arc related phenomena in terms of the tectonic stress; especially the recent
results of DSDP drilling at active margins. Finally, the possible importance of
~ollision/accretion and erosion in the process of subduction was emphasized.
156

REFERENCES

Abe, K. and Kanamori, H., 1970. Mantle structure beneath the Japan Sea as revealed by surface waves.
Bull. Earthquake Res. Inst., 48: 101 I-1021.
Anderson, R.N., 1980. update of heat flow in the east and southeast Asian Sea. In D. Hayes (Editor),
Tectonic/Geologic Evolution of Southeast Asia. Geophys. Monogr., Am. Geophys. Union, 23:
319-326.
Artyushkov. E., 1981 Mechanism of formation of active margin. Oceanol. Acta, 4 (Suppi.), Colloque C3.
26th Geol. Congr., Paris, pp. 245-250.
Chase, C., 1978. Extension behind island arcs and motions relative to hot-spots. J. Geophys, Res.. X3:
5385-5387.
Churkin, M. and Trexier, J.H., 1980. Circum-Pacific plate accretion-isolating part of a Pacific Plate to
form the nucleus of the Artic Basin. Earth Planet. Sci. Lett., 48: 356-362.
Coney, P.J., Jones, D.L. and Monger, J.W.H. 1980. Cordilleran suspect terranes. Nature. 28X: 329-333.
Dewey, J.L., 1980. Episodicity, sequence and style at convergent plate boundaries. In: D. Strangway
(Editor), The Continental Crust and its Mineral Resources. Geol. Assoc. Canada, Spec. Pap.. 20:
553-574.
Dewey, J.F. and Bird, F.M., 1970. Mountain belts and the new global tectonics. J. Geophys. Res.. 75:
2625-2647.
Engdahl, R. and Scholtz, C.H., 1977. A double Benioff zone beneath the Central Aleutians; an unbending
of the lithosphere. Geophys. Res. Lett.. 4: 473-476.
Fisher, 0.. 1881. Physics of the Earth’s Crust, MacMillan and Co., London, 299 pp.
Fitch, T.J., 1972. Plate convergence, transcurrent faults, and internal deformation adjacent to southeast
Asia and the western Pacific. J. Geophys. Res., 77: 44324460.
Fujita, K., 1978. Pre-Cenozoic tectonic evolution of Northeast Siberia. J. Geol., X6: 159- 172.
Goto, K. and Hamaguchi, H., 1978. A double-planed structure of the intermediate seismic zone-thermal
stress within the descending slab. Abstr. Annu. Meet. Seismol. Sot. Japan, 2: 36 (in Japanese).
Griggs, D., 1939. A theory of mountain building. Am. J. Sci., 237: 6 I I-650.
Hasebe, K., Fujii, N. and Uyeda, S., 1970. Thermal processes under island arcs. Tectonophysics. IO:
335-355.
Hasegawa, A., Umino, N. and Takagi, A., 1978. Double-planed deep seismic zone and upper-mantle
structure in the Northeastern Japan Arc. Geophys. J. R. Astron. Sot., 54: 281-296.
Hilde, T.W.C. and Sharman, G.F., 1978. Fault structure of the descending plate and its influence of the
subduction process (abstr.). EOS, Trans Am. Geophys. Union, 59: I 1X2.
Hobart, M.A., Anderson, R.N. and Uyeda, S., 1979. Heat transfer in the Mariana Trough (Abstr.). EOS,
Trans. Am. Geophys. Union, 60: 383.
Holmes, A.. 1929. Radioactivity and earth movements. Trans. Geol. Sot. Glasgow, IX: 559-606.
Holmes, A., 1965. Principles of Physical Geology. Nelson, London, 1288 pp.
Honza, E.. Tamaki, K. and Murakami, F. (compilers), 1978. Geological Map of the Japan and Kuril
Trenches and the Adjacent Areas. Geol. Surv. Japan, Mar. Geol. Map Ser.. I I.
Hsui, A. and Toksih, N.. 1979. The evolution of thermal structure beneath a subduction zone.
Tectonophysics, 60: 43-60.
Hussong, D.M., Edwards. P.B., Johnson, S.H., Campbell, J.F. and Sutton, G.H., 1976. Crustal structure
of Peru-Chile Trench: 8’- 12’S latitude. In: G. Sutton et al. (Editors), The Geophysics of the Pacific
Ocean Basin and its Margin. Geophys. Monogr., Amer. Geophys. Union, 19: 71 -X5.
Hussong, D., Uyeda, S. and Scientific Party, DSDP Leg 60, 1978. Leg 60 ends in Guam. Geotimes. 23
(IO): 19-22.
Jordan, T.H., 1975, The present-day motions of the Carribean plate. J. Geophys. Res., 80: 44334439.
Kanamori, H., 1971. Seismological evidence for a lithospheric normal faulting-The Sanriku earthquake
of 1933. Phys. Earth Planet. Inter., 4: 289-300.
157

Kanamori, H., 1977a. The energy release in great earthquakes. J. Geophys. Res., 82: 2981-2987.
Kanamori, H. 1977b. Seismic and aseismic slip along subduction zones and their tectonic implications. In:
M. Talwani and W.C. Pitman III, (Editors), Island Arcs, Deep Sea Trenches, and Back-Arc Basins.
Maurice Ewing Ser., Am. Geophys. Union, 1: 16%174.
Kanamori, H., 1978. Quantification of earthquakes. Nature, 271: 41 I-414.
Karig, D.E., 1971. Origin and development of the marginal basins in the western Pacific. J. Geophys.
Res., 76: X42-2561.
Karig, D.E., 1974. Tectonic erosion at trenches. Earth Planet. Sci. Lett.. 21: 209-212.
Karig, D., Suparka, S., Moore, G.F. and Hehanussa, P.E., 1979. Structure and Cenozoic evolution of the
Sunda arc in the Central Sumatra region. In: J. Watkins et al. (Editors), Geological and Geophysical
Investigation of Continental Margins. Mem. Am. Assoc. Pet. Geol., 29: 223-239.
Keller, G., 198 t . Benthic foraminifera and p~eobathymet~ of the Japan Trench area, DSDP Leg 57. In:
Initial Rep. DSDP, LVI-LVII, pt. 2, U.S. Govern. Printing Office, Washington. D.C., pp. 83.5-865.
Kerr, R.A., 1980. The bits and pieces of plate tectonics. Science, 207: 1059- 1061.
Kobayashi, K. and Isezaki, N., 1976. Magnetic anomalies in Japan Sea and Shikoku Basin and their
possible tectonic implications. In G. Sutton et al. (Editors), The Geophysics of Pacific Ocean Basin
and its Margins. Geoph$s. Monogr., Am. Geophys. Union, 19: 235-251.
Kroenke, L., Scott, R. et al., 1978. Old questions answered-and new ones asked (Leg 59). Geotimes, 23
(7): 20-23.
Kuijpers, E.P., 1980. The geologic history of the Nicoya Ophiolite Complex, Costa Rica, and its
geotectonic significance. Tectonophysics, 68: 233-255.
Kulm, L.D., Schweller, W.J. and Mathias, A., 1977. A preliminary analysis of the subduction processes
along the Andean continental margin, 6’ to 45”s. In M. Talwani and W.C. Pitman III (Editors),
Island Arcs, Deep Sea Trenches and Back-Arc Basins. Maurice Ewing Ser., Am. Geophys. Union, 1:
28.5-302.
Langseth, M.G., von Huene, R., Nasu, N. and Okada, H. 1981. Subsidence of the Japan Trench forearc
region of north Honshu. Oceanol. Acta, 4 (Suppl.), Colloque C3, 26th Geol. Congr, Paris, pp.
173-179.
Ludwig, W., Den, N. and Murauchi, S., 1973. Seismic reflection measurements of southwest Japan
margin. J. Geophys. Res., 78: 25082516.
Matsuda, T.. Nakamura, K. and Sugimura, A., 1967. Late Cenozoic orogeny in Japan. Tectonophysics, 4:
349-366.
Matsuda, T. and Uyeda, S., 1971. On the Pacific-type orogeny and its model--Extension of the paired
belts concept and possible origin of marginal seas. Tectonophysics, 1 I : 5-27.
McKenzie, D.E.. 1969. Speculations on the consequences and causes of plate motions. Geophys. J. R.
Astron. Sot., 18: l-32.
McMillen, K.J. and Bachman, S.B., 1982. Bathymet~c and tectonic evolution of the southern Mexico
active margin, DSDP Leg 66. In: Init. Rep. DSDP, 66, U.S. Govern. Printing Office, Washington,
D.C., in press.
Minster, J.G. and Jordan, T.H., 1978. Present-day plate motions. J. Geophys. Res., 83: 5331-5354.
Minster, J.G., Jordan, T.H., Molnar, P. and Haines, E., 1974. Numerical modelling of instantaneous plate
tectonics. Geophys. J. Astron. Sot.. 36, 541-576.
Miyashiro, A., 1975. Volcanic rock series and tectonic setting. Annu. Rev. Earth Planet. Sci., 3: 2.5 l-269.
Molnar. P. and Atwater, T., 1978. Interarc spreading and Cordilleran tectonics as aternates related to the
age of subducted oceanic lithosphere. Earth Planet. Sci. Lett., 41: 330-340.
Moore, J.C. 1979. Watkins, J.S., and Scientific Party of Leg 66, DSDP, 1979. Middle America Trench off
Mexico. Geotimes, 24 (9): 20-22.
Murauchi, S. and Ludwig, W.J., 1980. Crustal structure of the Japan Trench: the effect of subduction of
oceanic crust. In: Init. Rep. DSDP, LVI-LVII, pt. 1, U.S. Govern, Print. Off., Washington. D.C.. pp.
459-469.
158

Nakamura. K. and Uyeda, S., 1980. Stress gradient in arc-back arc regions and plate bubduction. J.
Geophys. Res., 85: 6419-6428.
Niitsuma, N.. 1978. Magnetic stratigraphy of the Japanese Neogene and the development of the island
arcs of Japan. J. Phys. Earth, 26 (Suppl.): 5367-7378.
Nur, A. and Ben-Avraham, Z., 1977. Mountain building and lost Pacifica continent. Nature. 270: 41-42.
Oxburgh, E.R. and Turcotte, D.L., 1970. Thermal structure of island arcs. Geol. Sot. Am. Bull., 81:
1665-1688.
Plafker, G., 1976.Tectonic aspects of Guatemala earthquake of 4, Feb. 1976. Science. 193: I201 - 1208.
Research Group for Quaternary Tectonic Map, 1968. Quaternary Tectonic Map of Japan. Quaternary
Res.. 7: 182- 187(in Japanese with English abstr.).
Samowitz, I. and Forsyth, D., 198 I. Double seismic zone beneath the Marianas Island Arc. J. ticophys.
Saito, Y. and Hashimoto, M., 1982. An aIl~hthonous terrane in Japan: South Kitakami Region. Res.. X6:
7013-7021.
Seely. D.. 1979. The evolution of structural highs bordering major forearc basins. In: .I. Watkins et al.
(Editors), Geological and Geophysical Investigations of Continental Margins. Mem. Am. Assoc. Pet.
Geol.. 29: 245-260.
Sillitoe, R., 1980. Are porphyry copper and Kuroko-type massive sulphide deposits incompatible?
Geology, 8: I I - 14.
Sleep. N. and Toksoz, M.N., 1973. Evolution of marginal basins. Nature, 233: 548550.
Sugi, N., Chinzei, K. and Uyeda, S., 1982. Vertical crustal movements of the Northeast Japan since
middle Miocene. Geodynamics Ser., Am. Geophys. Union /Geol. Sot. Am., in press.
Topper, R.E., 1979. Fine Structure of the Benioff Zone beneath the Central Aleutian Arc. M.S. Thesis,
Univ. of Colorado (unpubl.).
Tsumura, K.. 1973. ~cr~~thqu~e activity in the Kanto District. In: Pub. 50th Ann. Great Kanto
Earthq. 1923. Earthquake Res. Inst., Tokyo, pp. 67-87.
Utsu, T.. 1971. Seismological evidence for anomalous structure of island arcs with special reference to the
Japanese region. Rev. Geophys. Space Phys., 9: 839-890.
Uyeda, S., 1977. Some basic problems in the trench-arc-back arc system. In: M. Talwani and W.C.
Pitman III (Editors), Island Arc, Deep Sea Trenches and Back-arc Basins. Maurice Ewing Ser.. Am.
Geophys. Union, 1: I - 14.
Uyeda, S., 1979. Subduction zones: facts, ideas and speculations. Oceanus, 22: 52-62.
Uyeda, S., 1981. Subduction zones and back arc basins-A review. Geol. Rundsch., 70: 552-569.
Uyeda, S., and Horai, K., 1964. Terrestrial heat flow in Japan. J. Geophys. Res.. 69: 2 I2 I-2 I4 I.
Uyeda, S. and Kanamori, H., 1979. Back-arc opening and the mode of subduction. J. Gcophys. Res.. X4:
1049-1061.
Uyeda, S. and Miyashiro, A.. 1974. Plate tectonics and the Japanese Islands: A synthesis, Gcol. Sot. Am.
Bull.. 85: 1l59- 1170.
Uyeda, S. and Nishiwaki, C.. 1980. Stress field, metallogenesis and mode of subduction. In: D.W
Strangway (Editor), The Continental Crust and its Mineral Deposits. Geol. Sot. Can.. Spec. Pap.. 20:
323-339.
Veith, K.F., 1977. The nature of the dual zone of seismicity in the Kuriles arc. EOS, Trans. Am Cieophys.
Union, 58: 1232.
Vogt, P.R., Lowrie, A., Bracey, D.R. and Hey, R.N.. 1976. Subduction of aseismic ridges: Effects on
shape, seismicity. and other characteristics of consuming plate boundaries. Geol. Sot. Am. Spcc. Pap..
172, 59 pp.
von Huene. R., Langseth, M., Nasu, N. and Okada, H., 1980. Summary, Japan Trench Transect. In: Init.
Rep. DSDP LVI-LVII, pt. 1, US Govern. Printing. Office., Washington., D.C., pp. 473-48X.
von Huene, R. and Uyeda, S.. 1981. A summary of results from the IPOD active margin transects across
the Japan, Mariana, and Mid-American convergent margins. Oceanol. Acta. 4 (Suppl.), Colloque C3.
26th Geol. Con@., Paris, pp. 233-239.
159

van Huene, R., Nasu, N., and Scientific Party of DSDP Leg 57, 1978. Japan Trench transected. Geotimes,
23 (4): 16-20.
van Huene, R.. Aubouin, J., Azema, J., Blackington, G., Carter, J., Coulbourn, W., Cowan, D., Curiale, J..
Dengo, C., Faas. R., Harrison, W., Hesse, R., Hussong, D., Ladd, J., Muzlov, N.. Shiki, T.. Thompson,
P. and Westberg, J., 1980. The IPOD Middle-America Trench transect off Guatemala. Geol. Sot. Am.
Bull., 91: 421-432.
Wadati, K., 1935. On the activity of deep-focus earthquakes in the Japan Islands and neighborhoods,
Geophys. Mag., 8: 305-325.
Yasui, M., Kishii, T., Watenabe, T. and Uyeda, S., Heat flow in the Sea of Japan. In: L. Knopoff et al.
(Editors), The Crust and Upper Mantle of the Pacific Area. Geophys. Monogr. Am. Geophys. Union,
12: 3-16.
Yoshii, T., 1975. Proposal of the “Aseismic Front”. Zisin, 28: 365-367 (in Japanese).
Yoshii, T., 1979a. A detailed cross-section of the deep seismic zone beneath northeastern Hanshu, Japan,
Tectonophysics, 55: 349-360.
Yoshii, T. 1979b. Complication of geophysical data around the Japanese Islands (I). Bull. Earthquake
Res. Inst., 54: 75- I17 (in Japanese with English abstract).

You might also like