Thermal evolution of the Tertiary Shimanto Belt, Muroto Peninsula, Shikoku, Japan

17
The Island Arc (1992) 1, 116-132 Thematic Article Thermal evolution of the Tertiary Shimanto Belt, Muroto Peninsula, S hikoku, Japan MICHAEL B. UNDERWOOD, MATTHEW M. LAUGHLAND, * TIM BYRNE,~ J. P. HIBBARD3 AND LEE DITULLIo4 'Department of Geological Sciences, University of Missouri, Columbia, MO 652 1 1, 'Department of Geology and Geophysics, University ofconnecticut, Storrs, CT 06268, 3MEAS, Box 8208, North Carolina State University, Rakigh, NC 27695, USA and 4Department of Geology, Kochi University, Kochi 780 Japan Abstract The Shimanto accretionary complex on the Muroto Peninsula of Shikoku comprises two major units of Tertiary strata: the Murotohanto Sub-belt (Eocene-Oligocene) and the Nabae Sub-belt (Oligocene-Miocene). Both sub-belts have been affected by thermal overprints following the peak of accretion-related deformation. Palaeotemperatures for the entire Tertiary section range from - 140 to 315"C, based upon mean vitrinite reflectance values of 0.9-5.0%Rm. Values of illite crystallinity index are consistent with conditions of advanced diagenesis and anchimetamorphism. Illite/mica b, lattice dimensions indicate that burial pressures were probably no greater than 2.5 kbar. In general, levels of thermal maturity are higher for the Murotohanto Sub-belt than for the Nabae Sub-belt. The Eocene- Oligocene strata also display a spatial decrease in thermal maturity from south to north and this pattern probably was caused by regional-scale differential uplift following peak heating. Conversely, the palaeothermal structure within the Nabae Sub-belt is fairly uniform, except for the local effects of mafic intrusions at the tip of Cape Muroto. There is a paleotemperature difference of - 90°C across the boundary between the Murotohanto and Nabae Sub-belts (Shiina-Narashi fault), and this contrast is consistent with approximately 1200 m of post-metamorphic vertical offset. Subduction prior to Middle Miocene probably involved the Kula or fused Kula-Pacific plate and the background geothermal gradient during the Eocene-Oligocene phase of accretion was - 30-35"Ckm. Rapid heating of the Shimanto Belt evidently occurred immediately after a Middle Miocene reorganiza- tion of the subduction boundary. Hot oceanic lithosphere from the Shikoku Basin first entered the subduction zone at - 15 Ma; this event also coincided with the opening of the Sea of Japan and the rapid clockwise rotation of southwest Japan. The background geothermal gradient at that time was - 70"Ckm. Whether or not all portions of the inherited (Eocene-Oligocene) palaeothermal structure were overprint- ed during the Middle Miocene remains controversial. Key words: accretionary prism, ikte crystallinity, palaeotemperature, thermal structure, vitrinite reflectance. INTRODUCTION The Shimanto Belt of southwest Japan (Fig. 1) represents the youngest subaerial portion of a long-lived accretionary margin (Taira et al. 1982, 1988, 1989). These strata make up a large part of the Outer Zone of Japan (Hada & Suzuki 1983) and range in age from Cretaceous to Miocene. In general, metamorphic mineral assemblages within the Outer Zone conform to the zeolite and prehnite-pumpellyite facies (Toriumi & Teruya 1988). Conceptual and strati- graphic correlations have been made between the offshore accretionary domain (modern Nankai Trough) and the Shimanto Belt (Taira 1985; Taira et al. 1988). Although the paradigm of an accretionary prism seems quite appropri- ate, the history of interaction between the Japanese margin and subducting plates of the Pacific Basin actually has been quite complicated (Uyeda & Miyashiro 1974; Maruyama et al. 1989; Taira et al. 1989). For example, the Miocene epoch, alone, was punctuated by the cessation of sea floor spreading in the Shikoku Basin, the opening of the Sea of Accepted for publication 21 March 1992 Japan, clockwise rotation of southwest Japan, widespread anomalous near-trench magmatism and incipient collision between the Izu-Bonin Arc and the Honshu Arc (Seno & Maruyama 1984; Niitsuma & Akiba 1985; Otofuji & Mat- suda 1987; Niitsuma 1988). Studies by Mori and Taguchi (1988) showed that the level of thermal maturity displayed by Tertiary strata of the Shimanto Belt is anomalously high when compared with most coeval accretionary terranes of the Pacific Rim, at least those exposed to similar (relatively shallow) depths of burial. For example, the average vitrinite reflectance (%R,) values for Palaeogene pelites within the Franciscan Com- plex of northern California (Yager and Coastal terranes) are only 0.7%R,,,, which equates to an approximate palaeo- temperature of only 110°C (Underwood et al. 1988). Simi- larly, most of the Tertiary rocks exposed on the Kodiak Is., in the western Gulf of Alaska, fall below l.O%R, (Moore & Allwardt 1980). The Tertiary Shimanto, Franciscan and Kodiak sequences lack high-pressure mineral assemblages (such as lawsonite), so burial pressures were evidently

Transcript of Thermal evolution of the Tertiary Shimanto Belt, Muroto Peninsula, Shikoku, Japan

The Island Arc (1992) 1, 116-132

Thematic Article Thermal evolution of the Tertiary Shimanto Belt, Muroto Peninsula,

S hi ko ku, Japan

MICHAEL B. UNDERWOOD, MATTHEW M. LAUGHLAND, * TIM BYRNE,~ J. P. HIBBARD3 AND LEE DITULLIo4

'Department of Geological Sciences, University of Missouri, Columbia, MO 652 1 1, 'Department of Geology and Geophysics, University ofconnecticut, Storrs, CT 06268, 3MEAS, Box 8208, North Carolina State University, Rakigh, NC 27695, USA and

4Department of Geology, Kochi University, Kochi 780 Japan

Abstract The Shimanto accretionary complex on the Muroto Peninsula of Shikoku comprises two major units of Tertiary strata: the Murotohanto Sub-belt (Eocene-Oligocene) and the Nabae Sub-belt (Oligocene-Miocene). Both sub-belts have been affected by thermal overprints following the peak of accretion-related deformation. Palaeotemperatures for the entire Tertiary section range from - 140 to 315"C, based upon mean vitrinite reflectance values of 0.9-5.0%Rm. Values of illite crystallinity index are consistent with conditions of advanced diagenesis and anchimetamorphism. Illite/mica b, lattice dimensions indicate that burial pressures were probably no greater than 2.5 kbar. In general, levels of thermal maturity are higher for the Murotohanto Sub-belt than for the Nabae Sub-belt. The Eocene- Oligocene strata also display a spatial decrease in thermal maturity from south to north and this pattern probably was caused by regional-scale differential uplift following peak heating. Conversely, the palaeothermal structure within the Nabae Sub-belt is fairly uniform, except for the local effects of mafic intrusions at the tip of Cape Muroto. There is a paleotemperature difference of - 90°C across the boundary between the Murotohanto and Nabae Sub-belts (Shiina-Narashi fault), and this contrast is consistent with approximately 1200 m of post-metamorphic vertical offset.

Subduction prior to Middle Miocene probably involved the Kula or fused Kula-Pacific plate and the background geothermal gradient during the Eocene-Oligocene phase of accretion was - 30-35"Ckm. Rapid heating of the Shimanto Belt evidently occurred immediately after a Middle Miocene reorganiza- tion of the subduction boundary. Hot oceanic lithosphere from the Shikoku Basin first entered the subduction zone at - 15 Ma; this event also coincided with the opening of the Sea of Japan and the rapid clockwise rotation of southwest Japan. The background geothermal gradient at that time was - 70"Ckm. Whether or not all portions of the inherited (Eocene-Oligocene) palaeothermal structure were overprint- ed during the Middle Miocene remains controversial.

Key words: accretionary prism, ikte crystallinity, palaeotemperature, thermal structure, vitrinite reflectance.

INTRODUCTION The Shimanto Belt of southwest Japan (Fig. 1) represents the youngest subaerial portion of a long-lived accretionary margin (Taira et al. 1982, 1988, 1989). These strata make up a large part of the Outer Zone of Japan (Hada & Suzuki 1983) and range in age from Cretaceous to Miocene. In general, metamorphic mineral assemblages within the Outer Zone conform to the zeolite and prehnite-pumpellyite facies (Toriumi & Teruya 1988). Conceptual and strati- graphic correlations have been made between the offshore accretionary domain (modern Nankai Trough) and the Shimanto Belt (Taira 1985; Taira et al. 1988). Although the paradigm of an accretionary prism seems quite appropri- ate, the history of interaction between the Japanese margin and subducting plates of the Pacific Basin actually has been quite complicated (Uyeda & Miyashiro 1974; Maruyama et al. 1989; Taira et al. 1989). For example, the Miocene epoch, alone, was punctuated by the cessation of sea floor spreading in the Shikoku Basin, the opening of the Sea of

Accepted for publication 21 March 1992

Japan, clockwise rotation of southwest Japan, widespread anomalous near-trench magmatism and incipient collision between the Izu-Bonin Arc and the Honshu Arc (Seno & Maruyama 1984; Niitsuma & Akiba 1985; Otofuji & Mat- suda 1987; Niitsuma 1988).

Studies by Mori and Taguchi (1988) showed that the level of thermal maturity displayed by Tertiary strata of the Shimanto Belt is anomalously high when compared with most coeval accretionary terranes of the Pacific Rim, at least those exposed to similar (relatively shallow) depths of burial. For example, the average vitrinite reflectance (%R,) values for Palaeogene pelites within the Franciscan Com- plex of northern California (Yager and Coastal terranes) are only 0.7%R,,,, which equates to an approximate palaeo- temperature of only 110°C (Underwood et al. 1988). Simi- larly, most of the Tertiary rocks exposed on the Kodiak Is., in the western Gulf of Alaska, fall below l.O%R, (Moore & Allwardt 1980). The Tertiary Shimanto, Franciscan and Kodiak sequences lack high-pressure mineral assemblages (such as lawsonite), so burial pressures were evidently

Thermal evolution of the Tertiary Shimunto Belt 117

Fig.1 Modern plate setting and major geological elements of southwest Japan. Fossil Shikoku Basin spreading ridge is from Chamot-Rooke et a/. (1987). Heavy arrows indicate approximate relative plate motions. BTL = Butsuzo Tectonic Line (northern boundary of Shimanto Belt). Fine lines on the Philippine Sea plate are 4000 m bathymetric contours. From Hib- bard & Karig (1 990a)

<- 3 kbar. Therefore, geothermal gradients appear to have been higher in southeast Japan during the phase of peak heating, compared with California and Alaska. The cause of this abnormal heating, in a general sense, must be related to the subduction of young oceanic lithosphere (Underwood et al. 1992~) . When examined in detail, however, several unknowns remain in our reconstructions of thermal history, including the absolute timing of the peak heating event(s) and the identity of the lithospheric plate(s) responsible for the thermal overprint.

The purpose of this paper was to summarize the critical observations and interpretations of a multidisciplinary study of the Shimanto Belt on the Muroto Peninsula of Shikoku (Fig. 1). The principal rock units include the Murotohanto Sub-belt (Eocene-Oligocene) and the Nabae Sub-belt (Oligocene-Miocene). The boundary between the two sub-belts is marked by an out-of-sequence thrust, the Shiina-Narashi Fault. Readers should refer to papers cited in Underwood (1992) for in-depth discussions and citations regarding the regional geology, laboratory results and analytical techniques.

SHIMANTO METAMORPHISM

METHODS

Analyses of organic metamorphism on the Muroto Penin- sula (Laughland & Underwood 1992) amplify and confirm the results of previous research by Mori and Taguchi (1988), who covered a much larger portion of the Shimanto Belt at a reconnaissance scale and included Cretaceous rocks located to the north of the present study area (Fig. 2). All of the data reported herein are expressed as values of mean random reflectance in oil (%R,), based on measure-

ments of - 50 individual vitrinite particles per specimen. Palaeotemperature estimates have been calculated using the equation of Barker (1988), which relates vitrinite reflectance to borehole temperature without considering the possible effects of heating duration or rate. This equa- tion is T ("C) = 148 + (104[In %R,]) and the error is - + 30°C. The resulting palaeotemperatures fall within the regional constraints imposed by inorganic mineral phases assigned to the zeolite and prehnite-pumpellyite facies (Toriumi & Teruya 1988). In addition, %R, values have been compared with illite crystallinity (Underwood et al. 199213). The illite crystallinity index (CI) is reported here in units of E 2 8 , and this measure of peak width on X-ray diffractograms decreases with increasing thermal maturity. X-ray diffraction measurements were made after oriented clay-sized aggregates (< 2pm) had been saturated with ethylene glycol. Following the criteria of Blenkinsop (1988), the boundary between advanced diagenesis and anchimeta- morphism (transition into lowermost greenschist facies) corresponds to a CI value of 0.42D28, and the upper limit of the anchizone is set at 0.25U28. Finally, as a constraint on maximum burial pressure, selected samples were ana- lysed for illitdmica b, lattice spacing, using random-powder mounts and an internal quartz standard (Underwood et al. 199213). As described by Sassi and Scolari (1974), the b, dimension increases as a function of burial pressure in response to Mg+Fe substitution in the octahedral site of white K-mica.

Most of our palaeotemperature estimates agree favourably with constraints imposed by the annealing temperatures of apatite and zircon fission tracks; the bulk of the Tertiary Shimanto Belt appears to have been heated below the upper limit of the zircon partial annealing zone. This boundary corresponds to a temperature maximum of - 260"C, assum- ing a relatively modest heating rate of lO"C/m/Ma (Hasebe

118 M. B. Underwood et al.

et ul. 1992). With faster rates of heating and/or smaller durations of effective heating time, this limit to the anneal- ing zone would be higher (Zaun & Wagner 1985; Naeser et al. 1989). It should be emphasized here that the apparent temperature constraint of 260°C is equivalent to a value of %R, of - 3.0%, using the model of organic metamorphism proposed by Barker (1988). Values as high as 3.7%Rm, moreover, still fall within an error bar of k 20°C with respect to the adopted limit of the zircon partial annealing zone. Additional discussion of errors inherent in the palaeo- temperature estimates appears in Laughland and Under- wood (1992).

MUROTOHANTO SUB-BELT

The Murotohanto Sub-belt on the Muroto Peninsula has been subdivided into two structural domains (Shiina and Gyoto), plus three fault-bounded zones of tectonic melange (Fig. 2). DiTullio and Byrne (1990) showed that stage 1 accretion-related deformation features within these rocks were overprinted by younger structures. Intermediate stage 2 deformation, which was caused by intraprism shortening, involved the successive development of: ( 1) seaward-vergent folding and cleavage development within the Shiina domain, in a regime of north-south shortening (with respect to present-day geographic co-ordinates); ( 2 ) out-of-sequence thrusting of the Gyoto domain over the Shiina domain; and (3) folding and cleavage development in the Gyoto domain, in a regime of northwest-southeast shortening (DiTullio 61 Byrne 1990).

Fig. 2 Geological framework of the Ter- tiary Shimanto Belt, Muroto Peninsula, Shikoku. Stratigraphic nomenclature and structural boundaries are based upon field mapping by Hibbard (1 988), DiTullio (1 989), DiTullio and Byrne (1990) and Hibbard eta/. (1992a).

Figure 3 shows that %R, values for the Eocendligocene strata range from 1.4 to 5.0%Rm. According to Barker ( 1988), these data correspond to palaeotemperatures of 180-3 15°C. In terms of regional-scale spatial changes, a limited number of analyses from fault-bounded units in the rugged interior of the peninsula suggest that the interior has been heated slightly more than the coastal sections (Fig. 4). Thermal maturity on the east coast decreases steadily toward the north (Fig. 5 ) . If comparisons are restricted to localities immediately to the north of the Shiina-Narashi Fault, there are no significant differences in thermal maturityfrom the west coast to the east (Fig. 4). On the other hand, farther from the Shiina-Narashi Fault, the west-coast exposures of the Shiina domain are signifi- cantly higher in organic rank than comparable rocks of the Shiina domain on the east coast (Fig. 4).

The intensity of cleavage development in the Shiina and Gyoto domains clearly correlates, in a spatial sense, with organic metamorphism. In particular, the highest-rank rocks located just north of the Shiina-Narashi Fault display the best cleavage, so it seems logical to conclude that peak heating of the Murotohanto Sub-belt was contemporaneous with the formation of these fabrics (DiTullio et al. 1992). Figure 6 shows that the isoreflectance surfaces are not folded with the limbs of map-scale F2 folds (for additional examples, see DiTullio et al. 1992), which means that most or all of the levels of organic metamorphism were attained after the culmination of stage 2 deformation. In addition, neither the Shiina domain nor the Gyoto domain south of Cape Hane (west coast) displays consistent tilting or offset

Thermal evolution of the Tertiary Shimanto Belt 119

of isoreflectance surfaces (Fig. 4). This uniformity of %R, values is likewise consistent with a thermal event after the peak in deformation.

The lowest-rank rocks of the Murotohanto Sub-belt coincide with a sandstone-rich sedimentary facies on both the east and west coasts (Figs 2 and 4). Although one might argue that the lowest-rank successions represent part of a slope-apron or slope-basin depositional facies (which would explain the shallow burial depths and lower temperatures), structural analyses of one such example tend to refute that interpretation (DiTullio & Byrne 1990). Regardless of the primary depositional setting, the palaeothermal data cer- tainly support the contention that these sandstone-rich sections remained at the shallowest levels of the accretion- ary prism during the peak heating event. Hasebe et al. (1992) used apatite fission-track data to show that the northern part of the Eocene-Oligocene section passed through the 100°C isotherm at - 14.6Ma, which was earlier than the coeval higher-rank rocks farther to the south (8-1 1 Ma). Presumably, this difference in cooling history is a combined response to contrasts in burial posi- tion at the time of peak heating (that is, deeper to the south), together with differential uplift following organic metamorphism.

Perhaps the most important structural feature within the Murotohanto Sub-belt is the out-of-sequence thrust which places the Gyoto domain over the Shiina domain (Fig. 2). The east half of the Muroto Peninsula displays the same high ranks of organic metamorphism in both the hanging wall and the southern part of the footwall (Fig. 4). The same lack of palaeotemperature contrast between structural domains exists along most of the west coast (Fig. 4). Only at Cape Hane, on the west coast, was a clear offset of the thermal structure across the apparent domain boundary documented; and the Gyoto hanging wall there is actually lower in rank than the Shiina footwall (DiTullio et d. 1992). Several scenarios involving postmetamorphic out-of-sequence thrusting could explain this relationship (e.g., two phases of post-peak thermal faulting, or folding of isoreflectance surfaces followed by thrusting). However, the simplest interpretation calls for uplift of the Shiina domain relative to the Gyoto domain. A late-stage, northeast-striking, high-angle fault (Hane River Fault Zone) has been mapped through this same locality (Taira et al. 1980; Sakai 1987), and this younger fault probably is responsible for a late- stage offset of the palaeothermal structure.

The apatite fission-track data of Hasebe et al. (1992) show that the Shiina and Gyoto domains were uplifted and cooled through the 100°C isotherm at roughly the same time. Some additional deformation and/or differential uplift could have affected the domains after they passed above the 100°C isotherm, but this would be difficult to detect using fission-track techniques. Based upon the available data, we conclude that organic metamorphism in the Murotohanto Sub-belt occurred after the culmination of G y o t d h i i n a out-of-sequence thrusting. Furthermore, thrust displace- ment was not responsible for any detectable offsets of the palaeothermal structure, as defined by surface patterns of %R, contours.

NABAE SUB-BELT

The Nabae Sub-belt (Oligocene-Miocene Shimanto Belt) contains two units of melange (Sakamoto and Hioki), two

1'1 Eocene-Oligocene Shimanto

Murotohanto sub-belt

X = 3.29% 8

, I I I I

160 200 225 250 275 300

? O l 15 -I Oligocene-Miocene Shimanto

Nabae sub-belt

P=1.81% n = 1 0 2

0 - 10

0

5

1.0 1.4 1.8 2.2 2.6 3.0 3.4 3.8 Mean random vitrinite reflectance ("A)

, , 150 175 200 225 250 275

Palaeotemperature ("C)

Fig.3 Histograms of %R, values for the Murotohanto Sub-belt (Eocene- Oligocene) and the Nabae Sub-belt (Oligocene-Miocene), Muroto Peninsula, Shikoku, Japan. Temperature scale is based on the model of Barker (1988).

assemblages of complexly deformed sedimentary rocks (Misaki and Tsuro), inferred slope-basin deposits of the Shijuji- yama Formation and mafic intrusions of the Maruyama igneous suite (Fig. 2). Just as in the Murotohanto Sub-belt to the north, deformed rocks within the Nabae Complex show clear evidence that the early-stage structural fabrics (due to sediment offscraping) were overprinted by an inter- mediate stage of intraprism shortening (Hibbard et al. 1992a). It is important to note, however, that the first two stages of deformation in the Nabae Sub-belt occurred after broadly analogous stage 1 and stage 2 deformations docu- mented in the Murotohanto Sub-belt.

Overall, the Nabae Sub-belt is considerably lower in thermal maturity than the older Murotohanto rocks (Fig. 3). Moreover, there is no obvious relationship between the degree or style of stage 1 stratal disruption and the magni- tude of thermal maturity (Hibbard et al. 199213). The %R, values within most of the Nabae Complex range from 0.9 to 2.1% (Fig. 3) and most of the structural and stratigraphic units display similar levels of thermal maturity (Fig. 4). O n the west coast of Cape Muroto (Fig. 7), two windows of melange contain high-rank anomalies (2.2 and 2.6%Rm) compared with values within the overlying Tsuro Assem- blage (1.4-2.1%Rm). These anomalies probably have been retained from an early phase of hydrothermal activity (Hibbard et al. 199213). Overall, however, the Tsuro sec- tion displays a remarkable degree of uniformity in thermal

120 M. B. Underwood et al.

Shijujiyama Formation

Maruyama instrusive suite

B

Kabuka

" O m North - KABUK! FAULT 5 3

Fig.4 Map of spatial trends in %R, values within the Tertiary Shimanto Belt, Muroto Peninsula, Shikoku, Japan. The boundary between the Murotohanto Sub- belt (M) and the Nabae Sub-belt (N) is marked by the Shiina-Narashi Fault. See Fig. 9 for conversion of these values into estimates of palaeotemperature and Laugh- land and Underwood (1 992) for data tables and sample localities.

Fig.5 Geological map and down-plunge section for the Kabuka segment of the Shiina structural domain, east coast of the Muroto Peninsula (see Fig. 4 for location). Numbers in boxes are values of mean vitrinite reflectance (%R,). Numbers in ovals are values of illite Cl (A"26'). With the exception of the Flat Rock Fault, there are no significant offsets or uniform gradients in thermal maturity. From DiTullioet a/ . (1992).

Thermal evolution of the Tertiary Shimanto Belt 121

Gyoto

It North 500 rn

4 1

Fig. 6 Geological map and cross-section for the Gyoto section, west coast of the Muroto Peninsula (see Fig. 4 for location). Numbers in boxes are values of mean vitrinite reflectance (%R,). Numbers in ovals are values of illite CI (A"20). Note the absence of significant variations in thermal maturity across the fold limbs and across the Moudodanigawa Fault. From DiTullioet a/ . (1992).

maturity across a complicated succession of early-stage and intermediate-stage folds and faults (Fig. 7). This pattern is clearly the result of a late-stage thermal overprint and erosion down to a common level within the palaeothermal structure.

Thermal overprints during late-stage deformation of the Nabae Sub-belt are particularly obvious near the southern tip of Cape Muroto (Fig. 4), where abnormally high %R, values for the Misaki assemblage are genetically related to the emplacement of the Maruyama intrusive suite at - 14Ma (Hamamoto & Sakai 1987). The temperatures of vitrinite equilibration adjacent to the mafic intrusions were 2 285"C, based upon %R, values of 3.7%; as explained by Laughland and Underwood (1992), the maximum wall- rock temperatures probably were considerably higher. Hasebe et al. (1992) showed that detrital zircons in the Misaki sandstone beds were completely reset and then passed through the annealing window immediately after the intru- sion (14.1 Ma cooling age). We believe that peak heating and organic metamorphism throughout the Nabae Sub-belt probably were synchronous with this magmatic activity (Hibbard et al. 1992b). However, the spatial extent of the contact aureole is restricted to roughly twice the outcrop width of the Maruyama intrusion (Laughland & Underwood 1992).

At first glance, one might expect the Shijujiyama basin- fill to be lower in thermal maturity than the adjacent

melange, based on models of trench-slope basins (e.g. Moore & Karig 1976). However, this anticipated palaeo- thermal contrast with respect to the underlying Hioki melange (inferred accretionary basement) cannot he sub- stantiated. In fact, %R, values for several of the melange sites, located within only a few hundred metres of the contact, are significantly lower (down to 0.9%Rm) than those of the Shijujiyama Formation, which are consistently - 1.6%R, (Fig. 4). We believe this is because the primary tectonostratigraphic architecture (slope basin over melange basement) has been structurally overprinted. The slope- basin hypothesis, therefore, must include some element of fault dislocation along the Shijujiyama-Hioki contact. One possibility is that melange-cored ridges bordering the slope basin were uplifted sufficiently along thrust faults prior to peak heating; this syndepositional uplift could have prevented resetting of %R, values even if the regional geothermal gradient subsequently increased and subsurface isotherms rose higher in the structural section. Alternatively, vertical offset along the slope-basin margin might have occurred after peak heating, but then the Shijujiyama strata would have to be uplifted with respect to the Hioki melange.

ESTIMATES OF GEOTHERMAL GRADIENT

As discussed by Underwood et al. (1992b), the intensely cleaved rocks of the Murotohanto Sub-belt are the only ones to have reached high enough burial temperatures to yield reliable values of illite/mica b, lattice spacing. Lower rank strata in the Nabae Sub-belt are questionable, in this regard, because of the likelihood of contamination of the authigenic signal by detrital white K-mica. Most of the illite b , values range from 9.010 to 9.020A (Fig. 8). Based upon comparisons with other regions of low-temperature metamorphism (e.g. Sassi & Scolari 1974; Padan et al. 1982; Guidotti & Sassi 1986), the Shimanto data are consistent with modest amounts of phengitic substitution (i.e. increases of octahedral Fe + Mg in the muscovite- celadonite series). Burial conditions correspond to the intermediate-pressure facies series, and the strata probably were not exposed to pressures >- 2.5 kbar. Assuming a mean bulk density of 2.60g/cm3, this limit on burial pressure equates to a maximum depth of - 9 km. If we divide the maximum palaeotemperature values within the Murotohanto Sub-belt by this depth limit (and assume a minimum sea floor temperature of roughly VC), then the geothermal gradient at the time of organic metamorphism equates to 30-35"Ckm.

Our estimate of the geothermal gradient during peak heating of the Nabae Sub-belt is based upon several argu- ments and empirical comparisons presented by Hibbard et al. (199213). The gradient within inferred slope-basin deposits of the Shijujiyama Formation was probably - 70"Ckm at the time of peak heating, which is qualitatively consistent with the existence of interfingered mafic volcanic rocks and siliciclastic mudstones near the base of the Shijujiyama section (Hibbard & Karig 1990a). Palaeotemperature gradients calculated for other rock units within the Nabae Complex are conservatively estimated to range from 60 to 90'Ckm (Hibbard et al. 1992b).

SHIINA-NARASHI FAULT

The most conspicuous discontinuity in palaeotemperature on the Muroto Peninsula occurs across the Shiina-Narashi

122 M. B. Underwood et al.

TSURO -T-

WABAE N A B 3 1

+- younging directions in strata Tsuro assemblage Nabae Complex melange mainly alternating sandstone mudstone

mainly hempelagic rocks

,,/ decollsment beneath the Tsuro assemblage

0 100 m - Fig. 7 Detailed schematic cross-section of the Tsuro assemblage between Tsuro and Nabae (see Fig. 4 for location). Numbers refer to values of mean vitrinite reflectance. Note the lack of correspondence between %R, values and intermediate-stage deformation features; this is indicative of a later-stage thermal overprint. Note also the two anomalously high values from windows of melanse (2.2 and 2.6%Rm). Both anomalous sites are associated with early-stage fluidized sediment-injection features. From Hibbard eta/. (l992b)

Mica bo lattice spacing ( A )

Fig.8 Cumulative frequency curve of illite/mica b, data from the Tertiary Shimanto Belt, Muroto Peninsula, Shikoku, Japan (heavy line and open dots). Also shown are reference curves for crystalline K-mica (from Sassi & Scolari 1974) for the Bosost facies series (high-temperature, low-pressure), northern New Hamp- shire (low-pressure, intermediate-temperature), the Otago schist (Barrovian-type metamorphism), and the Sanbagawa Belt of Japan (high-pressure, glauco- phane-greenschist facies series). See Padan et a / . (1 982) for additional reference curves.

Fault (Fig. 9). Eocene-Oligocene (Murotohanto) rocks immediately to the north of the fault display better cleavage and yield %R,, values that average - 3.7% (- 285"C), whereas weakly cleaved and uncleaved rocks to the south (Hioki melange and Shijujiyama Formation) average only 1.6%Rm (- 195°C). This difference in organic metamor- phism corresponds to an increase in palaeotemperature of - 90°C.

We acknowledge that displacement along the Shiina- Narashi Fault may have started prior to the culmination of peak heating in the Nabae footwall. The fault truncation and vertical offset of isoreflectance surfaces, however, requires substantial postmetamorphic displacement. Furthermore, there is no evidence of anomalously high %R, values immediately to the south of the fault, which might be expected if there had been conductive transfer of heat from the hanging wall into a footwall aureole (e.g. Brewer 1981; Furlong & Edman 1984; Shi 6 Wang 1987). Consequently, either the hanging wall was relatively cool at the time of thrusting or fault-zone temperatures were moderated by high rates of hanging-wall erosion.

It seems reasonable to conclude that the apatite' cooling ages (8-1 1 Ma) documented for the Eocene-Oligocene rocks by Hasebe et al. (1992) are a response to the uplift that was accommodated by the Shiina-Narashi Fault. Given the evidence for neotectonic activity on the Muroto Penin- sula, including the exposure and deformation of Upper Pliocene marine sediments and the formation of prominent marine terraces (Sakai 1987), there is also a strong possibil- ity of protracted offset through much of the Late Cenozoic. This contention of late-stage displacements within the Shimanto Belt is supported by the postmetamorphic slip along the Hane River Fault Zone, which is superimposed on the Gyote-Shiina out-of-sequence thrust south of Cape Hane (Fig. 9). For both faults (Hane River and Shiina- Narashi), the structural levels observed today (on opposite sides of the fault) were juxtaposed after both fault blocks had reached their maximum burial temperatures.

If one accepts the idea that the regional geothermal gradient was as high as 70°C/km during the Middle Miocene, and that both the Murotohanto and Nabae Sub-belts attained their maximum burial temperatures within that thermal regime, then the minimum amount of postmeta- morphic vertical offset on the Shiina-Narashi Fault equates to - 1200 m. In other words, this amount of slip is required to account for the contrast in vitrinite reflectance. As discussed later, however, peak heating of the Murotohanto section could have occurred during the Oligocene under the influence of a lower geothermal gradient. If so, then the

Thermal evolution of the Tertiary Shimanto Belt 123

0 kilometres 4 - - I

Fig. 9 Distribution of palaeotemperature values (T) in the Tertiaty Shimanto Belt, Muroto Peninsula, Shikoku. See Fig. 2 for identification of tectonostratigraphic units. The equation [T("C) = 104 (In %R,) + 1481 was used to calculate palaeotem- perature based on values of mean random vitrinite reflectance (from Barker 1988).

PHILIPPINE

cumulative vertical offset along the Shiina-Narashi Fault may be significantly greater than our minimum estimate of 1200 m.

MUROTO FLEXURE The Muroto Flexure is a major north-trending cross-fold that affects all of the strata analysed on the Muroto Peninsula (Hibbard & Karig 1990a). Within the EoceneOligocene part of the section, the structure appears to maintain a subhorizontal axis (DiTullio & Byrne 1990). Based on mapping within the Nabae Sub-belt, Hibbard and Karig (1990a) concluded that the fold axis is steeper towards the south, where the limbs also tighten. The flexure may be responsible for the pronounced bend which appears in the surface trace of the Shiina-Narashi Fault (Fig. 2), assuming that the fault trace was originally straight. In addition, there is widespread evidence of quartz-vein precipitation and displacement on a system of high-angle, mesoscale faults and tension fractures during the same stage of regional deformation (Hibbard & Karig 1990a).

Thermal maturity data do not define a monotonic post- metamorphic warping of palaeotemperature gradients across the core of the fold or its limbs. For example, %R, values within the Nabae Sub-belt are actually somewhat higher on the west coast (on the limb of the flexure) than on the east coast near the core (Hibbard et al. 1992b); this spatial change in thermal maturity is opposite to that expected for a postmetamorphic anticline, particularly if the plunge of the axial trace is truly subhorizontal. In addition, the

uniformity of palaeotemperature values along an east-west transect just north of the Shiina-Narashi Fault suggest little if any modification by the flexure (Fig. 9). On the other hand, farther to the north, especially on the east coast within the Shiina domain, the gradual northward reduction of thermal maturity values can be broken into subtle, step- like shifts; this pattern has been explained through small amounts of offset along high-angle tear faults on the limbs of the Muroto Flexure (DiTullio et al. 1992). Some of the fault-bounded blocks display fairly uniform levels of ther- mal maturity within, whereas others display an obvious tilting of isoreflectance surfaces (Figs 5, 9). In most cases, the offsets of organic rank across the faults are subtle, at best. One exception occurs across the Flat Rock Fault, where values change from 1.5-1.8 to 2.3-2.9%Rm (Fig. 5). Finally, if the Shiina rocks south of Cape Hane (on the west coast and in the interior) are compared with rocks on the east coast at equivalent distances from the Shiina- Narashi Fault, there is an obvious contrast in thermal maturity (Fig. 9). DiTullio et al. (1992) attributed this spatial change to an important north-south high-angle fault (east-side down); this fault is located near the axis of, and presumably is related to, the Muroto Flexure (Fig. 9).

In conclusion, the Muroto Flexure exerts neither consist- ent nor dramatic control on the orientation of isoreflec- tance surfaces within the Tertiary Shimanto Belt, even though the fold appears to be post-peak thermal. One explanation is that the magnitude of vertical offset of isoreflectance surfaces, from the core of the flexure to either limb, was modest, particularly close to and south of

124 M. B. Underwood et al.

the Shiina-Narashi Fault. With a subvertical fold axis, we would expect the east-west warping of palaeotemperature horizons to be negligible. Another possibility, at least within the Nabae Sub-belt, is that the flexure formed during peak heating (Hibbard et al. 1992b).

TEMPERATURE CALIBRATION OF THE ANCHIZONE DiTullio and Hada (1992), DiTullio et al. (1992) and Underwood et al. (199211) used measurements of the illite CI to assist in our regional and local assessments of thermal history in the Shimanto Belt. Analyses of vitrinite reflectance prove that the Shimanto strata on the Muroto Peninsula have been subjected to a wide range of palaeotemperatures and, in general, the values of CI confirm these results (Figs 5, 6). On the other hand, the statistical match between %R,,, and CI data displays a degree of scatter that cannot be explained entirely in terms of the errors inherent in each analytical technique (Fig. 10). This imperfect match is consistent with numerous studies that have demonstrated that illitic reactions are affected by a much more extensive group of internal and external variables (in addition to maximum temperature), including the original bulk-rock and clay-mineral compositions, heating rate and duration, fluid compositions and rates of fluid flux, the content of organic matter and tectonic stress (e.g. Kisch 1983, 1987; Kemp et al. 1985; Merriman & Roberts 1985; Frey 1987; Yang & Hesse 1991).

In the case of the Shimanto Belt, most of the internal controls that affect the crystallization of illite and white K-mica appear to have been relatively constant. For example, there are no obvious differences in bulk-rock composition among the rock units studied. Because of this, the correla- tion coefficient ( r = 0.84) for the logarithmic regression line shown in Fig. 10 is significantly larger than comparable values associated with other correlations of this type (e.g. Guthrie et al. 1986; Underwood et al. 1992a). In general, pelitic rocks with better cleavage (southern part of the Murotohanto Sub-belt) yield lower values of CI, as do samples from the Misaki assemblages within the contact aureole of the Maruyama intrusive suite (Underwood et al. 199213). Because vein sets have not been analysed, we have no means of assessing the possible impact of variations in hydrogeological parameters.

In terms of CI values, the established boundaries for the anchizone have been set at 0.42 and 0.25P26 (Blenkinsop 1988; Kisch 1990; Robinson et d. 1990). The regression curve for the Shimanto Belt intersects the anchizone domain at %R, values of 2.8 and 4.2, respectively (Fig. 10). Assuming that the palaeotemperature conversion method of Barker ( 1988) is accurate, the temperature intersections with the anchizone equate to -255 and 300°C. When considered as a whole, the Shimanto curve conforms rea- sonably well to the global norm for conditions of anchi- metamorphism, as summarized by Kisch (1987), and it follows the same general trend as curves from other circum- Pacific orogenic belts (Underwood et al. 1991).

DISCUSSION

REGIONAL THERMAL HISTORY

As reviewed by Underwood et al. (1992c), there are many critical observations and data sets from elsewhere in the

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 l l l i te c rys ta l l i n i t y i n d e x (A"28)

Fig. 10 Statistical correlation between values of mean vitrinite reflectance (%R,) and values of illite CI, Tertiary Shimanto Belt, Muroto Peninsula, Shikoku, Japan. The correlation coefficient (r) for a best-fit logarithmic regression is 0.84. Boundaries for the anchizone are based on Blenkinsop (1 988) and Underwood et a / , (1 991). See Underwoodet a/ . (1 992b) for a complete listing data and discussion of statistical trends.

Japanese islands to consider during any interpretation of results from the Muroto Peninsula. For example, Agar et al. (1989) dated early (accretion-related) cleavages within Shimanto strata on the Hata Peninsula of Shikoku, using K-Ar radiometric methods; their dates for Eocene-Oligocene rocks range from - 43 to 18 Ma (Fig. 11) and most values cluster between 34 and 26 Ma. In addition, for strata that were heated above the apatite partial annealing zone (- 125"C), the resultant fission-track cooling ages for both Late Cretaceous and Palaeogene strata range from - 11 to 9Ma (Fig. 11). Igneous activity on the Hata Peninsula (Ashizuri granites) has been dated at - 13 Ma (Shibata & Nozawa 1967; Oba, 1977), and the emplacement of these granitic intrusions must have coincided, at least locally, with organic metamorphism and recrystallization of illite/ mica (DiTullio & Hada 1992).

K-Ar dates for Shimanto cleavage in eastern Kyushu (Fig. 11) are - 48Ma, which is < lOMa after the fossil- controlled age of deposition (Mackenzie et d. 1990). Values of illite crystallinity for these rocks are generally consistent with conditions of advanced diagenesis and anchizone metamorphism (DiTullio &a Hada 1992). Just as in the Shimanto Belt of southern Shikoku, the rocks on Kyushu were affected by intrusive activity during a time interval of 13-15Ma (Shibata & Nozawa 1967; Oba 1977). The Palaeogene strata there yield %R, values as high as 4.0% (- 290°C); vertical maturation gradients are in the order of 0. l%R,/100 m, and spatial patterns of organic metamorph- ism clearly overprint the structural architecture of the accreted section (Aihara 1989).

On the Kii Peninsula of Honshu (Fig. l l ) , there are pronounced angular unconformities between the intensely deformed Shimanto strata and overlying slope-basin and fore-arc-basin deposits (Chijiwa 1988; Hisatomi 1988; Kumon et ul. 1988). Cleavage within the accreted strata evidently formed prior to the development of the unconfor- mities, because overlying strata (of Middle Miocene or Late Miocene age) are not cleaved; however, we do not have any data to establish the absolute age of cleavage in this region. On the other hand, maximum burial temperatures

Thermal evolution of the Tertimy Shimanto Belt 125

A Shimanto cleavage age (Ma) rl

0

0

Acidic igneous body (radiometric age in Ma) Basic igneous body (radiometric age in Ma) Setouchi volcanic rocks (12-14 Ma) - a Shimanto Belt

100 km

(13- 14)

@ - 30°N

Fig.11 Map showing the distributions of anomalous basic and acidic igneous bodies of the Middle Miocene within the Outer Zone of Japan, together with complementary radiometric age dates (Ma). Occurrences of the somewhat younger Setouchi volcanic series are shown north of the Median Tectonic Line (i.e. the northern limit of the Outer Zone). Also depicted are control points and corresponding absolute K-Ar ages (Ma) of cleavage formation within the Shimanto Belt, plus uplift/ cooling ages (Ma) based on apatite fission tracks. Data for this figure were obtained from the following sources: Shibata and Nozawa (1 967); Oba (1 977); Miyake (1 985); Hamamoto and Sakai (1987); Terakadoet a / . (1988); Agareta/. (1989); Hibbard and Karig (1990b); MacKenzieet a/ . (1990).

over much of the Kii Peninsula must have coincided with the emplacement of plutonic and volcanic rocks at 14- 15 Ma (Shibata & Nozawa 1967; Oba 1977; Miyake 1988; Terakado et al. 1988). This contention is supported by the %R, values documented by Chijiwa (1988) for the Miocene fore-arc-basin succession, which are as high as 6.0%Rm,,. In addition, the ranks of fore-arc-basin coals overlap data from dispersed organic matter within the underlying Shimanto rocks, which are up to nearly 7.0%Rm,, in the subsurface (Aihara 1989).

Collectively, the data cited above show that the entire Outer Zone of southwest Japan was affected by an unusual episode of high geothermal gradients and near-trench mag- matic activity during Middle Miocene time. One of the major disagreements in the interpretation of palaeothermal data is whether or not palaeotemperature gradients inherited from the Eocene-Oligocene phase of plate convergence were overprinted by the effects of the Middle Miocene magmatic activity and associated heat flux through the Shimanto accretionary prism. To address this issue, we must first look at the options for Tertiary plate-tectonic reconstructions of the western Pacific Basin.

PALAEOGENE PLATE MODELS

Interpretations of the Eocene-Oligocene thermal history are tenuous, in part because most of the sea floor record of earlier plate motions has been lost to subduction. Among the recognized complications, events in the Late Eocene and the Oligocene took place before the Middle Miocene clockwise rotation of southwest Japan (Otofuji & Matsuda 1984; Hayashida 1986; Nakajima & Hirooka 1986). In addition, the position of the Kula-Pacific spreading ridge prior to the cessation of sea-floor spreading remains unknown (Fig. 12). Most plate reconstructions locate the Kula-Pacific boundary well to the north of the Japanese islands at - 43 Ma (e.g. Engebretson et al. 1985; Lonsdale 1988; Jolivet et al. 1989), but DiTullio (1989) and Byrne and Di Tullio (1992) speculated that the triple junction may have been close to the Shimanto Subduction Zone (Fig. 12). We simply do not know which of several possible plates (Pacific, Kula, western Philippine or fused Kula-Pacific) was sub- ducted beneath Shikoku during the latest Eocene and Early Oligocene.

An unresolved issue within the context of our study is

126 M. €3. Underwood et al.

whether or not the Early Oligocene epoch was a time of relatively high heat flow, and if so, what was the source of the heat that was transferred through the Murotohanto accretionary prism? One line of evidence favouring the subduction of relatively young (and warm) lithosphere towards the end of the Eocene is the small time gap between turbidite fossil ages and the ages of basalts within the Murotohanto Sub-belt (Taira 1985). These temporal gaps by themselves, however, do not help us determine which plate was subducted. According to Sakai (1988), the rocks at Cape Gyoto contain microfossils whose range extends into Early Oligocene (perhaps as young as 34Ma); fossils as young as 33 Ma also have been identified within the Oligocene Shimanto section of the Hata Peninsula (Agar et al. 1989). The minimum depositional age for this portion of the Shimanto Belt, therefore, seems to be significantly younger than the plate reorganization de- scribed by Lonsdale (1988). Thus, even if the Kula-Pacific ridge had been in close proximity to southwest Japan when spreading ceased at 43 Ma (Fig. 12), the subsequent Oligo- cene subduction history must have involved the fused Kula-Pacific crust, and the slab age must have been at least 10 million years at the time of accretion of the youngest Murotohanto sediment. Nevertheless, by analogy with the present-day Nankai accretionary prism (e.g. Yamano et al. 1984; Kinoshita & Yamano 1986), a slab age of 10-15 million years should have been young enough to produce a background geothermal gradient well in excess of 3O0C/km, particularly near the toe of the prism.

THERMAL OVERPRINT OF PALAEOGENE CLEAVAGE?

DiTullio and Byrne (1990) contend that the observed divergence in cleavage orientations within the Gyoto and Shiina domains is related to a change in shortening direction (from north-south to north-southeast), which occurred during their intermediate stage of out-of-sequence thrusting (Fig. 13). DiTullio and Byrne (1990) attributed the rota- tion of shortening directions to the reorganization in plate motions at - 43 Ma (i.e. Kula subduction followed by fused Kula-Pacific subduction). In other words, according to this model, the D, component of stage 2 deformation began during the late Middle Eocene and the DZ component extended from 43 Ma into the Early Oligocene.

(bl - 4 0 Ma

GYOTO sNF 14 Ma (a .

Fig. 13 Schematic illustration of the tectonostratigraphic evolution of the Tertiary Shimanto Belt on the Muroto Peninsula of Muroto. OST = out-of- sequence thrust (Shiina-Gyoto domain boundary); SNF = Shiina-Narashi Fault (Murotohanto-Nabae Sub-belt boundary). Based largely upon the structural interpretations of DiTullio and Byrne (1 990) and Hibbard et a / . (1 992a).

Fig.12 Comparison of plate tectonic reconstructions for the north Pacific basin and southwest Japan during the Late Eocene (a) From Lonsdale (1988), with the reconstruction set at 56.5 Ma. An earlier configuration of the Kula- Pacific spreading ridge (at 57 Ma) is shown by the dashed line. Fusion of the Kula-Pacific boundary occurred at 44-43 Ma. (b) From Jolivet et al. (1989), is set at 43 Ma and shows subduction of the Pacific plate beneath Japan. (c) From DiTullio (1989), shows segments of the Kula-Pacific ridge (at 45 Ma) in close proximity to the Shimanto subduction front just prior to the demise of sea floor spreading.

Thermal evolution of the Tertiary Shimanto Belt 127

None of the cleavages in the Murotohanto Sub-belt has been dated using radiometric methods. Because the rocks with better cleavage also yield the highest values of %R,, it stands to reason that the cleavages formed during phases of peak heating (DiTullio et al. 1992). Nevertheless, to retain the Eocene and/or Early Oligocene thermal structure (im- parted, for example, by a geothermal gradient of - 30"C/ km), the rocks must have been uplifted considerably with respect to the sea floor prior to the Middle Miocene. Otherwise, palaeotemperature values at a given structural position within the Shimanto prism would have been reset as the geothermal gradients increased across the strike- length of the accretionary margin. The key to this relation- ship, for any particular unit-volume of rock, is the absolute magnitude of uplift and erosion of overburden balanced against the absolute vertical shift in isotherms. Spatial overlap of the temporal palaeothermal domains certainly is possible. However, as discussed by Underwood et al. (1992d), the magnitude of uplift and erosion would have to be > 3 km to prevent a thermal overprint of a typical parcel of Murotohanto strata.

The best evidence for regional uplift prior to the Middle Miocene, as summarized by Sakai (1988) and Hibbard and Karig (1990b), comes from the unconformities that sep- arate the Shimanto Belt from overlying fore-arc-basin successions, such as the Kumano Group (Kii Peninsula), the Shijujiyama Formation (Muroto Peninsula) and the Misaki Group (Hata Peninsula). We believe, however, that these contacts are submarine unconformities, particu- larly because the fore-arc and slope-basin successions grade up-section from basal units that were deposited at middle bathyal depths into shallow-water facies (Kimura 1985; Chijiwa 1988; Hisatomi 1988; Osozawa 1988). Submarine uplift of the Eocene-Oligocene accretionary prism almost certainly did occur, perhaps in response to underplating of the Oligocene-Miocene sediments (Fig. 13), but the removal of overburden was not sufficient to keep pace with the rise in isotherms at 15 Ma.

One alternative interpretation worth considering is that organic metamorphism and D2 cleavage in the Muroto- hanto Sub-belt were, indeed, synchronous, but they oc- curred during the Middle Miocene rather than the Early Oligocene. If so, then the change in shortening direction documented by DiTullio and Byrne (1990) may be a response to prerotation versus post-rotation convergence, combined with a shift from Eurasia-Pacific subduction to Eurasia-Philippine Sea subduction at - 15 Ma. According to Otofuji and Matsuda (1987), the clockwise sense and magnitude of block rotation (- 50") are perfectly consist- ent with the angular deflection in the fabric elements (DiTullio & Byme 1990). As a final possibility, the D1 and D2 cleavages may have formed during the Eocene-Oligocene, but ranks of organic metamorphism were still overprinted during the Middle Miocene without forming a cross-cutting set of cleavage.

Two lines of evidence refute the idea of Miocene cleav- age formation within the Murotohanto Sub-belt. First, the S2 cleavages in the Eocene-Oligocene strata are discordant to Miocene fabrics in the Nabae Sub-belt, and the respect- ive deformation histories of the two sub-belts cannot be correlated. Second, the K-Ar dates for composite illite populations on the Hata Peninsula seem to be 28-3Ma older than the onset of Outer Zone magmatic activity (Agar et al. 1989). One pitfall associated with radiometric dating of illite/mica in low-grade pelitic rocks is the pos-

sibility of mixing among detrital populations and one or more populations of authigenic illite and/or K-mica. The critical temperature required to reset argon completely in these minerals is - 260 k 30°C (Hunziker et al. 1986). Several variables control the resulting K-Ar dates, includ- ing the particle size, detrital provenance, the relative proportions of 2 MI and 1 Md mica polymorphs, and burial temperature (Hunziker et al. 1986). In response to these factors, inverse correlations commonly exist between illite K-Ar ages and metamorphic grade (i.e. higher-grade rocks yield younger cleavage ages). In other words, composite- particle K-Ar ages from the Shimanto Belt do not neces- sarily represent the culmination of peak heating unless all of the crystals were heated beyond the illite/muscovite blocking temperature. This important boundary condition definitely did not exist within the study area of Agar et al. (1989), as they intentionally avoided the highest-grade zones closer to the Ashizuri granitic intrusions (Fig. 11); in fact, many of their specimens did not even pass through the temperature window for total annealing of apatite fission tracks (- 125°C). In conclusion, without unequivocal iso- lation of the youngest population of authigenic illite, the K-Ar results only show that the cleavage on the Hata Peninsula (and, by analogy, the Muroto Peninsula) is no older than 43-18 Ma.

OLIGOCENE-MIOCENE PLATE MODELS

At the present time, subduction of the Shikoku Basin lithosphere along the Nankai Trough involves the Eurasian and Philippine Sea plates (Ranken et al. 1984; Seno & Maruyama 1984). The initiation of rifting and sea floor spreading in the Shikoku Basin cannot be dated directly through analysis of sea floor magnetic anomalies, because the older intervals of oceanic crust have been consumed. The oldest observable anomaly in Shikoku Basin is anomaly 7, which has been dated at 26 Ma (Watts & Weissel 1975). Several phases of spreading and realignment of the ridge can be recognized and the last occurrence of volcanic activity evidently took place between 14 and 12Ma (Chamot- Rooke et al. 1987).

According to the retreating trench model of Seno and Maruyama ( 1984), subduction beneath Nankai Trough during Late Oligocene (30 Ma) involved older lithosphere of the Pacific plate (Fig. 14). Westward migration of the Izu-Bonin Arc-Trench System eventually brought the cen- tral ridge of the Shikoku Basin into a position almost directly offshore of the Muroto Peninsula by - 17-15 Ma. A model by Otsuki (1990) is similar except that the central ridge is placed farther to the east in the Middle Miocene and then migrated to the west.

An important modification of the Seno and Maruyama (1984) plate reconstruction was proposed recently by Hib- bard and Karig (1990b). The critical difference is the addition of a transform boundary between the Pacific and Philippine Sea plates, together with a protrusion of older Pacific crust separating the Shimanto subduction front from the active Shikoku Basin spreading ridge (Fig. 14). This configuration is very appealing because it allows for a lower geothermal gradient through the Oligocene and Early Mio- cene, with a rapid transition to higher geothermal gradients when the Shikoku Basin lithosphere entered the subduction zone. The radiometric dates summarized on Fig. 11 prove that Miocene magmatic activity did not sweep gradually across the Shimanto fore-arc in the wake of a migrating

128 M. 8. Underwoodet al.

w I I

Fig. 14 Plate tectonic reconstructions for southwest Japan during the Oligocene, earliest Miocene and Middle Miocene. PAC = Pacific plate. Dotted line in the Shikoku Basin represents the approximate limit of oceanic crust that has been subducted since 14 Ma. Arrows indicate directions of motion for the Pacific and Philippine Sea plates relative to Eurasia. For the 24 Ma reconstruction, an allowance has been made for opening of the Sea of Japan and 50" of clockwise block rotation. At 14 Ma, the locus of collision between the central ridge of the Shikoku Basin and the subduction front is placed near the Kii Peninsula of Honshu. The 24 and 1 4 Ma reconstructions are modified from Hibbard and Karig (1990b). The 30 Ma reconstruction is modified from Sen0 and Maruyama (1984) and does not take into account the block rotations associated with opening of the Sea of Japan.

Shikoku Basin spreading centre, as would be expected with a migrating triple junction (e.g. Marshak & Karig 1977). Instead, high regional heat flow and widespread near- trench magmatism began suddenly across the entire fore- arc, within a time window of - 15-13Ma. A transform boundary might also account for the inferences of wide- spread slumping and olistostromes along the margin of southwest Japan during the same time interval (Sakai 1988).

Palaeomagnetic evidence shows that the Outer Zone magmatic activity occurred immediately after the culmina- tion of 42-56" of rapid clockwise rotation of southwest Japan, which itself was caused by the opening of the Sea of Japan (Otofuji & Matsuda 1987). We believe that the thermal regime throughout the Shimanto Belt changed abruptly at this same time because of a shift from subduc- tion of Pacific plate to subduction of very young and unusually hot Philippine Sea plate (Fig. 14). Immediately after the plate reorganization, Early Miocene strata within the accretionary wedge would have been very close to the subduction front, so background geothermal gradients of 70"C/km, or more, certainly would be expected (see, for a modem analogue, Cande et al. 1987). Even higher gradients must have developed across the Outer Zone within the contact aureoles of granitic and mafic intrusions such as the Maruyama gabbros (Fig. 13).

LOCUS OF RIDGE COLLISION

According to Hibbard and Karig (1990a), the Muroto Flexure formed soon after the 15 Ma reorganization of the subduction boundary, in a mechanical response to collision between the Shimanto accretionary prism and a rigid indenter of significant size. Although the indenter may have been the central ridge of the Shikoku Basin (which would also explain the occurrence of cogenetic mafic mag- matism), active seamounts located off the spreading axis

could have accomplished the same thing in terms of both deformation style and magmatic activity. The Muroto Flexure is only one of several so-called 'megakinks' mapped within the Outer Zone of southwest Japan; superficially similar structures have been documented in Kyushu, the Hata Peninsula and the Kii Peninsula (e.g. Kano et al. 1990), where they have been attributed to horizontal compression at shallow crustal levels during opening of the Sea of Japan. Future investigations might test whether or not these other cross-folds and flexures display unusual gradients in thermal maturity, but at the present time, the combination of structural geometry and coeval mafic mag- matism does set the Muroto Flexure apart from the other megafolds.

Based solely on regional trends in thermal maturity, one might argue that the locus of collision between the Shikoku Basin ridge crest and the Shimanto subduction front was located to the east of Cape Muroto, in the general vicinity of the Kii Peninsula (Fig. 11). Our provisional identifica- tion of this 'hot spot' is based on three lines of evidence: (1) values of organic metamorphism within the Tertiary Shimanto section (Aihara 1989); (2) coal rank within the Miocene Kumano Group (Chijiwa 1988); and ( 3 ) the widespread distribution of both extrusive and intrusive magma bodies. The remarkably high levels of coal rank on the Kii Peninsula cannot be dismissed to comparatively larger amounts of uplift and erosion into deeper stratigra- phic levels of the accretionary fore-arc; this is because the coal-bearing successions of the Kumano Group were depos- ited within shallow-marine and non-marine facies, which should be the first deposits lost to subaerial erosion. Accord- ing to Chijiwa (1988), the Middle Miocene palaeogeothermal gradient was 80-1 10"Ckm in the Kumano Fore-arc Basin, but we view these numbers as conservative estimates be- cause they are based on an outdated time-dependent model of thermal maturation. If the Barker (1988) model of

Thermal evolution of the Tertiary Shimanto Belt 129

vitrinite maturation is applied to the data instead, then an estimate for a typical geothermal gradient increases to - 140°C/km, which is substantially higher than compa- rable estimates made by Hibbard et al. (199213) for the Shijujiyama Formation on the Muroto Peninsula.

Another factor to consider is the volume and composi- tion of Miocene igneous bodies within the Outer Zone. Compared with the Maruyama intrusive suite at Cape Muroto, the Kii Peninsula igneous rocks are much more voluminous and chemically more diverse, ranging from MORB-type ophiolites to high-magnesian andesite, rhyo- litic tuff and S-type granite (Murata 1984; Miyake 1985, 1988; Torii & Ishikawa 1986; Terakado et al. 1988). As with other granitoid bodies of the Outer Zone, chemical data, occurrences of xenoliths and zones of migmatite suggest that the magmas probably originated deep in the crust as upwelling mantle convection cells passed beneath the Nankai accretionary prism (Oba 1977; Nakada & Takahashi 1979; Shibata & Ishihara 1979; Nakada 1983; Terakado et aE. 1988). Mixing between igneous and sedi- mentary components evidently occurred deep in the source region, prior to chemical differentiation of the magma (Terakado et al. 1988). In comparison, the flux of mantle heat beneath the Muroto Peninsula evidently was not sufficient to melt the lower crust. Instead, the MORB-type bodies of the Maruyama intrusive suite (Fig. 13) probably were injected into the accretionary prism directly from the upper mantle (Hibbard & Karig 1990a).

To link all of the Miocene igneous bodies of the Outer Zone to a single point of ridge-trench intersection leads to an enigma because magmatism was nearly simultaneous along a strike-length of 700km (Fig. 11). One way to widen the spatial extent of magmatism within a narrow time window (15-12 Ma) is to allow for several east-west transform offsets of spreading segments in the Shikoku Basin. For example, one ridge segment may have collided with the accretionary prism at the Kii Peninsula at 15 Ma, with a second segment providing an intersection point at Cape Muroto at 14Ma. A better explanation, perhaps, involves plate instability and disorganization of the mag- matic locus as the spreading ridge first approached the trench. Chamot-Rooke et al. (1987) speculated that a phase of north-south spreading occurred between 15 and 12 Ma, in response to either a simple rotation of the ridge axis or a ridge propagatiodjumping process. We suggest that protracted adjustments of an unstable spreading ridge, complete with extensive off-axis magmatism, provide the most compelling explanation for what must have been an unusually large zone of mantle upwelling and heat flux into the Shimanto accretionary prism.

CONCLUSIONS Low-temperature, high-pressure metamorphic conditions, which typify blueschist-facies subduction complexes, obvi- ously did not develop along the margin of southwest Japan during the Late Cenozoic. Comparisons among regional data sets from the Outer Zone of Japan show that palaeo- temperature patterns are fairly consistent for hundreds of kilometres along the strike. Our results from the Muroto Peninsula of Shikoku are representative of the Tertiary accretionary prism in general; we have not documented an along-strike anomaly.

Values of mean vitrinite reflectance for strata within the

Murotohanto Sub-belt (Eocene to Early Oligocene) range between 1.4 and 5.0%Rm; corresponding estimates of peak palaeotemperature are 180-3 15°C. Rocks with higher ranks of organic metamorphism generally display better pressure-solution cleavage, and thermal maturity decreases from south to north. Values of illite crystallinity index are consistent with conditions of advanced diagenesis and anchimetamorphism. Maximum burial depths were prob- ably < 9km and the geothermal gradient at the time of peak heating within this sub-belt was at least 30-35"C/km.

Vitrinite reflectance values for the Nabae Sub-belt (Late Oligocene to Early Miocene) range from 0.9 to 3.7%Rm; our palaeotemperature estimates for these rocks are 140- 285°C. Most values of illite crystallinity fall within the field of advanced diagenesis. The palaeothermal structure within the Nabae section overprints all but the final stage of regional deformation. The highest-rank rocks (anchizone CI values) occur in close proximity to gabbroic bodies of the Maruyama intrusive suite, which penetrated the Shimanto accretionary prism at - 14 Ma. Thus, local rock temperatures (near intrusions or related zones of hydrother- mal discharge) deviated substantially from the regional Miocene gradient, which itself was probably 3 70"Ckrn.

The thermal peak in the Murotohanto Sub-belt post-dated the first two phases of deformation, including out-of-sequence thrusting between the Shiina and Gyoto structural domains. The qualitative correlation between intensities of pressure- solution cleavage and ranks of organic metamorphism indi- cates that the two are temporally related. However, the absolute ages of organic metamorphism and cleavage forma- tion within the Murotohanto Sub-belt remain unknown. Because of this uncertainty, there are two possible scen- arios: (1) a progressive heating event, ranging from Early Eocene to Late Oligocene, followed by uplift and signifi- cant erosion of overburden prior to 15Ma; (2) a single, relatively short-lived thermal overprint at - 15-13 Ma that coincided with anomalous near-trench igneous activity throughout the Outer Zone of southwest Japan. With option 1, the documented rotation in shortening direction within the Murotohanto Sub-belt would be linked to a change in the convergence vector of the Pacific-Kula lithosphere at - 43 Ma. With scenario 2, the shortening direction shifted because of the opening of the Sea of Japan and clockwise block rotation of southwest Japan at 15 Ma.

The heat source responsible for the Eocene-Oligocene thermal structure (whether subsequently overprinted, or not) remains unconstrained. Subduction of relatively young crust from the Kula plate (or fused Kula-Pacific plate) may have generated moderately high geothermal gradients within the Eocene-Oligocene prism. However, with respect to the Shimanto subduction front, the exact position of the Kula- Pacific spreading ridge at 43 Ma remains unknown.

Throughout the Outer Zone, a dramatic Middle Miocene perturbation of geothermal gradients was punctuated by widespread and voluminous near-trench acidic volcanism, emplacement of S-type granites, local mafic intrusions and the formation of anthracite coals. At Cape Muroto, this anomalous event involved collision between the accretion- ary prism and a rigid indentor of significant size. In addition to mafic magmatism, the mechanical responses to the collision included high-angle faulting, differential uplift of the Murotohanto Sub-belt, late-stage mesoscopic faulting and quartz-veining within the Nabae Complex, and forma- tion of a major cross-fold (the Muroto Flexure).

130 M. B. Underwood et al.

A final controversy involves the heat source responsible for the Middle Miocene thermal overprint. There are four possible explanations: (1) subduction of an active spreading segment directly beneath Cape Muroto; (2) subduction of an active spreading ridge beneath the Kii Peninsula, with off-axis magmatism (and seamount collision) at Cape Muroto; (3) development of an unstable and disorganized Shikoku Basin spreading system, with several off-axis mantle con- vection cells passing beneath the subduction front; and (4) a sudden change from Pacific-plate convergence to subduc- tion of a very young oceanic crust of the Philippine Sea plate; this change occurred as southwest Japan experienced clockwise rotation and the margin of Shikoku Basin changed from strike-slip to consuming. We emphasize that these scenarios are not mutually exclusive.

The only faults to show unequivocal evidence of substan- tial post-metamorphic displacement (i.e. after 14 Ma) are the Shiina-Narashi Fault (Nabae-Murotohanto boundary), with > 1 km of post-peak-thermal offset and the Hane River Fault Zone. Apatite fission-track cooling ages of 11- 8Ma indicate that cooling through the 100°C isotherm occurred soon after the cessation of the Middle Miocene magmatic activity. Neotectonic features, including high- angle faults and uplifted marine terranes, show that vertical offset of the Muroto Peninsula has continued into the Quaternary.

ACKNOWLEDGEMENTS Funding for field and laboratory-based elements of this project was provided by a wide variety of sources: NSF Grants EAR-8509461 and EAR-8720743 (to D. E. K. and T. B.); NSF Grant EAR-8706784 (to M. B. U,) ; two GSA Harold Stearns awards (to L. D. and J. P. H.); and a Sigma Xi Grant in Aid of Research (to L. D,). We also acknowledge the Donors of the Petroleum Research Fund, administered by the American Chemical Society, for partial support of this research (Grant No. 19187-AC2 to M. B. U.). Re- views of earlier papers by D. Karig, D. Orange, J . C. Moore, A. Taira, D. Cowan, M. Cloos, J . Sample and D. Fisher helped clarify both the form and the substance of this manuscript. Finally, we extend special thanks to Asahiko Taira for inviting our contribution to this inaugural journal issue.

REFERENCES AGAR S. M., CLIFF R. A., DUDDY I. R. & REX D. C. 1989.

Accretion and uplift in the Shimanto Belt, SW Japan. Journal of the Geological Society, London 146, 893-96.

AIHARA A. 1989. Paleogeothermal influence on organic meta- morphism in the neotectonics of the Japanese Islands. Tectono- physics 159, 291-305.

BARKER C. E. 1988. Geothermics of petroleum systems: Implications of the stabilization of kerogen thermal maturation after a geologically brief heating duration at peak temperature. In Magoon L. B. ed. Petroleum Systems of the United States. U.S. Geological Survey Bulletin 1870, 26-9.

BLENKINSOP T. 0 . 1988. Definition of low-grade metamorphic zones using illite crystallinity. Journul of Metamorphic Geology 6, 623-36.

BREWER J. 1981. Thermal effects of thrust faulting. Earth and Planetary Science Letters 56, 309-28.

RYRNE, T. & DITULLIO L. 1992. Evidence for changing plate motions in southwest Japan and reconstructions of the Philip- pine Sea Plate. The Island Arc, this volume.

CANDE S. C., LESLIE R. B., PARRA J. C. & HOBAR M. 1987. Interaction between the Chile Ridge and Chile Trench: Geo- physical and geothermal evidence. Journal of Geophysical Re- search 92, 495-520.

CHAMOT-ROOKE N., RENARD V. & LEPICHON X. 1987. Magnetic anomalies in the Shikoku Basin: A new interpretation. Earth a d Planetary Science Letters 83, 214-28.

CHIJIWA K. 1988. Post-Shimanto sedimentation and organic metamorphism: An example of the Miocene Kumano Group, Kii Peninsula. Modern Geology 12, 363-87.

DITULLIO L. D. 1989. Evolution of the Eocene accretionary prism in SW Japan: Evidence from structural geology, thermal altera- tion, and plate reconstructions. PhD thesis (unpubl.). 161 pp. Brown University, Providence, RI, USA.

DITULLIO L. D. & BYRNE T. 1990. Deformation paths in the shallow levels of an accretionary prism: The Eocene Shimanto belt of southwest Japan. Geological Society of America BuUetin 102, 1420-38.

~TULLIO L. D. & HADA S. 1992. Regional and local variations in the thermal history of the Shimanto Belt, southwest Japan. In Underwood M. B. ed. Thermal Evolution of the Tertiary Shimanto Belt, Southwest Japan: An Example of Ridge- Trench Interaction. Geological Society of America, Special Paper, Vol. 273 (in press), Boulder, Colorado.

)ITULLIO L. D., LAUGHLAND M. M. & BYRNE T. 1992. Thermal maturity and constraints on deformation from illite crystallin- ity and vitrinite reflectance in the shallow levels of an ac- cretionary prism: Eocene Shimanto Belt, southwest Japan. In Underwood M. B. ed. Thermal Evolution of the Tertiary Shimanto Belt, Southwest Japan: An Example of Ridge- Trench Interaction. Geological Society o f America, Special Paper, Vol. 273 (in press), Boulder, Colorado.

ENGEBRETSON D. C. , COX A. & GORDON R. G. 1985. Relative motions between oceanic and continental plates in the Pacific basin. Geological Society of America, Special Paper 206, 59 pp.

FREY M. 1987. Very low-grade metamorphism of clastic sedimen- tary rocks. In Frey M. ed. Low-Temperature Metamorphism, pp. 9-58. Chapman and Hall, New York.

FURLONG K. P. & EDMAN J. D. 1984. Graphic approach to deter- mination of hydrocarbon maturation in overthrust terrains. American Ass&tion of Petroleum Geologists Bulletin 68, 1818-24.

GUlDOTTl c. v. & SASS1 F. P. 1986. Classification and correla- tion of metamorphic facies series by means of muscovite b, data from low-grade metapelites. News Jahrbuch fiir Minerdogie Abh. 153, 363-80.

GUTHRIE J . M., HOUSEKNECHT D. W. & JOHNS W. D. 1986. Relationships among vitrinite reflectance, illite crystallinity, and organic geochemistry in Carboniferous strata, Ouachita Mountains, Oklahoma and Arkansas. American Association of Petroleum Geologists Bulletin 70, 26-33.

HADA S. & SUZUKI T. 1983. Tectonic environments and crustal section of the Outer Zone of southwest Japan. In Hashimoto M. & Uyed S. eds. Acoetion Tectonics in the Circum-Pacific Regions, pp. 207-18. Terra Scientific Publishing Company, Tokyo.

HAMAMOTO R. & SAKAI H. 1987. Rb-Sr age of granophyre associated with the Cape Muroto gabbroic complex. Science Reports of the Department of Geology, Vol. 15, pp. 1-5. Kyushu University.

HASERE N., TAGAMI T. & NISHIMURA s. 1992. Evolution of the Shimanto accretionary complex; a fission-track thermochrono- logic study. In Underwood M. B. ed. Thermal Evolution of the Tertiary Shimanto Belt, Southwest Japan: An Example of Ridge-Trench Interaction. Geological Society of America, Special Paper, Vol. 273 (in press), Boulder, Colorado.

HAYASHIDA H. 1986. Timing of rotational motion of southwest Japan inferred from paleomagnetism of the Setouchi Miocene Series. Journal ofGeomagnetism and Geoelectricity 38, 295-310.

HIBBARD J. P. 1988. Evolution of anomalous structural fabrics in an accretionary prism; the Oligocene-Miocene portion of the

Thermal evolution of the Tertiary Shimanto Belt 13 1

Shimanto Belt at Cape Muroto, southwest Japan. PhD thesis (unpubl.), 227 pp. Cornell University, Ithaca, NY, USA.

HIBBARD J. P. & KARIG D. E. 1990a. Structural and magmatic responses to spreading ridge subduction; An example from southwest Japan. Tectonics 9, 207-30.

HIBBARD J. P. & KARIG D. E. 1990b. An alternative plate model for the Early Miocene evolution of the SW Japan margin. Geology 18, 170-4.

HIBBARD J. P., KARIG D. E. & TAIRA A. 1992a. Oligocene- Miocene geology of the Shimanto accretionary prism, Muro- tomisaki, Shikoku, Japan. The Island Arc, 1 , 133-47.

HIBBARD J. P., LAUGHLAND M. M., KANG S. M. & KARIG D. E. 1992b. The thermal imprint of spreading ridge subduction on the upper structural levels of an accretionary prism, southwest Japan. In Underwood M. B. ed. Thermal Evolution of the Tertiary Shimanto Belt, Southwest Japan: An Example of Ridge-Trench Interaction. Geological Society of America, Spe- cial Paper, Vol. 273 (in press), Boulder, Colorado.

HISOTAMI K. 1988. The Miocene forearc basin of southwest Japan and the Kumano Group of the Kii Peninsula. Modern Geology 12, 389-408.

HUNZIKER J. C., FREY M., CLAUE N., et al. 1986. The evolution of illite to muscovite: Mineralogical and isotopic data from the Glarus Alps, Switzerland. Contributions to Mineralogy and Pet- rology 92, 157-80.

JOLIVET L., HUCHON P. & RANGIN C. 1989. Tectonic setting of Western Pacific marginal basins. Tectonophysics 160, 23-47.

KANO K., KOSAKA K., MURATA A. & YANAI S. 1990. Intra-arc deformations with vertical rotation axes: The case of the pre- Middle Miocene terranes of southwest Japan. Tectonophysics 176,333-54.

KEMP A. E. S., OLIVER G. H. J. & BALDWIN J. R. 1985. Low-grade metamorphism and accretion tectonics: Southern Uplands terrain, Scotland. Mineralogical Magazine 49, 335-44.

KIMURA K. 1985. Sedimentation and sedimentary facies of the Tertiary Shimizu and Misaki Formations in the southwestern part of Shikoku. Geological Society ofJapanJournal91, 815-31.

KINOSHITA H. & YAMANO M. 1986. The heat flow anomaly in the Nankai Trough area. In Kagami H., Karig D. E. & Coul- bourn W. T. eds. Initial Reports of the Deep Sea Drilling Project, vol. 87, pp. 737-43. US Government Printing Office, Wash- ington, DC.

KISCH H. J. 1983. Mineralogy and petrology of burial diagenesis (burial metamorphism) and incipient metamorphism in clastic rocks. In Larsen G. & Chilingar G. V. eds. Diagenesis in Sediments and Sedimentary Rocks, Part 2. pp. 289-493. Elsevier, Amsterdam.

K~SCH H. J. 1987. Correlation between indicators of very low-grade metamorphism. In Frey M. ed. Low-Tempemure Metamorphism, pp. 227-300. Chapman & Hall, New York.

KISCH H. J. 1990. Calibration of the anchizone: A critical com- parison of illite ‘crystallinity’ scales used for definition. Journal of Metamorphic Geology 8, 31-46.

KUMON F., SUZUKI H., NAKAZAWA K. et al. 1988. Shimanto Belt in the Kii Peninsula, SW Japan. Modern Geology 12, 71-96.

LAUGHLAND M. M. & UNDERWOOD M. B. 1992. Vitrinite reflect- ance and estimates of paleotemperature within the Upper Shimanto Group, Muroto Peninsula, Shikoku, Japan. In Under- wood M. B. ed. Thermal Evolution of the Tertiary Shimanto Belt, Southwest Japan: An Example of Ridge-Trench Interac- tion. Geological Society of America, Special Paper, Vol. 273 (in press), Boulder, Colorado.

LONSDALE P. 1988. Paleogene history of the Kula plate: Offshore evidence and onshore implications. Geological Society uf Ameri- ca Bulktin 100, 733-54.

MACKENZIE J. S., TAGUCHI S. & ITAYA T. 1990. Cleavage dating by K-Ar isotopic analysis in the Paleogene Shimanto Belt of eastern Kyushu, S. W. Japan. Journal of Mineralogy, Petrology and Economic Geology 85, 161-7.

MARSHAK S. & KARIG D. E. 1977. Triple junctions as a cause for

anomalously near-trench igneous activity between the trench and volcanic arc. Geology 5, 233-6.

MARUYAMA S., LIOU J. G. & SENO T. 1989. Mesozoic and Cenozoic evolution of Asia. In Ben-Avraham 2. ed. The Evolution of the Pacific Ocean Margin. pp. 75-99. Oxford Uni- versity Press, New York.

MERRIMAN R. J. & ROBERTS B. 1985. A survey of white mica crystallinity and polytypes in pelitic rocks of Snowdonia and Llyn, North Wales. Mineralogical Magazine 49, 305-19.

MIYAKE Y. 1985. MORB-like tholeiites formed within the Miocene forearc basin, southwest Japan. Lithos 18, 23-34.

MIYAKE Y. 1988. Petrology of the Shionomisaki igneous complex, southwest Japan and its implication for the ophiolite genera- tion. Modern Geology 12, 283-302.

MCORE G. F. & KARIG D. E. 1976. Development of sedimentary basins on the lower trench slope. Geology 4, 693-7.

MCORE J. C. & ALLWARDT A. 1980. Progressive deformation of a Tertiary trench slope, Kodiak Islands, Alaska. Journal of Geo- physical Research 85, 4741-56.

MORI K. & TAGUCHI K. 1988. Examination of the low-grade metamorphism in the Shimanto Belt by vitrinite reflectance. Modem Geology 12, 325-39.

MURATA M. 1984. Petrology of Miocene I-type and S-type gran- itic rocks in the Ohmine district, central Kii Peninsula. Journal Japan Association Mineralogy, Petrology, Economic Geology 79, 351-69.

NAESER N. D., NAESER C. W. & MCCULLOH T. H. 1989. The application of fission-track dating to the depositional and thermal history of rocks in sedimentary basins. In Naeser N. D. & McCulloh T. H. eds. Thermal History of Sedimentary Basins, Methods and Case Histories, pp. 157-80. Springer-Verlag, New York.

NAKADA S. 1983. Zoned magma chamber of the Osuzuyama acid rocks, southwest Japan. Journal of Petrology 24, 471-94.

NAKADA S. & TAKAHASHI M. 1979. Regional variations in chemistry of the Miocene intermediate to felsic magmas in the Outer Zone and the Setouchi Province of southwest Japan. Journal of the Geological Society ofJapan 85, 571-82.

NAKAJIMA T. & HIRCOKA K. 1986. Clockwise rotation of south- west Japan inferred from paleomagnetism of Miocene rocks in Fukui Prefecture. Journal of Geomagnetism and Geoelectricity 38, 513-22.

NIITSUMA N. 1988. Neogene tectonic evolution of southwest Japan. Modern Geology 12, 497-532.

NIITSUMA N. & AKlBA F. 1985. Neogene tectonic evolution and plate subjection in the Japanese island arc. In Nasu N. et al. eds. Formation of Active Ocean margins, pp. 75-108. Terra Scientific Publishing Company, Tokyo.

OBA N. 1977. Emplacement of granitic rocks in the Outer Zone of - southwest Japan and geologic significance. Journal of Geology 85, 383-93.

OSOZAWA S. 1988. Accretionary process of the Tertiary Setogawa and Mikasa Groups, southwest Japan. Journal of Geology 96, 199-208.

OTOFUJI Y. & MATSUDA T. 1984. Timing of rotational motion of southwest Japan inferred from paleomagnetism. Earth and Plan- etary Science Letters 70, 373-82.

OTOFUJI Y. & MATSUDA T. 1987. Amount of rotation of south- west Japan - fan shaped opening of the southwestern part of the Japan Sea. Earth and Planetary Science Letters 85, 289-301.

OTSUKI K. 1990. Westward migration of the Izu-Bonin Trench, northward motion of the Philippine Sea Plate, and their relationships to the Cenozoic tectonics of Japanese island arcs. Tectonophysics 180, 35 1-67.

PADAN A., KISCH H. J. & SHAGAM R. 1982. Use of the lattice parameter b, of dioctahedral illitehnuscovite for characteriza- tion of P R gradients of incipient metamorphism. Contributions to Mineralogy and Petrology 79, 85-95.

RANKEN B., CARDWELL R. K. & KARIG D. E. 1984. Kinematics of the Philippine Sea plate. Tectonics 3, 555-76.

132 M. B. Underwood et al.

ROBINSON D., WARR L.N. & BEVINS R. E. 1990. The illite ‘crystallinity’ technique: A critical appraisal of its precision. Journal of Metamorphic Geology 8, 333-44.

SAKAI H. 1987. Active faults in the Muroto Peninsula of ‘non- active fault province’. Journal of the Geological Society of Japan 93, 513-16.

SAKA~ H. 1988. Origin of the Misaki Olistostrome Belt and re- examination of the Takachiho Orogeny. Journal of the Geologi- cal Society ofJapan 94, 945-61.

SASS F. P. & SCOLARI A. 1974. The b, value of the potassic white micas as a barometric indicator in low-grade metamorphism of pelitic schists. Contributions to Mineralogy and Petrology 45, 143-5 2.

SENO T. & MARUYAMA S. 1984. Paleogeographic reconstruction and origin of the Philippine Sea. Tectonophysics 102, 53-84.

SHIBATA K. & ISHIHARA S. 1979. Initial 87Sr/86Sr ratios of plutonic rocks from Japan. Contributions to Mineralogy and Petrology 70, 381-90.

SHIBATA K. & NOZAWA T. 1967. K-Ar ages of granitic rocks from the Outer Zone of southwest Japan: Geochemical Journal 1 , 131-7.

SHI Y. & WANG C. Y. 1987. Two-dimensional modeling of the P-T-t paths of regional metamorphism in simple overthrust terrains. Geology 15, 1048-51.

TAIRA A. 1985. Sedimentary evolution of Shikoku subduction zone: The Shimanto Belt and Nankai Trough. In Nasu N. et al. eds. F m t i o n of Active Ocean Margins, pp. 835-51. Terra Scientific Publishing Company, Tokyo.

TAIRA A, , TASHIRO M., OKAMURA M. & KAITO J. 1980. The geology of the Shimanto Belt in Kochi Prefecture, Shikoku, Japan. In Taira A. & Tashiro M. eds. Geology and Paleontology of the Shimanto Belt, Kochi, pp. 319-89. Rinya-Kosakai Press, Kochi, Japan.

TAIRA A., OKADA H., WHITAKER J. H. McD. & SMITH A. J . 1982. The Shimanto Belt of Japan: Cretaceous to lower Miocene active margin sedimentation. In Leggett J . K. ed. Trench-Forearc Geology. Geological Society of London, Special Publication 10, 5-26.

TAIRA A., KATE J., TASHIRO M., OKAMURA M. & KODAMA K. 1988. The Shimanto Belt in Shikoku, Japan - Evolution of Cretaceous to Miocene accretionary prism. Modern Geology 12, 5-46.

TAIRA A., TOKUYAMA 13. & SOH W. 1989. Accretion tectonics and evolution of Japan. In Ben-Avraham 2. ed. The Evolution of the Pacific Ocean Margins, pp. 100-23. Oxford University Press, New York.

TERAKADC Y., SHIMIZU H. & MASUDA A. 1988. Nd and Sr isotopic variations in acidic rocks formed under a peculiar tectonic environment in Miocene southwest Japan. Contribu- tions to Mineralogy and Pertrology 99, 1-10.

TORII M. & ISHIKAWA N. 1986. Age estimation of a high-magnesian andesite from the Kii Peninsula, southwest Japan: Possible interpretation of anomalous magnetic direction. Journal of Geomagnetism a d Geoelectricity 38, 523-8.

TORIUMI M. & TERUYA J. 1988. Tectono-metamorphism of the Shimanto Belt. Modern Geology 12, 303-24.

UNDERWOOI) M. B., (ed.) 1992. Thermal Evolution of the Ter- tiary Shimanto Belt, Southwest Japan: An Example of Ridge- Trench Interaction. Geological Society of America, Specud Paper, Vol. 273 (in press), Boulder, Colorado.

U N D E R W ~ D M. B., OLEARY J. D. & STRONG R. H. 1988. Con- trasts in thermal maturity within terranes and across terrane boundaries of the Franciscan Complex, northern California. Journal of Geology 96, 399-416.

UNDERWOOD M. B, LAUGHLAND M., SHELTON K. et al. 1991. Correlations among paleotemperature indicators within orogenic belts: Examples from pelitic rocks of the Franciscan Complex (California), the Shimanto Belt (Japan), and the Kandik Basin (Alaska). American Geophysical Union Transactions (EOS) 72, 549.

U N D E R W ~ D M. B., BROCCULER~ T., BERGFELD D., HOWELL 1). G. & PAWLEWICZ M. 1992a. Statistical comparison between illite crystallinity and vitrinite reflectance, Kandik region of east- central Alaska. In Bradley D. & Dusel-Bacon C. eds. Geologic Studies in Alaska by the U.S. Geological Survey During 1991. US Geological Survey Bulletin (in press).

UNDERWOOD M. B., LAUGIILAND M. M. & KANG S. M. 1992b. A comparison among organic and inorganic indicators of diagenesis and low-temperature metamorphism, Tertiary Shi- manto Belt, Shikoku, Japan. In Underwood M. B. ed. Thermal Evolution of the Tertiary Shimanto Belt, Southwest Japan: An Example of Ridge-Trench Interaction. Geological Society of America, Special Paper, Vol. 273 (in press), Boulder, Colorado.

UNDERWOOD M. B., HIBBARD J. P. & DITULLIO L. 1992c. Geo- logic summary and conceptual framework for the study of thermal maturity within the Eocene-Miocene Shimanto Belt, Shikoku, Japan. In Underwood M. B. ed. Thermal Evolution of the Tertiary Shimanto Belt, Southwest Japan: An Example of Ridge-Trench Interaction. Geological Society of America, Special Paper, Vol. 273 (in press), Boulder, Colorado.

LAUGHLAND M. M. 1992d. The effects of ridge subduction on the thermal structure of accretionary prisms: A Tertiary exam- ple from the Shimanto Belt of Japan. In Underwood M. B. ed. Thermal Evolution of the Tertiary Shimanto Bclt, Southwest Japan: An Example of Ridge-Trench Interaction. Geological Society of America, Special Paper, Vol. 273 (in press), Boulder, Colorado.

UYEDA S. & MIYASHIRO A. 1974. Plate tectonics and the Japa- nese islands: A synthesis. Geological Society of America Bulktin 85, 1159-70.

YANG C. & HESSE R. 1991. Clay minerals as indicators of diagenetic and anchimetamorphic grade in an overthrust belt, External Domain of southern Canadian Appalachians. Clay Minerals 26, 211-31.

WATTS A. B. & WEISSELL J. K. 1975. Tectonic history of the Shikoku marginal basin. Earth a d Planetary Science Letters 25, 239-50.

YAMANO M., HONDA S. & UYEDA S. 1984. Nankai Trough: a hot trench? Marine Geophysical Researches 6, 187-203.

ZAUN P. E. & WAGNER G. A. 1985. Fission-track stability in zircons under geologic conditions. Nuclear Tracks 10, 303-7.

UNDERWOOD M. B., BYRNE T., HIBBARD J. P., DITULLIO L. &