BioOne.org will be down briefly for maintenance on 12 February 2025 between 18:00-21:00 Pacific Time US. We apologize for any inconvenience.
Open Access
How to translate text using browser tools
1 January 2015 An in situ Vesicomyid-dominated Cold-seep Assemblage from the Lowermost Pleistocene Urago Formation, Kazusa Group, Forearc Basin Fill on the Northern Miura Peninsula, Pacific Side of Central Japan
Masayuki Utsunomiya, Ryuichi Majima, Kiminori Taguchi, Hideki Wada
Author Affiliations +
Abstract

We describe the modes of occurrence of a vesicomyid-dominated fossil assemblage in the lowermost Pleistocene Urago Formation, a forearc basin fill on the Miura Peninsula, Pacific side of central Japan. The assemblage consists mostly of Calyptogena (Archivesica) kawamurai (Kuroda) with a minor amount of Conchocele bisecta (Conrad). The shells occur in cross-bedded and massive sandstones of an outcrop that is approximately 25 m wide and 10 m high. The sandstones are interpreted to have been formed by migration of dunes under northward- to eastward-directed bottom currents, judging from the dips of their foreset laminae. Many of the shells are disarticulated and show evidence of reworking by bottom currents, as indicated by their convex-up positions with their commissure planes dipping southwestward (up-current direction) or concordant with the cross laminae. In the massive sandstones, some articulated vesicomyids are preserved in their life position, that is, perpendicular to the bedding plane with their anterior parts oriented downward. Massive authigenic carbonates, which are developed in nearly all horizons of the vesicomyid-bearing sandstones, consist exclusively of Ca-rich dolomites (δ13C, -37.78‰ to -24.16‰ VPDB; δ180, 0.69‰ to 4.35‰ VPDB). Biogenic carbonates have entirely dissolved, and some of the resulting molds are filled with dolomite-cemented fine clastics.

We consider this fossil assemblage to be an in situ cold-seep dependent assemblage because of its association with 13C-depleted authigenic carbonates, which clearly reflect the influence of the anaerobic oxidation of methane. We infer that the vesicomyids lived on sand dunes during time intervals between intermittent sanddeposition events under the unidirectional bottom currents.

Introduction

Many modern vesicomyids are deep-sea chemosymbiotic. semi-infaunal bivalves living at cold-seep sites (Sibuet and Olu, 1998; Kojima, 2002; Krylova and Sahling, 2010), and vesicomyids are also a major component of Cenozoic seep-related fossil assemblages (Majima, 1999; Majima et al., 2005; Campbell, 2006). The occurrence of vesicomyid fossils is thus both a spatial and temporal indicator of the ancient methane seepages. Therefore, because atmospheric methane is an exceedingly powerful greenhouse gas (Hanson and Hanson, 1996), vesicomyid fossils are very useful for evaluating the influence of ancient carbon cycles on global environmental history.

On the Miura and Boso peninsulas, Pacific side of central Japan, many fossil vesicomyid-dominated assemblages have been reported from Miocene-Pleistocene marine strata (Majima et al., 2005). The abundant occurrence of vesicomyid assemblages reflects the drastic tectonic histories controlled by the presence of a trench-trench-trench-type triple junction off the Boso Peninsula, where the Philippine Sea plate is subducting underneath the North American plate, and the Pacific plate is subducting underneath both the North American and Philippine Sea plates (Figure 1: McKenzie and Morgan, 1969). Due to the presence of this triple junction, the Izu-Bonin arc on the Philippine Sea plate is colliding with the Honshu arc on the North American plate; this collision has been forming a bent structure, called the Kanto syntaxis, on the Honshu arc since Miocene (Takahashi and Saito. 1997). and inducing both tectonic compression and stretching in the Miura and Boso regions, located on the eastern wing of the syntaxis. These tectonic activities in the Miura and Boso regions since Miocene have resulted in active methane seepages that have maintained the cold-seep assemblages.

Figure 1.

Geological map of the northern part of the Miura Peninsula, central Japan (after Eto, 1986a; Eto et al., 1998). Abundant vesicomyid-dominated fossil assemblages have been reported from fossil localities 1–12 (Niitsuma et al., 1989), 13 (Hirata et al., 1991), and 14 (Utsunomiya et al., 2014). The fossil localities were grouped according to their modes of occurrence into Type A (in association with submarine landslide deposits), Type B (reworked shells in a sandstone bed deposited by a sediment gravity flow), Type C (in situ occurrence in submarine sand-ridge deposits), or unknown.

f01_01.jpg

We focus here on the forearc basin fills exposed on the northern Miura Peninsula, that is, the Zushi and Ikego fonnations of the Miura Group and the overlying Kazusa Group (Figure 1). In these forearc basin fills, vesicomyiddominated fossil assemblages are abundant (Niitsuma et al., 1989; Majima et al., 2005). However, there is no evidence hitherto for determining the presence of an in situ fossil assemblage. Here we newly report an in situ vesicomyid-dominated fossil assemblage from the Urago Formation. Kazusa Group, that occurs in sand-ridge deposits cemented with 13C-depleted authigenic carbonates.

Remarks on the fossil localities of vesicomyids

Geological setting

The stratigraphy of the forearc basin fills on the northern Miura Peninsula is summarized in Figures 1 and 2. mainly following Eto (1986a, b) and Eto et al. (1998). and also the other recent stratigraphic studies referred below. They are composed of the Miura and Kazusa groups. The Miura Group consists of the upper Miocene to lower Pliocene Zushi Formation and the Pliocene Ikego Formation, in ascending order. The Zushi Formation is composed mainly of alternating beds of sandstone and mudstone that were deposited at water depths between 500 and 2000 m, as estimated from the benthic foraminiferal assemblages (Eto et al., 1987). The Ikego Formation consists of a main part and two basal members, the Takatoriyama and Jimmuji members. These two members have a restricted distribution, and they are time-equivalent to the main part (Figure 1). The main part of the formation consists of alternating beds of tuffaceous sandstone and mudstone that were deposited at water depths between 500 and 2000 m, as estimated from the benthic foraminiferal assemblages (Eto et al., 1987). The Takatoriyama Member consists of massive and cross-bedded tuffaceous sandstones and conglomerates that have been interpreted as canyon-fill deposits because of the convex-down shape of the deposits inferred from their geographic distribution (Soh et al., 1991). The Jimmuji Member consists of chaotic deposits containing slide-generated sandstone and mudstone blocks (Eto et al., 1998). Both members bear abundant vesicomyiddominated fossil assemblages (Niitsuma et al., 1989; Yokohama Defense Facilities Administration Bureau. 1993).

Figure 2.

Ages, stratigraphy, lithologies, and paleobathymetries of the Miura and Kazusa groups, forearc basin fills exposed on the northern Miura Peninsula. The stratigraphy is mainly following Eto (1986a) and Eto et al. (1998), and modified in ages following Tamura and Yamazaki (2010), and in lithology following Tate and Majima (1998). The paleobathymetries are from *1, Oyama (1951); *2, Mitsunashi and Kikuchi (1982); *3, Eto et al. (1987); *4. Tate and Majima (1998); and *5. Utsunomiya and Majima (2012).

f02_01.jpg

The Kazusa Group consists of the Urago, Nojima, Ofuna, Koshiba, Nakazato, and Tomioka formations in ascending order. The Kazusa Group seems to overlie conformably the Miura Group on the Miura and western Boso peninsulas, whereas on the eastern Boso Peninsula, the basal conglomerates of the Kazusa Group unconformably overlie the Miura Group, the upper surface of which is distinctly erosional (Koike. 1951). The Urago Formation is composed of tuffaceous sandstones and muddy sandstones that were deposited at water depths between 400 and 600 m (Utsunomiya and Majima, 2012). It has been dated to the latest Pliocene to earliest Pleistocene on the basis of a widespread tephra (KGP). dated to 2.5 Ma by tephrochronology (Tamura and Yamazaki, 2010). that is intercalated in the middle horizon of the formation (Inagaki et al., 2007; this study). The formation's sandstones contain a vesicomyid-dominated assemblage (Shikama and Masujima, 1969; Niitsuma et al., 1989; Hirata et al., 1991; Majima et al., 2005). The Nojima Formation is composed mainly of sandy mudstones, muddy sandstones, and sandstones (Mitsunashi and Kikuchi, 1982; Eto et al., 1998). The Imaizumi Member, consisting of well-bedded sandstones and conglomerates that are intercalated in the lower part of the fonnation, bears vesicomyids (Shikama and Masujima, 1969; Niitsuma et al., 1989; Hirata et al., 1991; Majima et al., 2005; Utsunomiya et al., 2014). Utsunomiya and Majima (2012) estimated that the lower part of the formation was deposited at water depths between 400 and 500 m from the molluscan assemblage, whereas Eto et al. (1987) indicated that it was deposited at water depths between 500 and 1000 m, on the basis of the benthic foraminiferal assemblage. The Ofuna Formation consists of mudstones, and the overlying Koshiba Formation consists of sandy mudstones, muddy sandstones, and sandstones in ascending order (Tate and Majima, 1998; Kitazaki and Majima, 2003). The upper Ofuna and the lower Koshiba formations bear lucinid- or tyasirid-dominated cold-seep assemblages (Majima et al., 1996; Tate and Majima, 1998; Kitazaki and Majima, 2003). Their paleobathymetries, estimated from the molluscan assemblages, are water depths between 200 and 300 m, and between 100 and 200 m, respectively (Majima et al., 1996; Tate and Majima, 1998; Kitazaki and Majima, 2003). The differences in the dominant chemosymbiotic bivalves among these forearc basin fills probably reflect the water depth: the Ikego, Urago, and lower Nojima formations, which bear vesicomyid-dominated assemblages, were deposited at water depths exceeding 400 m, whereas the upper Ofuna and Koshiba formations, which bear lucinid- or thyasirid-dominated assemblages, were deposited at water depths shallower than 300 m.

The modes of fossil occurrence

Vesicomyid-dominated assemblages have been observed at fossil localities 1 to 12 (Niitsuma et al., 1989). 13 (Hirata et al., 1991). and 14 (Utsunomiya et al., 2014) (Figure 1). At Locs. 9 to 11 (Takatoriyama Member), Loc. 8 (Jimmuji Member). Loc. 6 (Ikego Formation main part). Locs. 1 and 2 (Imaizumi Member), and Locs. 12 and 13 (Nojima Formation main part) (Figure 1). the modes of occurrence of vesicomyids cannot be determined because of scarce data and because the outcrops are no longer exposed. In the other localities, the vesicomyids occur exclusively in sandstone beds, some of which show distinct sedimentary structures such as parallel or cross-lamination. The fossil localities of the vesicomyid-dominated assemblages can be classified into three types according to their occurrence modes: Type A. those associated with submarine landslide deposits; Type B. reworked shells in a sandstone bed deposited by a sediment gravity flow; and Type C. in situ shells in submarine sand-ridge deposits. Among them, Type C shows a peculiar mode of fossil occurrence as mentioned below. In order to compare the mode of occurrence among these types and emphasize the peculiarity of Type C, the modes of types A and B as well as Type C are briefly reviewed in this section.

Type A.—Niitsuma et al. (1989). Eto (1993). and Yokohama Defense Facilities Administration Bureau (1993) observed vesicomyid-dominated assemblages contained in slide-generated sandstone blocks of the Jimmuji Member exposed at Locs. 4, 5, and 7. For example, at Loc. 4, many vesicomyid fossils were observed in large outcrops exposed by residential construction (Figures 1, 3.1–3.3). The vesicomyid fossils occur mostly in centimeter- to decimeter-sized sandstone blocks, some of which are upside down relative to the general bedding orientation in the area. Cross lamination is frequently developed in the vesicomyid-bearing sandstone blocks. Vesicomyids also occur in the sandstone and mudstone matrix among the blocks. We consider the probable derivation of the shells within the matrix to be the sandstone blocks. One of the sandstone blocks contains dense aggregates of vesicomyid fossils (20 individuals per 500 cm2) (Figures 3.2, 3.3) with more than 90% articulated. The parts of the blocks with shell aggregates are cemented with 13C-depleted (δ13C = -30‰ to -40‰ PDB) authigenic calcites.

Yokohama Defense Facilities Administration Bureau (1993) interpreted the Takatoriyama Member as submarine canyon-fill deposits, and the adjacent Jimmuji Member as submarine landslide deposits derived from both the canyon fill (Takatoriyama Member) and the canyon wall (Zushi Formation and the main part of the Ikego Formation). They suggested that methane-bearing fluid seepage, evidenced by the occurrences of vesicomyid fossils and 13C-depleted authigenic carbonates, caused the submarine canyon fills and the walls to be unstable and led to a huge submarine landslide, resulting in the vesicomyid-bearing blocks of the Jimmuji Member.

Type B.—Utsunomiya et al. (2014) observed the occurrence of a vesicomyid-bearing sandstone deposited by a sediment gravity flow in the lower part of the Nojima Formation exposed at Loc. 14, Kuden, Sakae-ku, Yokohama City (Figures 3.4–3.7). At this location, pebbly sandstones, fine- to coarse-grained sandstones, and muddy sandstones are exposed in an outcrop, 3 m high and 10 m wide. A 50- to 53-cm-thick pebbly sandstone bed at this locality bears vesicomyids (Figures 3.5, 3.6) along with shells of shallow-water mollusks such as Scapharca satowi (Dunker) and Patinopecten yessoensis (Jay). These molluscan fossils have commonly suffered abrasion and fragmentation. The pebbly sandstone bed consists of three parts: a reverse-graded part, a normally graded part, and a parallel-laminated part, in ascending order. Utsunomiya et al. (2014) interpreted the bed to have been deposited by a sediment gravity flow, inferring that the reverse-graded part resulted from the debris flow, and that the normally graded and parallel-laminated parts (Bouma divisions Ta and Tb; Bouma. 1962) were deposited by a turbidity current. The sediment gravity flow presumably originated in a near-shore environment demonstrated by occurrences of shallow-water mollusks and entrained vesicomyid shells at depths below 400 m along its flow path, resulting in the mixed occurrence of vesicomyids and shallow-water mollusks.

Type C.—At Loc. 3, Ten-en. Kamakura City (Loc. 203 of Shikama and Masujima, 1969; Loc. 3 of Niitsuma et al., 1989; Loc. 60 of Majima et al., 2005). the vesicomyid fossils occur in tuffaceous medium- to coarse-grained sandstones in the middle part (lowermost Pleistocene) of the Urago Formation. Aggregates of both articulated and disarticulated vesicomyid fossils occur in association with well developed authigenic carbonate concretions (Niitsuma et al., 1989; Majima et al., 2005). We interpret these vesicomyid fossils to have lived on these sands; thus, these are the only examples of in situ vesicomyid fossils in the forearc basin fills exposed on the Miura Peninsula. In this paper, we describe in detail the modes of occurrence of the vesicomyid fossils at Loc. 3 to confirm that they lived on these sand-ridge deposits.

Figure 3.

Modes of fossil occurrence of the vesicomyid-dominated assemblages. 1–3, Type A at Loc. 4, after Eto (1993) and Yokohama Defense Facilities Administration Bureau (1993); 4–7, Type B at Loc. 14 (Utsunomiya et al., 2014). 1, Sketches of two outcrops exposed by residential construction. The vesicomyid fossils occur in decimeter-sized slide-generated blocks and in matrix (Jimmuji Member of the Ikego Formation). 2, Enlarged sketch of the black rectangle in 1. 3, Photograph of a portion of the area sketched in 2. A lens cap shown in the photo for scale. 4–6, Photographs of the outcrop (4) and of planes cut normal (5) and parallel (6) to the bedding plane at Loc. 14, where the Imaizumi Member of the Nojima Formation is exposed. 7, Sketch of the geologic column at Loc. 14. Most of the vesicomyids in the pebbly sandstone bed are fragmented, convex-down, and matrix-supported (5–7).

f03_01.jpg

Methods

Core boring

The two cores No. 1 (25.5 m long) and No. 2 (40 m) were bored normal to the bedding plane (strike N60°W. dip 15°NE) at Loc. 3 (Figure 5). by using an oil-feed-type rotary core drilling machine with an 86-mm-diameter coring bit. Both of the obtained cores were 7 cm in diameter. and their recovery rates are 100%.

Description of mode of fossil occurrence

At Loc. 3, we counted the shells in each of four preservation states (articulated shells, convex-up or convex-down disarticulated shells, and state unknown), then each disarticulated shell was counted as 0.5 for quantitative comparison (Figure 6). We measured the orientations of their commissure planes and shell's long axes which we show in lower hemisphere stereographic projections (Figure 6).

Mineralogical analysis

The authigenic carbonate minerals precipitated in the cores were identified by X-ray diffraction (XRD) (Rigaku LINT 2000 diffractometer at Yokohama National University) of powdered samples collected from the matrix by using a microdrill under a stereomicroscope. Ca amounts of the authigenic carbonates were estimated using the linear function between the dolomite d104 peak positions and the Ca amounts (Lumsden, 1979).

Stable isotope analysis

The same powdered samples used for the XRD analyses were used for stable carbon and oxygen isotope measurements. which were carried out using a Finnigan MAT 250 mass spectrometer at Shizuoka University. The CO2 was extracted from the dolomite samples for isotope measurements by reacting each sample with concentrated H3PO4 for 60 min at 100°C. The isotope ratios were calibrated relative to Vienna PeeDee Belemnite (VPDB) by using the NBS 19 and 20 standards from the U. S. National Bureau of Standards. The phosphoric acid fractionation factor (1000 1nα co2-dolomite) at 100°C (9.01; Nagai and Wada, 1993) was used for calibration of the oxygen isotope ratios. The standard deviation of the NBS 20 value relative to the reference standard gas used to determine reproducibility was better than 0.06‰ for carbon and 0.15‰ for oxygen.

Table 1.

δ13C and δ18O values of the authigenic carbonates of cores No. 1 and No. 2.

t01_01.gif

Sedimentary facies and sedimentary environments

Lithology

In the observed area around Loc. 3 (Figure 4). the Urago Formation is composed mainly of tuffaceous medium- to coarse-grained sandstones and tuffaceous muddy sandstones, among which fine (<0.063 mm in diameter) and coarse (0.063–2 mm in diameter) ash tuff beds and pumice- or scoria-rich lapilli tuff beds are frequently intercalated. We divided the Urago Formation exposed in the study area into the sandstone-dominated and muddy sandstone-dominated horizons (Figure 4).

Figure 4.

Geologic map (1) and columns (2) of the study area. Geologic columns were made along traverses 1–14 shown by red arrows on the map in (1). and of traverses 15–18 and cores No. 1 and No. 2 (locations shown in Figures 5.1 and 5.2).

f04_01.jpg

Figure 5.

Topographic map (1), cross section (2) and outcrop photographs (3–5) at Loc. 3. 1, Topographic map showing the extent of the outcrop at Loc. 3 (stippled area). The black arrows show the locations of the geologic columns [15]–[18] diagrammed in Figure 3.2. 2, Cross section along line A-A' in (1). 3, 4, Mode of occurrence of vesicomyid shells in the cross-bedded sandstones. White arrows indicate erosional surfaces, and black arrows indicate cross-laminae of the cross-bedded sandstones. Convex-up disarticulated shells (white dotted circles) occur sparsely and are oriented parallel to erosional surfaces (3) or the cross laminae (4). 5, Locations of outcrops shown in Figures 6.1 and 6.2 (95-cm-long shovel as scale).

f05_01.jpg

Figure 6.

Modes of occurrences of fossil vesicomyid shells at Loc. 3 in the upper massive sandstone horizons (1 and 2) and in the lower cross-bedded horizons (3). Pie charts show the proportion of shells classified as articulated, convex-up or convex-down disarticulated, or state unknown. A disarticulated shell was counted as 0.5 shell for quantitative comparison. Shell orientations are plotted in equal-angle projections on the lower hemisphere. In the stereonet, the shell commissure planes are shown as great circles, and shell's long axes are shown as three symbols, closed circles (articulated shell), open circles (disarticulated, convex up) and cross marks (disarticulated, convex down). For some articulated vesicomyids, the anterior directions were confirmable from their marginal shapes, and shown as black arrows extending from the projected points (filled circles) in the stereographic projections.

f06_01.jpg

At Loc. 3, vesicomyid fossils occur in cross-bedded and massive sandstones within restricted horizons of traverses 15 to 18 and the cores No. 1 and No. 2 (Figure 4.2; vertical red bars adjacent to the columns). The sand-stones are mainly composed of volcanogenic materials such as pumices, scorias, euhedral or subhedral plagio-clases. and volcanic-rock fragments, and they also contain minor amounts of other biogenic materials (mainly foraminifers), in addition to bivalves. The sands composing these sandstones prograded northward to eastward, because the trough and tabular cross-bedding (single set thickness. 5 to 60 cm) indicates northward to eastward directed paleocurrents (Figure 4.2; rose diagrams of traverses 15 and 17). These paleocurrent directions agree well with those reported by Naganuma et al. (1973). who measured foreset dip directions of the cross-bedded sandstones of the Urago Formation at other sites.

Figure 7.

Modes of occurrence of vesicomyid shells in the outcrop at Loc. 3. 1, Shells occur as matrix-supported molds (black arrows), some of which are partly filled with carbonate-cemented fine clastics (white arrows). 2, Even though some shells appear to retain their original shell material (white arrow), they actually are composed entirely of fine clastics cemented with authigenic carbonates, as seen in thin section (Figures 11.2, 11.3).

f07_01.jpg

Interpretation of sedimentary environments

We interpreted these sandstone deposits to have formed from dunes migrating under unidirectional bottom currents. We did not interpret them as channel-fill deposits because they lack characteristic features of channel-fill deposits, namely, distinct erosional surfaces between the bedded sandstones and the underlying muddy sandstones, and distinct upward-thinning and -fining successions within the sandstone beds (Walker, 1992). Nakayama and Masuda (1989) reported similar ancient sand-ridge or sand-wave deposits in the Kazusa Group on the Boso Peninsula. Wynn and Stow (2002) also showed that modern sand-ridge deposits are controlled by bottom currents at water depths of the upper continental slopes, similar to the depositional environment of the Urago Formation (between 400 and 600 m water depth; Utsunomiya and Majima. 2012).

Modes of occurrence of the vesicomyid-dominated assemblage

Mode of fossil occurrences

The outcrop of vesicomyid-bearing sandstones trends N-S and is approximately 25 m wide and 10 m high (Figure 5). The sandstones (Figure 4.2; traverses 15 to 18) are composed of lower cross-bedded (Figures 5.3, 5.4) and upper massive (Figures 5.5, 6.1, 6.2) horizons. At least some of the upper massive sandstone horizons appear to have primary cross-bedded structures only partly preserved as weak cross-bed relicts (Figures 6.1, 6.2). These cross-bed relicts suggest that the deposits suffered bioturbation, which erased some of the primary sedimentary structure.

Bivalves in the sandstones, which are generally matrix-supported (Figures 6.1, 6.2, 7), consist mostly of vesicomvids (Figures 8.1–8.4) with a minor amount of thyasirids (Figure 8.5). The vesicomyid fossils obtained from the outcrop have been identified as Calyptogena (Archivesica) kawamurai (Kuroda). which is characterized by an elongate ovate shell, a rounded posterior margin, a distinct hollow called the subumbonal pit (SP; Figure 8.4). and a steeply sloping posterior cardinal tooth (3b; Figure 8.4) (see Amano and Kiel, 2010). The thyasirid specimen (Figure 8.5). identified as Conchocele bisecta (Conrad), is characterized by a subquadrate shell, a nearly vertical anterior end, and a deep radial sulcus that extends from the beak to the posterior margin of the shell (see Coan et al., 2000). In the following descriptions. we use the terms “shell” or “chemosymbiotic bivalve” for the bivalve fossils from Loc. 3, because it was not practical to extract and identify all of the bivalve fossils in the outcrop.

Figure 8.

Vesicomyid (1–4) and thyasirid (5) fossils collected from Loc. 3. 1–4, Calyptogena (Archivesica) kawamurai (Kuroda); 5, Conchocele bisecta (Conrad). All photographs are of rubber casts. 1, KPM-NN0001466; 2, KPM-NN0001469; 3A (dorsal view), 3B, KPM-NN0001470; 4, KPM-NN0001471; 5, KPM-NN0001468. SP, subumbonal pit; 3a, anterior cardinal tooth; 1, central tooth; 3b, posterior cardinal tooth. The hinge structure terminology follows that of Amano and Kiel (2010). The described specimens are housed in Kanagawa Prefectural Museum of Natural History (KPM-NN).

f08_01.jpg

Chemosymbiotic bivalves occur sparsely in some of the lower cross-bedded sandstone horizons and abundantly in the upper massive sandstone horizons. Nearly all of the shells exhibit no evidence of shell fragmentation or abrasion. In Figure 6, pie charts show the proportion of shells classified as articulated, convex-up or convex-down disarticulated, or state unknown. Stereographic projections show the shell commissure planes as great circles, and shell's long axes as closed circles (articulated shell), open circles (disarticulated, convex up), and cross marks (disarticulated, convex down). For some articulated vesicomvids. the anterior directions were confirmable from their marginal shapes, and are shown as black arrows extending from the projected points (filled circles) in the stereographic projections (Figure 6).

Implications of the occurrence of the shells

In the lower cross-bedded sandstone horizons, convex-up shells dominate, and no articulated shells were observed (Figure 6.3). Flume experiments have shown that disarticulated shells are more likely to be buried in the convex-up position under a unidirectional current (Brenchley and Newall, 1970). The shells lie on the erosional base of each set (Figures 5.3, 5.4; white arrows) or on the foreset laminae (Figures 5.3, 5.4; black arrows). In both cases, the long axes are oriented northwest-south-east (Figure 6.3). indicating that they came to rest with their long axes nearly perpendicular to the current direction (northward to eastward; Figure 4.2). In flume experiments. convex-up shells ultimately become oriented with their long axes perpendicular to the current direction (Allen, 1992; Brenchley and Newall, 1970).

The commissure planes of two convex-up shells on the erosional surface dip southwestward, obliquely to the surface (Figure 6.3). Shells lying on the erosional surfaces would have been dragged and come to rest in the position of maximum friction on the sandy surface. The up-current dip of these convex-up shells represents a more stable position for shells being dragged by a unidirectional current. In contrast, the commissure planes of the two shells on the foreset laminae dip northeastward (Figure 6.3). concordant with the dip of the laminae. The opposite dip of the shells in the lower cross-bedded sandstone horizons can thus be explained by their depositional positions under a northeastward unidirectional current.

In the upper massive sandstone horizons (Figures 6.1, 6.2), most shells are disarticulated and convex-up, but there are moderate amounts of disarticulated convex-down and articulated shells. The valves of all of the observed articulated shells are entirely closed. Interpretation of the shell orientations in the upper massive sandstone horizons is difficult because of the erasure of primary sedimentary structures, probably by bioturbation. Although the commissure planes of the disarticulated shells dip in various directions, most dip southeastward to southwestward (Figures 6.1, 6.2), comparable to the dips of the shells on erosional surfaces in the lower cross-bedded sandstone horizons (Figure 6.3). The orientations of the shell's long axes are highly variable and seem to show no relation to the current direction (northward to eastward; Figure 4.2). The resting positions of the disarticulated shells in the upper massive sandstone horizons probably reflect both physical reworking under northward to eastward currents on erosional surfaces and the random effects of bioturbation. The different sedimentary processes between the lower cross-bedded and the upper massive sandstone horizons may be due to differences in the flow regime of the currents, because we observed no shells oriented concordantly with foreset laminae in the latter.

In the upper massive sandstone horizons, articulated shells occur commonly (Figures 6.1, 6.2), and we observed three articulated specimens (Figure 6.1; white arrows on the sketches and stereonet) orienting their long axes nearly perpendicular to the general bedding plane of the exposure and their anterior ends down. This orientation is very similar to the life position of vesicomyids living in soft sediments, in which juveniles are usually entirely, and adults partly, buried within the sediments (Fujikura et al., 2002). We thus infer that the orientations of these three articulated shells reflect their life positions, rather than the accidental outcome of physical reworking or random bioturbation.

Authigenic and biogenic carbonates

Lithology and mineralogy

Authigenic carbonates are developed in many horizons of the Loc. 3 outcrop and cores No. 1 and No. 2 (Figures 4.2, 9.1), In the cores, they are well developed in the horizons of the upper half, which correspond to the cross-bedded and massive sandstones of the outcrop, and they are sporadic in the horizons of the lower half, which are composed of massive sandstones and muddy sandstones that are not exposed in the outcrop (Figures 4.2, 9.1).

The authigenic carbonates were composed exclusively of dolomites with 54 to 58 mol% CaCO3. Authigenic Carich dolomite is often reported from modern cold-seep sites (Naehr et al., 2007; Takeuchi et al., 2007). along with aragonite and high-Mg calcite. The dolomites studied are composed of micrite (diameter < 5 µm; Figure 10.1) and microspar (5 µm < diameter < 50 µm; Figure 10.2).

So far as we could observe, biogenic carbonates (molluscan shells and foraminiferal tests) have dissolved entirely. Some of the dissolved shell molds have remained empty (Figures 7, 9.2; black arrows) and others are filled with dolomite-cemented fine clastics (white arrows). At 5.67 to 5.72 m in core No. 2. we observed one mold of an articulated bivalve shell filled with dolomite-cemented fine clastics (Figure 11) in which the internal mold is directly adjacent to the external mold on its lower (gravity direction) side (Figure 11.1; dashed line); showing the internal mold fell under the force of gravity after the dissolution of the shell. The remaining empty space and adjacent fractures in the matrix are filled with dolomite-cemented fine clastics (Figures 11.2, 11.3). Foraminiferal tests have also dissolved entirely (Figures 10.3, 10.4). The walls of the foraminiferal molds, which are composed of rhombic dolomite crystals, are sharply defined (Figures 10.3, 10.4).

Figure 9.

Geologic columns (1), stable isotope ratios of authigenic carbonates of cores No. 1 and No. 2 (1), and photographs of selected horizons of the core No. 2 (2). Matrix-supported shell molds are unfilled (black arrows) or filled with carbonate-cemented fine clastics (white arrows).

f09_01.jpg

Figure 10.

Scanning electron microscope images of authigenic dolomites precipitated in the vesicomyid-bearing horizons. 1, micritic cement, consisting of irregular-shaped dolomite grains less than 5 µm in diameter (from 5.69 m in core No. 2): 2, microspary cement consisting of sharp-edged rhombohedral dolomite crystals approximately 20 µm in diameter (from 5.69 m in core No. 2): 3, 4, a foraminiferal mold encrusted with platy rhombohedral dolomite crystals (from 3.80 m in core No. 1).

f10_01.jpg

Implications for the origin of dolomites and diagenesis

Given the sharply defined walls of the foraminiferal molds and the gravity fall of the internal bivalve mold, dolomite cementation of the sandy matrix must have occurred before dissolution of the aragonitic (vesicomyid and thyasirid bivalves) and calcitic (foraminifers) bioclasts. In the case of the bivalve, the fallen internal mold strongly suggests that the mold had already become consolidated (cemented by carbonates) before it fell. In the case of the foraminifers, dolomite growth was apparently constrained by the presence of the original foraminifer test. In addition, because fine clastics filling the shell molds are cemented with dolomite, dolomite precipitation must have occurred during at least two different stages, before and after the dissolution of the biogenic carbonates.

The complete dissolution of biogenic aragonites and calcites suggests that even if the authigenic aragonites and calcites were initially precipitated, they were entirely lost during the diagenetic history. Thus, the authigenic carbonate composition at this site may differ from the primary composition, which at present-day seep sites generally is some combination of aragonite, calcite, and dolomite (Naehr et al., 2007).

Isotope signatures of the authigenic carbonates

Results of isotopic analysis

All measured dolomites were depleted in 13C (δ13C = -37.78‰ to -24.16‰ VPDB; Figure 9.1 and Table 1). 13C-depleted carbonates were recognized not only in the shell-bearing horizons, where authigenic carbonate concretions were abundant, but also in those horizons without shells, below 12 m in core No. 1 and below 11 m in core No. 2 (Figure 9.1). where authigenic carbonate occurred sporadically. The oxygen isotope ratios (δ18O) ranged from 0.69‰ to 4.35‰ VPDB (Figure 9.1 and Table 1). and tended to be lower in the upper part of the cores.

Figure 11.

Longitudinal section of core No. 2 (5.62 to 5.79 m). 1, A shell mold and fractures are filled with dolomite-cemented fine clastics. The inner mold has dropped down to where it is directly adjacent to the external mold of the shell (dashed line). 2, Enlarged view of the boundary (white arrows) between the mold-filling clastics above and the inner mold of the shell below. 3, Enlarged view of the mold-filling clastics.

f11_01.jpg

Interpretation of carbon isotopic compositions

The carbon isotopic compositions of authigenic carbonates precipitated beneath the seafloor at seep sites reflect those of the dissolved inorganic carbon in the interstitial water. This carbon generally consists of a mixture of carbon from methane (δ13C = -110‰ to -50‰ for microbially mediated methane, and δ13C = -50‰ to -20‰ for thermogenic methane; Whiticar, 1999). from the decomposition of organic material (-25‰; Hoefs, 2004), and from bottom seawater (about 0‰). These values suggest that the authigenic carbonates with carbon isotope ratios of less than -25‰ were precipitated under the influence of methane oxidation (AOM: anaerobic oxidation of methane). 13C-depleted authigenic carbonates are known to be precipitated just below the seafloor at many modern seep sites (Canet et al., 2006; Campbell et al., 2010).

Interpretation of oxygen isotopic compositions

The oxygen isotope ratios (δ18O) (0.69‰ to 4.35‰ VPDB; Figure 9.1 and Table 1) are lighter than, or partly overlap, the expected range of values, as discussed below. We calculated the expected oxygen isotope ratios of dolomite under the assumption that (1) they were precipitated in the same oceanic setting as is found at present in the sea adjacent to the fossil locality; (2) they precipitated just below the seafloor, as suggested by the carbon isotope data; (3) the isotope ratio of the ambient pore water was 0‰ VSMOW (Standard Mean Ocean Water); and (4) the temperature of the ambient pore water was the same as that of the seawater just above the seep site.

Relational expressions between the water-dolomite isotope fractionation factor and water temperature have been estimated by inorganic experiments conducted at high temperatures (200 to 820°C; Northrop and Clayton, 1966; O'Neil and Epstein, 1966; Matthews and Katz, 1977) as well as low-temperature experiments conducted at 25 to 45°C (Vasconcelos et al., 2005). We assumed that the measured dolomites from Loc. 3 were precipitated in near-surface environments; therefore, for our calculations we used a relational equation obtained from low-temperature experiments (Vasconcelos et al., 2005). In the equation, the water-dolomite isotope fractionation factor is given in VSMOW scale, so we used the equation of Copien et al. (1983) to convert the fractionation factor from VSMOW to VPDB scale. For water temperatures at 400 and 600 m water depth (the paleobathymetry of the Urago Formation; Utsunomiya and Majima. 2012), we adopted values of 4.28 to 9.75°C, the temperature range at those depths in the sea adjacent to the Miura Peninsula (between 35° N and 36° N and between 139° E to 140° E) (Japan Oceanographic Data Center, 2013). Using these assumptions and estimates, we calculated the expected oxygen isotope ratios of the dolomites to be 3.36‰ to 4.68‰ PDB. These values are heavier than, or partly overlap the upper end of, the observed range in the dolomites (0.69‰ to 4.35‰ VPDB; Figure 9.1 and Table 1).

To explain the lighter isotope values of the observed dolomites, we consider two hypotheses: (1) warmer seep fluids, and (2) 18O-depleted seep fluids.

Hypothesis of warmer seep fluids.—The expected water temperatures were calculated to be 5.6 to 21.9°C from the oxygen isotope values of the observed dolomites, by using the relational equation of Vasconcelos et al. (2005). Present-day vesicomyid colonies and microbial mats occur at about 1200 m water depth off Hatsushima Island in western Sagami Bay. At this site, high geothermal gradients (up to 40°C/m) (Iwase et al., 2003; Kinoshita et al., 1991) are caused by a shallow magmatic intrusion near the volcanic front of the Izu-Bonin arc (Kinoshita et al., 1991). The carbonate precipitation temperatures (3.6 to 6.8°C). calculated from the oxygen isotopes of authigenic calcites from the cold seepage area off Hatsushima Island (δ18O = 3.24‰ to 4.09‰ PDB; Hattori et al., 1994), are slightly higher than temperatures of the surrounding seawater (2.5 to 3.3°C; Kinoshita et al., 1991). In the late Pliocene to early Pleistocene, the Miura forearc basin fills, however, were far from the Plio-Pleistocene volcanic front (Mori et al., 2012), and in the present eastern Sagami Bay, which is similarly far from the volcanic front, there is no seep fluid temperature anomaly relative to the surrounding seawater at cold-seep sites (Fujikura et al., 1995). Thus, the oxygen isotope ratios of the observed dolomites cannot be explained by warmer seep fluids in the Miura forearc basin fills.

Hypothesis of 18O-depleted seep fluids.—It is possible that the seep fluids were influenced by land-derived groundwater or residual waters from gas-hydrate formation, either of which could cause seep fluids to be depleted in 18O. The occurrence of 18O-depleted ground-water at the Loc. 3 cold-seep site may be likely, because the fluid chemistry at the Hatsushima cold-seep site shows chloride depletion, suggesting a major effect of land-derived groundwater (Tsunogai et al., 1996).

The paleobathymetry of the Urago Formation (between 400 and 600 m; Utsunomiya and Majima. 2012) is at the upper limit of the water depth where gas hydrates are stable (Field and Kvenvolden, 1985). During gas hydrate formation, O18 is preferentially taken up by the gas hydrate (Maekawa, 2004), causing the residual waters to be depleted in 18O. Thus, we cannot reject either of these two explanations for the 18O-depleted seep fluids, but in the absence of direct evidence, no conclusion is possible.

Discussion on the occurrence modes of vesicomyiddominated assemblage at Loc. 3

The vesicomyid-dominated assemblage described in this paper is characterized by (1) an exclusive occurrence in cross-bedded and massive sandstones interpreted as sand-ridge deposits (Figure 4.2), (2) the presence of articulated bivalves in life position within the upper massive sandstone horizons (Figure 6.1), and (3) an association with 13C-depleted authigenic carbonates (δ13C = -37.78‰ to -24.16‰ VPDB; Figure 9.1 and Table 1). clearly suggesting methane seepage (Figure 9.1 and Table 1). We consider all of the chemosymbiotic bivalves from Loc. 3 described here to have lived on sand dunes and depended on the methane seepages evidenced by the 13C-depleted authigenic carbonates, but they were sometimes reworked by episodic bottom currents that produced the cross bedding. Many of the shells in the lower cross-bedded sandstone horizons might have been transported from beyond the exposure site, whereas most of those in the upper massive sandstone horizons were probably not transported very far, as some are preserved in their life position (Figure 6.1).

Among the vesicomyid-dominated fossil assemblages from the forearc basin fills exposed on the Miura Peninsula, the Loc. 3 assemblage is the only example preserved in the original cold seep site where they flourished, lacking any evidence of physical reworking by a submarine landslide (Type A: Figure 3) or by a sediment gravity flow (Type B: Figure 3). The mode of occurrence of the Loc. 3 assemblage newly suggests that the active cold seepages had occurred in the sand ridge deposits and maintained the seep-dependent organisms.

Living Calyptogena (Archivesica) kawamurai (Kuroda) have been sampled from cold seep sites at depths of 300 to 710 m, in the forearc regions of the southwestern Honshu arc and the Ryukyu arc (Fujikura et al., 2000; Numanami et al., 2002). Despite the fact that the water depths of their habitats are consistent with the paleobathymetry of the Urago Formation (between 400 and 600 m water depth; Utsunomiya and Majima, 2012). they could hardly be preserved in the sedimentary record and are not comparable with the Loc. 3 assemblage because the present-day colonies flourish on the top and slope of tectonically formed highs where surface sediments are mostly eroded and the brecciated basement rocks and authigenic carbonate concretions are exposed (Matsumoto et al., 1999; Machiyama et al., 2001a, b). In the fossil record, Calyptogena (Archivesica) kawamurai (Kuroda) associated with 13C-depleted authigenic carbonates have also been reported from the mudstones of the Pleistocene forearc basin fills in the Kakegawa area (Nobuhara and Tanaka, 1993; Nobuhara, 2003). but an in situ fossil occurrence like that of the Loc. 3 assemblage in sand ridge deposits has not been reported. Generally speaking, widespread permeable and coarse substrates prevent the flourishing of a cold-seep-dependent assemblage, because in that situation the methane from the seep easily diffuses into the water column and the adjacent permeable substrate and thus does not become concentrated. The methane diffusion decreases the AOM rate, and thus the concentration of hydrogen sulfide necessary to maintain seep-dependent organisms in interstitial waters near the seafloor (e.g. Borowski et al., 1996). To explain this apparent inconsistency, that is, chemosymbiotic bivalves living on coarse and permeable sand dunes, we speculate that the sandy substrate on which the chemosymbiotic bivalves lived was covered by thin offshore muds in between episodes of sand deposition. Such muds, which would have effectively impeded methane diffusion from the seafloor until the next depositional event, can thus explain the present-day modes of occurrence of the shells at Loc. 3. We therefore suggest that the Loc. 3 assemblage was able to live in the sand-ridge deposits formed by the migration of dunes under a unidirectional bottom current. This site is thus a rare example of cold-seep organisms that lived on dune deposits in a high-energy depositional environment.

Conclusions

The fossil localities of the vesicomyid-dominated fossil assemblages in forearc basin fills on the Miura Peninsula can be classified into three types: Type A, in association with submarine landslide deposits; Type B, reworked shells in a sandstone bed deposited by a sediment gravity flow; and Type C, in situ occurrence in sand-ridge deposits. The characteristics of the Type C assemblage and the lowermost Pleistocene Urago Formation where it was observed are summarized as follows: 1. The assemblage consists mostly of Calyptogena (Archivesica) kawamurai (Kuroda) with a minor amount of Conchocele bisecta (Conrad). In the studied outcrop, the shells occur in the lower cross-bedded and the upper massive sandstone horizons. Orientations of the shells in the cross-bedded sandstones are nearly concordant with those of erosional surfaces or foreset laminae. The long axes of the shells are aligned nearly perpendicular to the current direction estimated from the orientation of the foreset laminae, suggesting that they were dragged and came to rest in the position of maximum friction against the sandy bottom. Some articulated vesicomyid shells in the massive sandstones are oriented perpendicular to the bedding plane with their anterior ends pointing downward (life position).

2. Authigenic carbonates are well developed in the vesicomyid-bearing sandstones, whereas they occur sporadically in the underlying muddy sandstones and sandstones, neither of which contains shells. The biogenic carbonates have entirely dissolved, and some of the molds are filled with dolomite-cemented fine clastics. The authigenic carbonates are composed exclusively of dolomite with δ13C of -37.78‰ to -24.16‰ VPDB and δ18O of 0.69‰ to 4.35‰ VPDB.

We consider this fossil assemblage to consist of in situ cold-seep-dependent organisms living on sand dunes, because some shells appear to be in their original habitat, and because they are associated with 13C-depleted authigenic carbonates, indicating the influence of AOM.

Acknowledgements

For support of our field surveys, we are grateful to T. Ogawa, Y. Ogawa, I. Matsukawa and Y. Ebashi, who are residing in Kamakura City, and the Kamakura Public Golf Club, with particular thanks to R. Wani and A. Nozaki (Yokohama National University), R. G. Jenkins (Kanazawa University), T. Shibata (engineering geologist), D. Onda (Higashi Yamato High School), S. Inoue (Asia Air Survey. Co., Ltd.), K. Nishida (AIST), T. Saito (Nichirei Foods Inc.) and K. Sato (University of Tokyo). We thank M. Satish-Kumar (Niigata University), N. Hamana (Shinkoshuppansha Keirinkan Co., Ltd.), Y. Osada (Japan Environment Research Co., Ltd.), and T. Tsuboi (Shizuoka Pref.) for their help with the isotope analyses and for helpful suggestions about the stable isotope data. We also thank S. Kawagata (Yokohama National University) for his advice on scanning electron microscopy, and M. Arima, M. Ishikawa, and T. Ichiki (Yokohama National University) for discussions in our seminars. K. Amano (Joetsu University of Education), Y. Iryu (Tohoku University), T. Ubukata (Kyoto University) and an anonymous reviewer are appreciated for their critical and helpful comments to improve our manuscript. This study was supported by a Grant-in-Aid for Scientific Research (A: No. 16204041 and B: No. 20403015) and a Grant-in-Aid for JSPS Fellows (25–75) from the Japan Society for the Promotion of Sciences.

References

1.

J. R. I. Allen , 1992: Transport hydrodynamics: shells. In , D. E. G. Briggs and P. R. Crowther eds., Palaeobiology: A Synthesis , p. 227–230. Blackwell Scientific Publications, Oxford. Google Scholar

2.

K. Amano and S. Kiel , 2010: Taxonomy and distribution of fossil Archivesica (Bivalvia: Vesicomyidae) in Japan. Nautilus , vol. 124, p. 155–165. Google Scholar

3.

W. S. Borowski , C. K. Paull and W. Ussler III , 1996: Marine porewater sulfate profiles indicate in situ methane flux from underlying gas hydrate. Geology , vol. 24, p. 655–658. Google Scholar

4.

A., H. Bouma, 1962: Sedimentology of Some Flysch Deposits: A Graphic Approach to Facies Interpretation , 168 p. Elsevier, Amsterdam. Google Scholar

5.

P. J. Brenchley and G. Newall , 1970: Flume experiments on the orientation and transport of models and shell valves. Palaeogeography, Palaeoclimatology, Palaeoecology , vol. 7, p. 185–220. Google Scholar

6.

K. A. Campbell , 2006: Hydrocarbon seep and hydrothermal vent paleoenvironments and paleontology: past developments and future research directions. Palaeogeography, Palaeoclimatology, Palaeoecology , vol. 232, p. 362–407. Google Scholar

7.

K. A. Campbell , C. S. Nelson , A. C. Alfaro , S. Boyd , J. Greinert , S. Nyman , E. Grosjean , G. A. Logan , M. R. Gregory , S. Cooke , P. Linke , S. Milloy and I. Wallis , 2010: Geological imprint of methane seepage on the seabed and biota of the convergent Hikurangi Margin, New Zealand: Box core and grab carbonate results. Marine Geology , vol. 272, p. 285–306. Google Scholar

8.

C. Canet , R. M. Prol-Ledesma , E. Escobar-Briones , C. Mortera-Gutierrez , R. Lozano-Santa Cruz , C. Linares , E. Cienfuegos and P. Morales-Puente , 2006: Mineralogical and geochemical characterization of hydrocarbon seep sediments from the Gulf of Mexico. Marine and Petroleum Geology , vol. 23, p. 605–619. Google Scholar

9.

E. V. Coan , P. H. Scott and F. R. Bernard , 2000: Bivalve Seashells of Western North America. Marine Bivalve Mollusks from Arctic Alaska to Baja California , 764 p. Santa Barbara Museum of Natural History, Santa Barbara. Google Scholar

10.

T. B. Coplen , C. Kendall and J. Hoppe , 1983: Comparison of stable isotope reference samples. Nature , vol. 302, p. 236–238. Google Scholar

11.

T. Eto , 1986a: Stratigraphy of the Hayama Group in the Miura Peninsula, Japan. Science Report of Yokohama National University, Section II , no. 33, p. 67–105. ( in Japanese with English abstractGoogle Scholar

12.

T. Eto , 1986b: Stratigraphy of the Miura and Kazusa Groups in the Miura Peninsula, Japan. Science Report of Yokohama National University, Section II , no. 33, p. 107–132. ( in Japanese with English abstractGoogle Scholar

13.

T. Eto , 1993: Geology of cutting outcrops in the Ikego U. S. facilities area in Zushi City, Kanagawa Prefecture. Science Report of Yokohama National University, Section II , no. 40, p. 31–51. ( in Japanese with English abstractGoogle Scholar

14.

T. Eto , M. Oda , S. Hasegawa , N. Honda and M. Funayama , 1987: Geologic age and paleoenvironment based upon microfossils of the Cenozoic sequence in the middle and northern part of the Miura Peninsula. Science Report of Yokohama National University, Section II , no. 34, p. 41–57. (in Japanese with English abstractGoogle Scholar

15.

T. Eto , K. Yazaki , A. Urabe and I. Isobe , 1998: Explanatory Text of the Geological Map of Japan, Scale 1:50,000, Yokosuka District (Tokyo-84) , 128 p. Geological Survey of Japan, Tsukuba. ( in Japanese with English abstractGoogle Scholar

16.

M. E. Field and K. A. Kvenvolden , 1985: Gas hydrates on the northern California continental margin. Geology , vol. 13, p. 517–520. Google Scholar

17.

K. Fujikura , J. Hashimoto , Y. Fujiwara and T. Okutani , 1995: Community ecology of the chemosynthetic community at off Hatsushima Site, Sagami Bay, Japan. JAMSTEC Journal of Deep Sea Research , no. 11, p. 227–241. ( in Japanese with English abstractGoogle Scholar

18.

K. Fujikura , J. Hashimoto and T. Okutani , 2002: Estimated population densities of megafauna in two chemosynthesis-based communities: a cold seep in Sagami Bay and a hydrothermal vent in the Okinawa Trough. Benthos Research , vol. 57, p. 21–30. Google Scholar

19.

K. Fujikura , S. Kojima , Y. Fujiwara , J. Hashimoto and T. Okutani , 2000: New distribution records of vesicomyid bivalves from deep-sea chemosynthesis-based communities in Japanese waters. Venus (Journal of the Malacological Society of Japan), vol. 59, p. 103–121. Google Scholar

20.

R. S. Hanson and T. E. Hanson , 1996: Methanotrophic bacteria. Microbiological Reviews , vol. 60, p. 439–471. Google Scholar

21.

M. Hattori , T. Oba , Y. Kanie and K. Akimoto , 1994: Authigenic carbonates collected from cold seepage area off Hatsushima Island, Sagami Bay, central Japan. JAMSTEC Journal of Deep Sea Research , no. 10, p. 405–416. ( in Japanese with English abstractGoogle Scholar

22.

D. Hirata , Y. Matsushima and S. Asaga , 1991: Distributions and mode of occurrences of fossil Calyptogena in Miura and Boso peninsulas. Chikyu Monthly , no. 139, p. 47–52. ( in Japanese; original title translatedGoogle Scholar

23.

J. Hoefs , 2004: Stable Isotope Geochemistry, Fifth Revised and Updated Edition , 244 p. Springer-Verlag, Berlin. Google Scholar

24.

S. Inagaki , T. Nishikawa , T. Mitsuoka and M. Yasuno , 2007: Garnetbearing pumice bed, KGP, in the lower Kazusa Group from the northeastern area of Kamakura City, Kanagawa Prefecture, central Japan. Earth Science (Chikyu Kagaku) , vol. 61, p. 143–148. ( in Japanese with English abstractGoogle Scholar

25.

R. Iwase , H. Machiyama , W. Soh , K. Brown and M. Tryon , 2003: Detailed mapping of heat flow and flow rate observation at the seepage site off Hatsushima Island in Sagami Bay. Abstracts of Japan Geoscience Union Meeting 2003 , no. T032-002, Chiba, JapanGoogle Scholar

26.

Japan Oceanographic Data Center, 2013: J-DOSS, Oceanographic Data and Information Download Service (Temperature, Current, Depth, Biology, Marine Information) [online]. [Cited 15 August 2013]. Available from:  http://www.jodc.go.jp/index_j.htmlGoogle Scholar

27.

M. Kinoshita , M. Yamano and S. Makita , 1991: High heat flow anomaly around Hatsushima biological community in the western Sagami Bay, Japan. Journal of Physics of the Earth , vol. 39, p. 553–571. Google Scholar

28.

T. Kitazaki and R. Majima , 2003: A slope to outer-shelf cold-seep assemblage in the Plio-Pleistocene Kazusa Group, Pacific side of central Japan. Paleontological Research , vol. 7, p. 279–296. Google Scholar

29.

K. Koike , 1951: On “The Kurotaki Unconformity”. Journal of the Geological Society of Japan , vol. 57, p. 143–156. ( in Japanese with English abstractGoogle Scholar

30.

S. Kojima , 2002: Deep-sea chemoautosynthesis-based communities in the northwestern Pacific. Journal of Oceanography , vol. 58, p. 343–363. Google Scholar

31.

E. M. Krylova and H. Sahling , 2010: Vesicomyidae (Bivalvia): Current taxonomy and distribution. PLoS ONE , vol. 5, e9957. Google Scholar

32.

D. N. Lumsden , 1979: Discrepancy between thin section and X-ray estimates of dolomite non-stoichiometry and carbonate facies parameters. Journal of Sedimentary Petrology , vol. 48, p. 429–436. Google Scholar

33.

H. Machiyama , T. Matsumoto , R. Matsumoto , M. Hattori , M. Okano , R. Iwase and H. Tomaru , 2001a: Outline of Shinkai 2000 dive surveys on the Kuroshima Knoll, off Ishigaki Island—Preliminary report of the NT00-05 Cruise—. JAMSTEC Journal of Deep Sea Research , no. 19, p. 45–60. ( in Japanese with English abstractGoogle Scholar

34.

H. Machiyama , R. Shinjo , M. Hattori , M. Okano , T. Matsumoto , M. Kimura , M. Nakamura , G. Toyama and T. Okada , 2001b: Outline of Dolphin-3K dive surveys on the Kuroshima Knoll and in the southern Okinawa Trough—Preliminary report of the NT00-05 Cruise—. JAMSTEC Journal of Deep Sea Research , no. 18, p. 15–30. ( in Japanese with English abstractGoogle Scholar

35.

T. Maekawa , 2004: Experimental study on isotopic fractionation in water during gas hydrate formation. Geochemical Journal , vol. 38, p. 129–138. Google Scholar

36.

R. Majima , 1999: Mode of occurrences of the Cenozoic chemosynthetic communities in Japan. Memoirs of the Geological Society of Japan , vol. 54, p. 117–129. ( in Japanese with English abstractGoogle Scholar

37.

R. Majima , T. Nobuhara and T. Kitazaki , 2005: Review of fossil chemosynthetic assemblages in Japan. Palaeogeography, Palaeoclimatology, Palaeoecology , vol. 227, p. 86–123. Google Scholar

38.

R. Majima , Y. Tate and T. Shibasaki , 1996: In situ fossil chemosynthetic community found from the Lower Pleistocene Kazusa Group, Yokohama City, central Japan. Fossils (Palaeontological Society of Japan) , vol. 61, p. 47–54. ( in Japanese with English abstractGoogle Scholar

39.

T. Matsumoto , M. Kimura , S. Nishida , T. Nakamura and T. Ono , 1999: Chemosynthetic communities and surface ruptures discovered on the Kuroshima Knoll south of Yaeyama Islands (NT97-14 Cruise). JAMSTEC Journal of Deep Sea Research , no. 14, p. 477–491. ( in Japanese with English abstractGoogle Scholar

40.

A. Matthews and A. Katz , 1977: Oxygen isotope fractionation during the dolomitization of calcium carbonate. Geochemica et Cosmochimica Acta , vol. 41 p. 1431–1438. Google Scholar

41.

D. P. McKenzie and W. J. Morgan , 1969: Evolution of triple junctions. Nature , vol. 224, p. 125–133. Google Scholar

42.

T. Mitsunashi and T. Kikuchi , 1982: Explanatory Text of the Geological Map of Japan, Scale 1:50,000, Yokohama District (Tokyo-74) , 105 p. Geological Survey of Japan, Tsukuba. ( in Japanese with English abstractGoogle Scholar

43.

S. Mori , H. Yamashita , M. Arima and K. Fujioka , 2012: K-Ar ages of volcanic rocks in the Tanzawa-Oiso area, central Japan: Implication for westward migration of the Pliocene volcanic front in the southern Fossa Magna region. Japanese Magazine of Mineralogical and Petrological Sciences , vol. 41. p. 67–86. ( in Japanese with English abstractGoogle Scholar

44.

T. H. Naehr , P. Eichhubl , V. J. Orphan , M. Hovland , C. K. Paull , W. Ussler III, T. D. Lorenson and H. G. Greene , 2007: Authigenic carbonate formation at hydrocarbon seeps in continental margin sediments: A comparative study. Deep-Sea Research II , vol. 54, p. 1268–1291. Google Scholar

45.

Y. Nagai and H. Wada , 1993: Development and applications of micro-scale isotopic analyses for metamorphic dolomites. Geoscience Reports of Shizuoka University , no. 19, p. 25–34. ( in Japanese with English abstractGoogle Scholar

46.

Y. Naganuma , J. Arai , Y. Kanie and H. Nagahama , 1973: Paleocurrent of the basal strata of the Kazusa Group, Miura Peninsula. The 80th Annual Meeting of the Geological Society of Japan, Abstract , p. 123. (in Japanese; original title translatedGoogle Scholar

47.

N. Nakayama and F. Masuda , 1989: Ocean current-controlled sedimentary facies of the Pleistocene Ichijiku Formation, Kazusa Group, Boso Peninsula, Japan. In , A. Taira and F. Masuda eds., Sedimentary Facies in the Active Plate Margin , p. 275–293. TERRAPUB, Tokyo. Google Scholar

48.

N. Niitsuma , Y. Matsushima and D. Hirata , 1989: Abyssal molluscan colony of Calyptogena in the Pliocene strata of the Miura Peninsula, central Japan. Palaeogeography, Palaeoclimatology, Palaeoecology , vol. 71, p. 193–203. Google Scholar

49.

T. Nobuhara , 2003: Cold seep carbonate mounds with Vesicomya (Calyptogena) kawamurai (Bivalvia: Vesicomyidae) in slope-mud facies of the Pliocene forearc basin of the Sagara-Kakegawa area, central Japan. Paleontological Research , vol. 7, p. 313–328. Google Scholar

50.

T. Nobuhara and T. Tanaka , 1993: Palaeoecology of Akebiconcha kawamurai (Bivalvia: Vesicomyidae) from the Pliocene Tamari Silt Formation in the Kakegawa area, central Japan. Palaeogeography, Palaeoclimatology, Palaeoecology , vol. 102, p. 27–40. Google Scholar

51.

D. A. Northrop and R. N. Clayton , 1966: Oxygen-isotope fractionations in systems containing dolomite. Journal of Geology , vol. 74, p. 174–196. Google Scholar

52.

H. Numanami , K. Fujikura and J. Hashimoto , 2002: Akebiconcha kawamurai collected from Suruga Bay, and a possibility of existence of chemosynthetic community. Venus (Journal of the Malacological Society of Japan) , vol. 61, p. 121–122. Google Scholar

53.

J. R. O'Neil and S. Epstein , 1966: Oxygen isotope fractionation factor in the system dolomite-calcite-carbon dioxide. Science , vol. 152, p. 198–201. Google Scholar

54.

K. Oyama , 1951: On the fossil biocoenosis of the Koshiba Formation. Miscellaneous Reports of Research Institute for Natural Resources , no. 24, p. 55–59. ( in Japanese with English abstractGoogle Scholar

55.

T. Shikama and A. Masujima , 1969: Quantitative studies of the molluscan assemblages in the Ikego-Nojima formations. Science Report of Yokohama National University, Section II , no. 15, p. 61–94. Google Scholar

56.

M. Sibuet and K. Olu , 1998: Biogeography, biodiversity and fluid dependence of deep-sea cold-seep communities at active and passive margins. Deep Sea Research II , vol. 45, p. 517–567. Google Scholar

57.

W. Soh , K. T. Pickering , A. Taira and H. Tokuyama , 1991: Basin evolution in the arc-arc Izu Collision Zone, Mio-Pliocene Miura Group, central Japan. Journal of the Geological Society (London) , vol. 148, p. 317–330. Google Scholar

58.

M. Takahashi and K. Saito , 1997: Miocene intra-arc bending at an arc-arc collision zone, central Japan. Island Arc , vol. 6, p. 168–182. Google Scholar

59.

R. Takeuchi , H. Machiyama and R. Matsumoto , 2007: The formation process of the cold seep carbonates at the Kuroshima Knoll. JAMSTEC Journal of Deep Sea Research , no. 19, p. 61–75. ( in Japanese with English abstractGoogle Scholar

60.

I. Tamura and H. Yamazaki , 2010: Significance of the remarkable unconformity in the Plio-Pleistocene of the Japanese islands. Quaternary International , vol. 219, p. 45–54. Google Scholar

61.

Y. Tate and R. Majima , 1998: A chemosynthetic fossil community related to cold seeps in the outer shelf environment. A case study in the Lower Pleistocene Koshiba Formation, Kazusa Group, central Japan. Journal of the Geological Society of Japan , vol. 104, p. 24–41. ( in Japanese with English abstractGoogle Scholar

62.

U. Tsunogai , J. Ishibashi , H. Wakita , T. Gamo , T. Masuzawa , T. Nakatsuka , Y. Nojiri and T. Nakamura , 1996: Fresh water seepage and pore water recycling on the seafloor: Sagami Trough subduction zone, Japan. Earth and Planetary Science Letters , vol. 138, p. 157–168. Google Scholar

63.

M. Utsunomiya and R. Majima , 2012: Paleobathymetries of the Plio-Pleistocene Urago and Nojima Formations, Kazusa Group, Miura Peninsula, central Japan: Revision on the basis of molluscan fossils from new localities. Fossils (Palaeontological Society of Japan) , vol. 91, p. 5–14. ( in Japanese with English abstractGoogle Scholar

64.

M. Utsunomiya , C. Nagahama , R. G. Jenkins, A. Nozaki and R. Majima , 2014: A shell-concentrated sandstone bed containing vesicomyid shells from the Lower Pleistocene Nojima Formation, Kazusa Group, Pacific side of central Japan. Journal of the Geological Society of Japan , vol. 120, p. 221–230. ( in Japanese with English abstractGoogle Scholar

65.

C. Vasconcelos , J. A. McKenzie , R. Warthmann and S. M. Bernasconi , 2005: Calibration of the δ18O paleothermometer for dolomite precipitated in microbial cultures and natural environments. Geology , vol. 33, p. 317–320. Google Scholar

66.

R. G. Walker , 1992: Turbidites and submarine fans. In , R. G. Walker and N. P. James eds., Facies Models: Response to Sea Level Change , p. 239–263. Geological Association of Canada, St. John's, Newfoundland. Google Scholar

67.

M. J. Whiticar , 1999: Carbon and hydrogen isotope systematics of bacterial formation and oxidation of methane. Chemical Geology , vol. 161, p. 291–314. Google Scholar

68.

R. B. Wynn and D. A. V. Stow , 2002: Classification and characterization of deep-water sediment waves. Marine Geology , vol. 192, p. 7–22. Google Scholar

69.

Yokohama Defense Facilities Administration Bureau, 1993: Investigation of Calyptogena Fossils at Ikego, Final Report , 470 p. Yokohama Defense Facilities Administration Bureau, Yokohama. (in Japanese; original title translatedGoogle Scholar
© by the Palaeontological Society of Japan
Masayuki Utsunomiya, Ryuichi Majima, Kiminori Taguchi, and Hideki Wada "An in situ Vesicomyid-dominated Cold-seep Assemblage from the Lowermost Pleistocene Urago Formation, Kazusa Group, Forearc Basin Fill on the Northern Miura Peninsula, Pacific Side of Central Japan," Paleontological Research 19(1), 1-20, (1 January 2015). https://doi.org/10.2517/2014PR024
Received: 27 April 2013; Accepted: 1 May 2014; Published: 1 January 2015
KEYWORDS
cold-seep fossil assemblage
Kazusa Group
Pleistocene
Urago Formation
vesicomyid
Back to Top