Journal of Petrology Pages 1827-1851 © 1999 Oxford University Press

Polybaric Petrogenesis of Mafic Layers in the Horoman Peridotite Complex, Japan
Introduction
Geological Background
Field Occurrence Of Mafic Layers
Petrography
   Rock types
   Type I and Type II mafic granulites
   Type V mafic granulite
   Plagioclase websterite
Analytical Procedure
Results
   Major elements and compatible trace elements
   Incompatible trace elements
   Sr, Nd and Pb isotope compositions
Discussion
   Do the mafic layers represent melt compositions?
   Effects of interaction between host peridotite and mafic layers
A cumulate model for the Horoman mafic layers
   Ages and source regions of the mafic layers in the Horoman peridotite
   Tectonic implications based on the Horoman mafic layers
Major Conclusions
Acknowledgements
References

Footnote Table

Polybaric Petrogenesis of Mafic Layers in the Horoman Peridotite Complex, Japan

E. TAKAZAWA1*, F. A. FREY1, N. SHIMIZU2, A. SAAL1,2 AND M. OBATA3

1MASSACHUSETTS INSTITUTE OF TECHNOLOGY, CAMBRIDGE, MA 02139, USA
2WOODS HOLE OCEANOGRAPHIC INSTITUTION, WOODS HOLE, MA 02543, USA
3KYOTO UNIVERSITY, KYOTO, 606-8502, JAPAN

RECEIVED MARCH 25, 1998; REVISED TYPESCRIPT ACCEPTED MAY 27, 1999

Two major types of mafic granulite layers occur within the Horoman peridotite, an 8 km * 10 km * 3 km orogenic lherzolite exposed in the high-T and low-P Hidaka metamorphic belt of Hokkaido, Japan. The mineral assemblages and textures of these layers reflect subsolidus reactions occurring during uplift from the upper mantle to the crust. Nevertheless, their whole-rock compositions can be used to infer the primary mineralogy of these layers, and a genetic relationship to melts geochemically similar to mid-ocean ridge basalts (MORB). The intralayer compositional variation of Type I layers (Al-Ti augite type mafic granulites) shows that the centers formed as garnet clinopyroxenites in equilibrium with an incompatible element depleted melt that crystallized to form the margins. In contrast, the Type II layers (Cr-diopside type mafic granulites) formed at relatively shallow depths and are much older, ~830 Ma, than the Type I garnet pyroxenites, which formed at ~80 Ma. The temporal sequence supports the hypothesis that the Horoman peridotite represents shallow MORB-related oceanic mantle that had subsided to deeper mantle depths before crustal emplacement.

Keywords: orogenic lherzolite;mafic layer; pyroxenite; trace element; isotope

INTRODUCTION

Upper-mantle peridotites exposed in the Earth's crust are commonly interlayered by `mafic layers' such as pyroxenites and olivine gabbros (mafic granulites), which may parallel or cross-cut the foliation of the host peridotite (e.g. Nicolas & Jackson, 1982). [In this paper we use `mafic layers' to describe pyroxenites to `gabbroic' rocks (mafic granulites) whose composition and mineral assemblage differs substantially from the host peridotites.] Hypotheses for the origin of these layers in orogenic peridotites include: (1) solidification of melt (Niida, 1984; Shiotani & Niida, 1997); (2) crystal cumulates in melt conduits or sills (Loubet & Allègre, 1982; Bodinier et al., 1987; Suen & Frey, 1987; Bodinier, 1988; Pearson et al., 1993; Kumar et al., 1996); (3) recycling of ancient subducted oceanic crust that was dispersed and thinned during mantle convection (Polvé & Allègre, 1980; Allègre & Turcotte, 1986). Subsequent to their formation these layers may have experienced partial melting (Loubet & Allègre, 1982) and interacted with metasomatic fluids (Garrido & Bodinier, 1999). Ultimately, some mafic layers may be important source components for basalts (e.g. Hirschmann & Stolper, 1996).

In addition to layers within massive peridotites, pyroxenites also occur as discrete xenoliths and as portions of composite pyroxenite-peridotite xenoliths in volcanic rocks (Nixon, 1987). Two major types of pyroxenites have been recognized, the Al-Ti-augite type and the Cr-diopside type (Wilshire & Shervais, 1975). These types of mafic layers occur in orogenic peridotites such as the Ronda (Obata, 1980; Suen & Frey, 1987; Garrido & Bodinier, 1999) and the Beni Bousera (Kornprobst, 1969; Kornprobst et al., 1990a; Pearson et al., 1993; Kumar et al., 1996). Similarly the Horoman peridotite, a fault-bounded 8 km*10 km *3 km mass of upper mantle exposed in Hokkaido, Japan (Niida, 1974) contains two dominant types of mafic rock, which were referred to as Gabbro I (Al-Ti augite type) and Gabbro II (Cr-diopside type) by Niida, (1984).

To determine the origin of these layers, including their depth of formation, the temporal sequence of the two types, and the interaction between the mafic layers and wall-rock peridotite, we determined major and trace element abundances and Sr, Nd and Pb isotopic data for mafic layers in the Horoman peridotite.

GEOLOGICAL BACKGROUND

The Horoman ultramafic complex occurs in the high-T and low-P Hidaka metamorphic belt of Hokkaido, Japan (Fig. 1). This metamorphic belt is subdivided into two zones, the Main and Western zones (Komatsu et al., 1983, , 1989). The Main Zone consists of metamorphic rocks, up to granulite facies rocks (Osanai, 1985; Osanai et al., 1991), associated with igneous intrusives (Maeda et al., 1986; Owada, 1989; Shimura et al., 1992) and ultramafic rocks (Komatsu & Nochi, 1966; Niida, 1974). It is interpreted as an island-arc type crustal section of Early Eocene to Early Miocene age, which has been thrust over the Western Zone along the Hidaka Main Thrust (Komatsu et al., 1982, , 1989; Toyoshima, 1991). The Western Zone consists of greenschist to amphibolite facies metamorphosed mafic rocks, ultramafic cumulates and ultramafic tectonite; it has a complete ophiolite sequence (Miyashita, 1983).


Figure 1. Simplified geological map of the Hidaka metamorphic belt after Komatsu et al., (1986). Inset shows location of this metamorphic belt on Hokkaido.


The Horoman complex occurs in the southwestern tip of the Main Zone (Fig. 1). In the west it is juxtaposed with the Cretaceous Hidaka Super Group, which is interpreted to be an accretional prism formed in a subduction zone (Fig. 2). There is a narrow fault zone of a few hundred meters thick between the peridotite and the Hidaka Super Group. This zone consists of sheared metagabbro, greenschist and blackschist, and is considered as a part of the Western Zone of the Hidaka metamorphic belt. On its eastern and southern sides, the Horoman complex is in a fault contact with metamorphic rocks and mafic intrusives of the Main Zone. Granulite facies metamorphism has been recognized in the Main Zone metamorphic rocks near the peridotite contact (Komatsu et al., 1981).


Figure 2. Simplified map of the Horoman peridotite showing sample locations for mafic layers. Estimated boundary between the Lower Zone and Upper Zone is marked by a dashed line. The north and south Fudo sections (Y-shaped set of parallel lines) were used to construct a representative section of the Upper Zone (Fig. 3). Inset shows location of the studied area in the Horoman peridotite.


In the Horoman complex, foliation and lineation are manifested chiefly by elongated direction of minerals, such as pyroxenes and spinel, and fine-grained mineral seams. Based on the foliation, the Horoman complex has a gently waving synclinal structure with its axis striking to the west in the central part of the complex (Komatsu & Nochi, 1966; Niida, 1974; Sawaguchi & Takagi, 1997). Towards the south, the foliation defines a large-scale half-dome structure opening southward, whereas to the north there is a nearly monoclinal structure; that is, the foliation strikes generally northwest and dips southwest (Niida, 1974). The lineation directions range from N15E to N15W with variable dips (Niida, 1975).

Takahashi, (1991) identified three petrographically distinct peridotite suites in the Horoman peridotite. The Main Harzburgite-Lherzolite (MHL) suite is the major part of the massif. The two other peridotite suites, Banded Dunite-Harzburgite and Spinel-rich Dunite-Wehrlite are inferred to be cumulate rocks (Takahashi, 1991). The peridotites of the complex show a well-developed layered structure. Based on the style of layering, the Horoman peridotite mass is subdivided into two main stratigraphic zones (Fig. 2), the Upper Zone (~600 m thick) and the Lower Zone (~2200 m thick) (Komatsu & Nochi, 1966; Niida, 1974). The boundary between the two zones is gradational and not a fault contact. The equilibrium temperatures recorded in the cores of orthopyroxene porphyroclasts increase from the Lower Zone to the Upper Zone. Ozawa & Takahashi, (1995) suggested that the temperature variation may reflect the thermal gradient from the core (Upper Zone) to the margin (Lower Zone) of an ascending mantle diapir.

The Lower Zone consists of several cyclic layers, tens to hundreds of meters in thickness, in the sequence: plagioclase lherzolite-lherzolite-harzburgite-dunite (not always present)-harzburgite-lherzolite-plagioclase lherzolite (Komatsu & Nochi, 1966; Niida, 1974; Obata & Nagahara, 1987; Takahashi, 1991). Mafic layers are scarce and are typically no more than a few centimeters thick, but a thick mafic layer, up to a few meters in thickness, occurs in a plagioclase lherzolite layer.

At a scale of tens of meters, the Upper Zone can be subdivided into units formed of the major rock types, plagioclase lherzolite and harzburgite (Fig. 3). Although each unit may be traced laterally over several kilometers, centimeter-scale individual layers within a unit gradually thin and disappear within a few hundred meters. At the centimeter scale, the Upper Zone is characterized by fine-scale layering of peridotites and abundant mafic layers (Fig. 4). Plagioclase lherzolite and harzburgite are the most dominant rock types in the Upper Zone, with subordinate amounts of lherzolite, dunite and mafic rocks; plagioclase lherzolite and mafic layers are more abundant in the Upper Zone than in the Lower Zone (Table 1). The lithological boundaries among plagioclase lherzolite, lherzolite and harzburgite are sharp in the field, particularly on weathered surfaces, but they are microscopically gradational over a few centimeters. Harzburgite close to lherzolite layers in the Upper Zone frequently contains fine-grained mineral seams, which consist of spinel, orthopyroxene, clinopyroxene and fine-grained spinel-pyroxene symplectite. These seams (<1 cm thick) range from a few centimeters to 20 cm in length.


Table 1. Estimated volume of rock types (values are percentages)


Figure 3. Representative section of the Upper Zone constructed along the Fudo section indicated in Fig. 2. Arrows A-E indicate horizons whose fine-scale layering is shown in Fig. 4.



Figure 4. Stratigraphic sections in the Upper Zone of the Horoman peridotite, showing the fine-scale lithological layering. The scales for sections B, C and D are similar to that of section A. See Fig. 3 for the location of each section. Arrows and sample numbers indicate the locations of analyzed samples (Table 2). Rock types are Type I mafic granulite, Type II mafic granulite and plagioclase websterite (Pl Web) (see the Petrography section of text).


FIELD OCCURRENCE OF MAFIC LAYERS

Mafic layers in the Horoman complex range from several millimeters to several meters in thickness and their abundance increases with decreasing thickness (Fig. 5); an observation that is typical of upper-mantle rocks (Allègre & Turcotte, 1986). In the Upper Zone of the Horoman complex there are at least five subzones containing abundant mafic layers ( Fig. 3). In such zones, many mafic layers of widely varying thickness are closely spaced and interlayered with peridotites (Fig. 4). The contacts between mafic layers and peridotites are typically sharp in the field.


Figure 5. Layer thickness (X) vs cumulative number (N) of layers that are thicker than the value (X) on the horizontal axis. D-H-L line is for layers in dunite, harzburgite or lherzolite, and PL line is for layers in plagioclase lherzolite.


Most of the mafic layers in the Horoman complex are parallel to the foliation plane of the peridotites, and many mafic layers occur as flat planes extending more than tens of meters parallel to other layers. Some mafic layers in the Upper Zone show folding structures, and wavy or asymmetric lithological boundaries on various scales. Folded layers contain late-stage foliation planes parallel to those in the peridotite and unfolded mafic layers. These structures may be caused by the top-to-the-south shear deformation discussed by Sawaguchi & Takagi, (1997).

In the Upper Zone, harzburgites are the dominant wallrocks of the mafic layers (Figs 3 and 4), typically forming a sequence: mafic layer-harzburgite-plagioclase lherzolite; whereas in the Lower Zone mafic layers generally occur within plagioclase lherzolite (Fig. 2). In the Upper Zone, dunite layers (a few centimeters to meters thick) may occur between harzburgite and mafic layers, particularly for thick mafic layers (Fig. 4, B section).

PETROGRAPHY

Rock types

Niida, (1984) recognized two dominant types of mafic layers with `gabbroic' mineral assemblage, Gabbro I (GB I) and Gabbro II (GB II). Recently, Shiotani & Niida, (1997) added two more types of gabbroic mineral assemblage (GB III and GB IV), which are characterized by larger amounts of orthopyroxene than in GB I and GB II. In this study we focus on the dominant types, GB I and GB II. As demonstrated below, the present mineral assemblage of `gabbros' reflects subsolidus reactions rather than primary igneous assemblages. To avoid misunderstanding, we refer to GB I-IV of Niida, (1984) and Shiotani & Niida, (1997) as `mafic granulite Types I-IV'. We also report a new type of mafic granulite, designated as Type V, which has the mineral assemblage of gabbro norite. On the basis of the mineral assemblages and textures, the mafic layers that we studied are classified into two major groups with four rock types as follows:

Group A: mafic granulite

Type I: plagioclase (pl) + clinopyroxene (cpx) + olivine (ol) + orthopyroxene (opx) + Ti-pargasite (parg) or kaersutite + green spinel (sp) + titaniferous magnetite + ilmenite + sulfide

Type II: pl + Cr-rich cpx + ol + opx + parg + green sp + magnetite + sulfide

Type V: opx + green sp + pl + cpx + ol + phlogopite (phl)

Group B: pyroxenite

plagioclase websterite: cpx + opx + ol + pl + sp + parg ± phl.

The petrographic characteristics of each rock type are summarized below.

Type I and Type II mafic granulites

These are the most abundant types of mafic layers in the Horoman complex. Although they have similar mineral assemblages, the two types are clearly distinguished by their whole-rock compositions; e.g. Type I has whole-rock TiO2 > 0·3 wt % but Type II has TiO2 < 0·1 wt % (Shiotani & Niida, 1997; this study). The former contains titaniferous diopside or augite (TiO2 0·5-2·5 wt %), titaniferous pargasite or kaersutite, ilmenite and titaniferous magnetite (TiO2 3-5 wt %) (Niida, 1984; Shiotani & Niida, 1997). On the other hand, Type II contains chromian diopside (Cr2O3 ~1·5 wt %), low-Ti pargasite and Ti-free magnetite.

In the field, these two types can easily be distinguished because Type I layers are dark, purplish brown, whereas Type II layers are light greenish. Although mafic layers are more abundant in the Upper Zone than in the Lower Zone, the relative proportions of Type I and II layers are similar in both zones. It is noted that, in the Upper Zone, Type II layers are traceable for more than a kilometer, whereas Type I layers are highly discontinuous. A thick Type I layer in the Upper Zone is bordered by augen-shaped orthopyroxenites (few tens of centimeters thick) (Niida, 1984; Shiotani & Niida, 1997). In the Lower Zone a symplectite-bearing wehrlitic zone (~7 cm thick) occurs between a Type II layer and adjacent harzburgite (Morishita & Arai, 1997). These symplectites contain fine grains of spinel + plagioclase + orthopyroxene.

A typical modal composition of both Type I and Type II mafic granulites is 35-55% plagioclase, 10-45% clinopyroxene, 6-40% olivine, and minor amounts of orthopyroxene, pargasite, green spinel and opaque phases. Figure 6 shows some examples of intralayer modal variations for thick Type I and Type II layers. Two Type I layers (70 and 110 cm thick) have relatively homogeneous mineral proportions, whereas a Type II layer (100 cm thick) shows pronounced symmetrical intralayer modal variation; i.e. the central part is enriched in plagioclase and clinopyroxene, and depleted in olivine, orthopyroxene and magnetite. Both Type I and II layers show mosaic equigranular to gneissic texture indicating solid-state plastic deformation and grain boundary migration (Mercier & Nicolas, 1975; Nicolas & Poirier, 1976). Typically there are alternating clinopyroxene-rich and plagioclase- and olivine-rich domains. Small amounts of orthopyroxene and pargasite occur interstitially in the clinopyroxene-rich domains. In Type I layers, large deformed clinopyroxene grains (>1 mm) contain exsolution lamellae of titaniferous magnetite and ilmenite. Olivine crystals are associated with opaque minerals (magnetite, titaniferous magnetite and ilmenite) in the fine-grained plagioclase-rich domain. Spindle-shaped or anhedral granular grains (0·2-5·0 mm) of green spinel may also occur in the plagioclase matrix. A Type I layer from the Lower Zone shows particularly well-developed gneissic texture with anhedral grains of plagioclase, clinopyroxene and olivine (0·1-0·3 mm in size).


Figure 6. Modal proportions (vol. %) in three thick Type I and II layers from the Horoman peridotite. Modes were determined by point counting, 1600-3600 points, per thin section. plag, plagioclase; cpx, clinopyroxene; ol, olivine; opx, orthopyroxene; opq, opaque minerals; Ti-parg, titaniferous pargasite; mag, magnetite; parg, pargasite. Narrow regions without mineral names are for opx-mag-opq-parg. Columns at right show expanded detail, and location of studied samples, T-35, etc.


Type V mafic granulite

The Type V mafic granulite occurs in lenticular form, about 1 m in width, in harzburgite near the Banded Dunite Harzburgite suite of Takahashi, (1991). It consists of orthopyroxene (~60%), green spinel (~20%),plagioclase (~15%), clinopyroxene (~5%) and minor amounts of olivine and phlogopite. It is characterized by symplectites of spinel, orthopyroxene and plagioclase with rare clinopyroxene. The texture is tabular mosaic equigranular with a typical grain size of 0·1 mm. Tiny plates of phlogopite rarely occur at orthopyroxene grain boundaries. The Type V lenses are typically mantled by a sequence of websterite (a few centimeters thick), orthopyroxenite and several tens of centimeters ofwehrlite in contact with the host harzburgite.

Plagioclase websterite

The plagioclase websterites consist of clinopyroxene,orthopyroxene, olivine, plagioclase, and small amounts of pargasite ± phlogopite. The texture is transitional between porphyroclastic and equigranular. Abundant pyroxene porphyroclasts are associated with smaller amounts of neoblasts; apparently, recrystallization was less complete in plagioclase websterites than in the mafic granulites. Orthopyroxenes show a pleochroism of pale pink to pale green. Plagioclase, from a few modal percent up to 40 modal percent, occurs as small interstitial grains (~0·5 mm) in pyroxene-rich matrices and sometimes forms plagioclase-rich seams together with brown or green spinel grains. Some clinopyroxenes contain thin lamellae of brown spinel. Symplectites (0·8-1·5 mm) that consist of clinopyroxene, orthopyroxene, brown spinel ± phlogopite occur in some samples. Small amounts (<0·5%) of strongly pleochroic pargasite (pale greenish gray) and phlogopite (pale brown) occur sporadically and interstitially among the pyroxene grain boundaries.

ANALYTICAL PROCEDURE

The whole-rock compositions of 30 mafic layers (18 Type I, seven Type II, one Type V, four plagioclase websterites) were determined using X-ray fluorescence (XRF), instrumental neutron activation analysis (INAA) and inductively coupled plasma-mass spectrometry (ICP-MS).

Abundances of major elements, Cr and Ni ( Table 2) were determined by XRF at Hokkaido University following the analytical procedure of Tsuchiya et al., (1989), and were previously reported by Takazawa, (1989). Abundances of V, Cr, Ni, Zn, Sr and Zr in 29 samples (Table 2) were analyzed by XRF at the University of Massachusetts (Amherst) using the procedures of Rhodes, (1983). Abundances of Sc and Cr in seven whole rocks (Table 2) were determined by INAA at MIT following the procedures of Ila & Frey, (1984).


Table 2. Whole-rock major and trace element compositions of mafic rocks

Abundances of Rb, Sr, Zr, Nb, Ba, Hf, Th and 14 REE (Table 3) were measured in 15 samples by ICP-MS at Université de Montpellier, France, following the procedures described by Ionov et al., (1992). The results for a standard RO-A1 (pyroxenite) are given in Table 3 and are compared with those of Remaïdi, (1993).


Table 3. Whole-rock trace element abundances determined by ICP-MS

Sr and Nd isotopic compositions for 11 whole rocks (not acid-leached) were determined by thermal ionization-mass spectrometry at MIT. Subsequently, the powders for six of these samples were leached for 1 h with 6·2 N HCl at 100°C. Acid leaching lowered 87Sr/86Sr, but no significant change was observed for 143Nd/144Nd (Table 4). Chemical separation and mass spectrometry of Sr and Nd isotopes at WHOI followed the procedures of Hart & Brooks, (1977) and Zindler et al., (1979), whereas those at MIT followed the procedures of Pin & Bassin, (1992) and Bowring & Housh, (1995). MIT and WHOI measurements of the same rock powders (G-1c and FU-14) agree within stated uncertainties ( Table 4). Six samples (not acid-leached) were analyzed for Pb isotopic ratios (Table 5) at MIT after chemical separation of Pb at WHOI, following the procedures of Manhes et al., (1978). The mass spectrometry of Pb isotopes at MIT followed that of Housh & Bowring, (1991).


Table 4. Sr, Nd isotope ratios for the Horoman mafic layers


Table 5. Pb isotope ratios for mafic layers from the Horoman peridotite

RESULTS

Major elements and compatible trace elements

Type I layers range in mg-number [100 Mg/(Mg + Fe)] from 55 to 89 whereas Type II range only from 84 to 88 (Table 2 and Fig. 7). The centers of thick (>75 cm) Type I layers (Fig. 6) have lower mg-number than the margins and thinner Type I layers (Table 2). Also the mg-number of the centers decreases as the distance from mafic layer-peridotite contact increases (Fig. 8). Niida, (1984) showed that Fo content across one thick Type I layer displayed a parabolic pattern with lowest Fo contents in the center of the layer. These results are consistent with diffusion-controlled Fe-Mg exchange between Type I mafic layers and their host peridotite. In contrast, even thick Type II and plagioclase websterite layers do not have low mg-number centers (Fig. 8).


Figure 7. Major oxides (in wt %) vs mg-number [100Mg/(Mg + Fe*), where Fe* is total Fe]. Circles and squares indicate data for Horoman Type I and Type II layers, respectively (their fields are enclosed and labeled in the TiO2 panel). Shaded symbols indicate samples collected from the center of thick layers (75-135 cm); tie-lines connect data for centers and margins of thick layers (Fig. 6). The shaded field is for Type I garnet clinopyroxenites in the Beni Bousera peridotite (Kornprobst, 1969; Pearson et al., 1993; Kumar et al., 1996) and Ronda peridotite [Groups A and C of Suen & Frey, (1987)]. The slashed field is for Type II garnet clinopyroxenites in Beni Bousera peridotite (Kornprobst et al., 1990b; Pearson et al., 1993) and olivine gabbros in Ronda peridotite [Groups E and F of Suen & Frey, (1987)]. The cross symbol indicates garnet clinopyroxenite R-127 from the Ronda peridotite (Obata & Dickey, 1976; Suen & Frey, 1987).



Figure 8. The mg-number of mafic layers vs the distance between sample location and peridotite-mafic layer contact. Shaded symbols indicate samples collected from the center of thick layers (75-135 cm); tie-lines connect data for centers and margins of thick layers (Fig. 6). Shaded rectangle indicates mg-number of melt in equilibrium with mantle olivine (Fo 90-92) assuming an Fe-Mg exchange coefficient of 0·3-0·35.


The major element compositions of Type I layers are variable, with SiO2 ranging from 45 to 48 wt %, Al2O3 from 10 to 17 wt %, CaO from 8 to 15 wt % and Na2O from 0·9 to 1·6 wt %. Abundances of TiO2, Cr and Ni in Type I layers also vary over large ranges: TiO2 from 0·27 to 0·7 wt %, Cr from 400 to 3300 ppm and Ni from 100 to 900 ppm (Figs 7 and 9). The Cr and Ni contents of Type I layers decrease with decreasing mg-number (Fig. 9). Figures 7 and 9 also show the fields for Type I garnet clinopyroxenite layers from Beni Bousera peridotite (Kornprobst, 1969; Pearson et al., 1993; Kumar et al., 1996) and the Ronda peridotite (Suen & Frey, 1987). These garnet clinopyroxenites have compositions similar to the Type I Horoman layers.


Figure 9. Trace element abundances (in ppm) vs mg-number. Shaded symbols indicate samples collected from the center of thick layers (75-135 cm) (see Fig. 6). The shaded field is for Type I garnet clinopyroxenites in Beni Bousera peridotite (Pearson et al., 1993; Kumar et al., 1996) and Ronda peridotite [Groups A and C of Suen & Frey, (1987)]. The slashed field indicates Type II garnet clinopyroxenites in Beni Bousera peridotite (Kornprobst et al., 1990b; Pearson et al., 1993) and olivine gabbros in Ronda peridotite [Groups E and F of Suen & Frey, (1987)].


Type II layers are characterized by very low TiO2 content (Fig. 7). Also, they range to higher Al2O3 and CaO contents than Type I layers. Figures 7 and 9 also show the fields for olivine gabbros from the Ronda peridotite (Suen & Frey, 1987) and garnet clinopyroxenites (Type II) from Beni Bousera peridotite (Kornprobst et al., 1990b; Pearson et al., 1993). The Type II garnet clinopyroxenites of Beni Bousera peridotite and olivine gabbros of Ronda peridotite have compositions similar to the Type II layers. In contrast, the Type V layer has very low SiO2 (42 wt %) and CaO (5 wt %) but high Al2O3 content (20 wt %). Plagioclase websterites (G-12 and T-25) are similar in major element composition (Table 2); i.e. relatively high SiO2 (~52 wt %), and low Al2O3 (~5 wt %), and moderate CaO (~13 wt %) contents. The relatively high SiO2 and Cr (>6000 ppm) in the high mg-number websterites reflect the high modal abundance of both pyroxenes (Figs 7 and 9). Compared with these samples, plagioclase websterite (G-13) has lower SiO2 (~48 wt %), Cr (~1700 ppm), Ni (~370 ppm) contents and mg-number (~79), whereas plagioclase websterite (G-14) has lower SiO2 (~48 wt %) and CaO (~9 wt %), and higher Al2O3 (~13 wt %) contents.

Incompatible trace elements

For Type I layers, Zn shows inverse correlation with mg-number whereas Sr shows positive correlation with mg-number (Fig. 9). Abundances of Zn, Sc, V, Sr and Zr in the Type I layers are generally similar to those of the pyroxenite layers from Ronda and Beni Bousera. Compared with Type I layers, Type II layers trend to lower Zn, Sc, V and Zr, and higher Sr contents (Fig. 9).

Figure 10a-c shows primitive mantle normalized abundance patterns for incompatible elements in the Horoman mafic layers. Many of the Type I mafic layers are relatively depleted in Ti and highly incompatible elements (i.e. Rb to Ce) (Fig. 10a). Samples from the center and margin of three thick layers (70, 75 and 110 cm in thickness; see Fig. 6) were analyzed. In one thick layer the center (FU-12c) and margin (FU-13m) have nearly identical abundances of incompatible elements, and they have similar mg-number (Table 2; Fig. 8). On the other hand, the two other thick layers show large intralayer compositional variation. Their centers (G-1c and G-3c) are enriched in heavy REE (HREE) and highly depleted in light REE (LREE) relative to the margins (G-2m and G-4m), and their patterns cross each other at intermediate REEs, i.e. Gd-Tb (Fig. 10a). In both of these layers, the incompatible element depleted centers have much lower mg-number than the margins (Table 2; Fig. 8). Shiotani & Niida, (1997) also reported similar intralayer variations of REE abundances across a thick Type I Horoman layer (their GB I).


Figure 10. Primitive mantle normalized patterns for (a) Type I mafic layers, (b) Type II mafic layers, (c) other mafic layers from the Horoman peridotite and (d) margins of thick (70-135 cm) Type I layers (G-2m and G-4m), and the ultra-depleted melt inclusion in MORB olivine phenocrysts (Sobolev & Shimizu, 1993). Shaded field indicates the range for seamount lavas near the East Pacific Rise (Niu & Batiza, 1997). Primitive mantle values are from Sun & McDonough, (1989).


Type II mafic layers are characterized by relatively flat patterns with lower abundances of moderately incompatible elements (i.e. Eu to Lu) compared with the Type I layers (Fig. 10b). Marked features are relative enrichment of Sr and Eu and depletion of Zr and Hf relative to Sm. The Type V layer is relatively enriched in HREE, strongly depleted in LREE and other highly incompatible elements, and it is relatively depleted in Ti and Th (Fig. 10c). Plagioclase websterites have relatively flat REE patterns (Fig. 10c) but two samples (G-12 and G-13) are depleted in Zr and Hf relative to Sm.

Sr, Nd and Pb isotope compositions

Figure 11 shows present-day Sr and Nd isotopic ratios for mafic layers. Several Type I layers have 87Sr/86Sr and 143Nd/144Nd ratios equivalent to present-day MORB. The thick Type I layers have significant differences in 87Sr/86Sr and 143Nd/144Nd ratios between the centers (G-1c and G-3c) and margins (G-2m and G-4m). Although the centers differ in 87Sr/86Sr and 143Nd/144Nd ratios, the margins plot within the MORB field. Compared with MORB the three Type II layers have similar 87Sr/86Sr ratios, and two of the three have similar 143Nd/144Nd ratios. The Type V layer (FU-14) has 143Nd/144Nd ratio similar to MORB, but its 87Sr/86Sr ratio (unleached) is higher than the MORB range. A plagioclase websterite (G-14) has a 87Sr/86Sr ratio within the range of the MORB but its 143Nd/144Nd ratio is distinctly lower than the MORB values.


Figure 11. 143Nd/144Nd ratio vs 87Sr/86Sr ratio for whole-rock mafic layers from the Horoman peridotite. Shaded symbols indicate samples collected from the center of thick layers (75-135 cm). Analytical precision is smaller than the size of symbol. Tie-lines connect the center and margin of two thick Type I layers. Shaded field indicates the variations of 143Nd/144Nd and 87Sr/86Sr ratios of clinopyroxenes and whole rocks in the Horoman peridotite (Takazawa, 1996; Yoshikawa & Nakamura, 1999). Field outlined with continuous line indicates the isotopic range for MORB (Staudigel et al., 1984).


Pb isotope ratios for six unleached mafic layers are plotted in Fig. 12. The Type I and II layers and plagioclase websterite have Pb isotope ratios near the field for MORB and the lower end of the OIB field. The Type V layer (FU-14) has more radiogenic Pb isotopic ratios relative to other samples; e.g. 206Pb/204Pb = 21·19. In the 207Pb/204Pb vs 206Pb/204Pb diagram (Fig. 12a), FU-14 is on the Northern Hemisphere Reference Line (NHRL) (Hart, 1984), but it is significantly offset to lower 208Pb/204Pb (Fig. 12b).


Figure 12. Lead isotopic ratios for whole-rock mafic layers from the Horoman peridotite. Field outlined with continuous line indicates the MORB field whereas shaded field is for ocean island basalts. The line labeled NHRL indicates the Northern Hemisphere Reference Line of Hart, (1984). The locations of MORB and OIB fields, DMM, EMI, EMII, N-MORB and HIMU are from Hart, (1988).


DISCUSSION

Do the mafic layers represent melt compositions?

A popular hypothesis for the origin of mafic layers is that they formed by melt flow through conduits in the upper mantle. For this hypothesis, a critical question is whether mafic layers represent frozen melts, possibly primary magmas (e.g. Loubet & Allègre, 1982; Bodinier et al., 1987; Suen & Frey, 1987; Bodinier, 1988; McDonough & Frey, 1989; Pearson et al., 1993; Kumar et al., 1996). To answer this question, we first compare the Horoman mafic layers with melts produced by peridotite melting experiments at high pressures (Fig. 13); for example, the melts generated by melting peridotite KLB-1 (Hirose & Kushiro, 1993), whose composition is similar to fertile Horoman peridotite. The mafic layers have lower SiO2 and higher CaO contents than the experimental melts but some Type I mafic layers with 16 wt % MgO have compositions close to melts produced at 2·0-2·5 GPa. Within this pressure range, primary melts should be in equilibrium with residual spinel peridotite. Using the procedures of Kinzler & Grove, (1992), we find that the mafic layer compositions (Table 2) are not in equilibrium with spinel peridotite.


Figure 13. Comparison of whole-rock major element compositions of mafic layers from the Horoman peridotite with experimental melt compositions derived from peridotite KLB-1 at various pressures (Hirose & Kushiro, 1993). Shaded symbols for Horoman mafic layers indicate samples collected from the center of thick layers (75-135 cm); tie-lines connect data for centers and margins of thick layers. Experimental data at each pressure are connected by lines that show increasing degree of melting as MgO increases.


We also evaluated if the mafic layers lie on melt extraction lines defined by the peridotites (Suen & Frey, 1987). If the linear trends shown by the Horoman peridotites (Fig. 14) resulted from extraction of melt, these melts must plot on extensions of the linear peridotite trends. In fact, estimates for primary MORB compositions cluster along these trends (Fig. 14). Also, Type I mafic layers with 16 wt % MgO have Al2O3, FeO* and CaO contents on the extraction lines defined by the peridotites, but their relatively low SiO2 contents do not lie on this extraction line (Fig. 14). We conclude that (1) the mafic layers are not crystallized primary melts and (2) they do not represent melts extracted from the peridotites.


Figure 14. Oxides (in wt %) vs MgO diagrams for whole-rock mafic layers and peridotites from the Horoman peridotite. Straight lines are linear regression lines for peridotites. Shaded symbols for Horoman mafic layers indicate samples collected from the center of thick layers (75-135 cm); tie-lines connect data for centers and margins of thick layers. Smaller open circles with dot indicate the Horoman peridotites from Takazawa et al., (1999). Plus symbols indicate proposed primary ocean floor basalts (Basaltic Volcanism Study Project, 1981). Shaded bar indicates compositional range of melts in equilibrium with residual peridotite with olivine Fo content from 89 to 92. Vertical dashed line indicates 16 wt % MgO.


In addition, the Horoman mafic layers have lower incompatible element abundances than most erupted magmas, although melts with even lower incompatible element contents occur as melt inclusions in olivine phenocrysts (Fig. 10d). The low incompatible element content of the mafic layers indicates that they are not related to the incompatible element enriched fluid or melt that reacted with the Horoman harzburgites and lherzolites (Takazawa et al., 1992, , 1996). Depletion in incompatible elements is typical of mafic layers in orogenic lherzolites. This result led to the conclusion that these mafic layers do not represent melts; rather they are crystal segregates from a cooling melt or residues formed by partial melting of pre-existing mafic layers (e.g. Loubet & Allègre, 1982; Suen & Frey, 1987).

Effects of interaction between host peridotite and mafic layers

We have already noted that the mg-number and olivine composition in thin mafic layers and the margins of thick layers equilibrated by diffusive Fe-Mg exchange with the surrounding peridotites. Also, the orthopyroxenite augen on the borders of some Type I layers probably reflect melt-rock reactions. Such reactions have been documented for the mantle section of ophiolites (e.g. Varfalvy et al., 1996). To avoid the effects of interactions between mafic layers and their host peridotites, Pearson et al., (1993) suggested that the mafic layers must be sampled >10 cm away from the contact. Most of the samples that we studied are from layers >25 cm thick (Table 2).

A cumulate model for the Horoman mafic layers

Evaluation of a cumulate model requires identification of primary igneous phases. The igneous mineralogy of the Horoman mafic layers, however, has been modified by subsolidus reactions. Nevertheless, the primary igneous mineral assemblages of metamorphosed cumulates can be inferred from their whole-rock compositions. Unlike crystallized melts, the whole-rock compositions ofcumulates are controlled by their cumulus phases; for example, garnet-rich cumulates have high Al2O3 contents, high ratios of HREE/LREE and low Sr contents, whereas plagioclase-rich cumulates have high Al2O3 contents accompanied by low REE contents and relative enrichments in Sr and Eu (e.g. Bodinier et al., 1987).

If these layers formed as crystal segregates, it is necessary to evaluate the amount of trapped melt. A first-order estimation of the amount of trapped melt was obtained using abundances of a highly incompatible element, such as Th, and assuming a melt Th content of 0·12 ppm, equal to that in average N-MORB (Sun & McDonough, 1989). With these assumptions, the trapped melt component was highly variable but generally <10% (Table 6); specifically, 0-14% in Type I layers with the largest amounts in the high mg-number margins, 2-22% in Type II layers, and up to 33% in two of the three plagioclase websterites. The strongest geochemical evidence for a cumulate origin arises from abundances of compatible elements; these amounts of trapped melt cannot mask the important geochemical characteristics of cumulate rocks.

Type I mafic granulite

Type I layers have 40-50 vol. % modal plagioclase, but Eu anomalies are absent in their primitive mantle normalized patterns and most samples have near-chondritic Sr/Nd ratios (Fig. 10a). Thus there is no geochemical evidence for cumulus plagioclase that causes enrichment in Eu and Sr; therefore, it is likely that the plagioclase in these layers formed by subsolidus reaction. This interpretation is consistent with the metamorphic texture of these rocks and previous arguments that plagioclase in the plagioclase lherzolites formed during subsolidus decompression reactions (Takahashi & Arai, 1989; Ozawa & Takahashi, 1995; Takazawa et al., 1996).

Despite compositional heterogeneity in some thick Type I layers, all Type I layers have major and trace element compositions very similar to garnet clinopyroxenites in other orogenic peridotites (Figs 7 and 9). Figure 15 plots the whole-rock compositions of Type I layers projected from clinopyroxene onto the olivine-Ca-Tschermak pyroxene (CaTs)-quartz plane (Suen & Frey, 1987) and from CaTs onto the clinopyroxene-garnet-quartz plane in oxygen units. It should be noted that the thick Type I layers from the Horoman peridotites plot between the garnet and clinopyroxene apices. The similarity of Type I layers to garnet clinopyroxenites (Type I) from the Beni Bousera peridotite (Kornprobst, 1969; Pearson et al., 1993; Kumar et al., 1996) and the Ronda peridotite (Suen & Frey, 1987), is also shown in Fig. 15. Figure 16 shows garnet and clinopyroxene compositions determined in basalt melting experiments at pressures ranging from 1·6 to 3·2 GPa and various temperatures (Yasuda et al., 1994; Rapp & Watson, 1995; Takahashi et al., 1998; Tsuruta & Takahashi, 1998). Because the Type I layers plot between the fields of garnet and clinopyroxene, we infer that the Horoman Type I layers formed as garnet clinopyroxenites. Consistent with this inference, at 2·5 GPa, between 1400°C and 1375°C, the equilibrium assemblage for a plagioclase-garnet clinopyroxenite from the Ronda peridotite, sample R127, is garnet, clinopyroxene and a liquid (Obata & Dickey, 1976). The major element composition of R127 is similar to some of the Type I layers in the Horoman peridotite although R127 is slightly higher in TiO2 and Na2O (Fig. 7).


Figure 15. Whole-rock normative compositions of mafic layers projected from clinopyroxene (Cpx: diopside-hedenbergite) onto olivine (forsterite-fayalite)-Ca-Tschermak pyroxene (CaTs: CaAl2SiO6)-quartz plane (upper panel) and from CaTs onto Cpx-(Py-Alm) (pyrope-almandine)-quartz plane (lower panel) in oxygen units. Shaded field is for Type I garnet clinopyroxenites in Beni Bousera peridotite (Kornprobst, 1969; Pearson et al., 1993; Kumar et al., 1996) and Ronda peridotite [Groups A and C of Suen & Frey, (1987)]. The Horoman peridotite data are from Takazawa et al., (1999). It should be noted that none of mafic layers lies on the extension of the regression line. The CaTs-Opx join is the trace of the garnet-pyroxene plane, which is a thermal divide at pressure >3 GPa (O'Hara & Yoder, 1967). Inset shows tetrahedral projection of a quaternary system Olivine-CaTs-Cpx-Quartz in oxygen units. An, anorthite; Sp, spinel; Gr, grossular; Mg-Ts, Mg-Tschermak pyroxene. Parentheses indicate projected endmembers.



Figure 16. Normative compositions of clinopyroxene, garnet and coexisting melt projected from Cpx onto the olivine-CaTs-quartz plane (upper panel) and from CaTs onto (Py-Alm)-Cpx-quartz plane (lower panel) in oxygen units. The data are compiled from anhydrous and hydrous basaltic melting experiments at pressures ranging between 1·6 and 3·2 GPa and various temperatures (Yasuda et al., 1994; Rapp & Watson, 1995; Takahashi et al., 1998; Tsuruta & Takahashi, 1998).


The intralayer compositional heterogeneity of some Type I layers provides important clues to the petrogenesis of these layers. As emphasized above, the centers of some Type I layers have lower mg-number, lower abundances of highly incompatible elements and higher abundances of HREE than their margins; therefore their primitive mantle normalized trace element patterns cross in the vicinity of the middle REEs (Fig. 10a). The ratio of incompatible element abundances in a center relative to its margin closely mimics the bulk-solid-melt partition coefficient for a cumulate with 50% clinopyroxene and 50% garnet (Fig. 17). A simple model consistent with these results is that the LREE-depleted centers represent garnet clinopyroxenite `cumulates' whereas the margins represent mostly a melt that equilibrated with garnet and clinopyroxene. This melt is highly depleted in highly incompatible elements (Fig. 10d). Shiotani & Niida, (1997) proposed that the margin of the Type I layers represents a crystallized primary N-MORB melt along the wall of melt conduit. In the context of a cumulate relationship between the centers and their margins, the absence of systematic differences in major element composition (e.g. Table 2 and mantle norms in Fig. 15) suggests a eutectic composition. O'Hara & Yoder, (1967) demonstrated that in the CMAS system the eutectic point crosses the garnet-pyroxene plane around 3 GPa. Under such conditions, major element abundances are not changed by fractionation of a clinopyroxene and garnet assemblage.


Figure 17. Ratio of incompatible element abundances in the center relative to associated margin of two thick Type I layers (75-135 cm) compared with bulk-solid-melt partition coefficients for a cumulate with 50% garnet and 50% clinopyroxene. Partition coefficients from Johnson, (1998). The mismatch at Sr and Ti may indicate that the centers also contained a small amount (<5%) of cumulate plagioclase and a Ti-bearing phase coexisting with garnet and clinopyroxene.


Type II mafic granulite

The whole-rock trace element abundances establish the importance (>50%) of cumulus plagioclase in Type II mafic layers; specifically, the positive Sr and Eu anomalies, low REE abundances (<1 * chondrite) and flat REE patterns (Fig. 10b). Even the relative depletions of Zr and Hf in the normalized trace element patterns (Fig. 10b) are consistent with a plagioclase-dominant cumulate because plagioclase has low Zr/Sr, Zr/Sm, Hf/Sr and Hf/Sm relative to an equilibrium melt (Fujimaki et al., 1984). Moreover, in the olivine-CaTs-quartz plane and in the clinopyroxene-garnet-quartz plane, Type II layers plot in a triangle connecting olivine, anorthite and clinopyroxene (Fig. 18). Thus, although Type II layers have extensively recrystallized textures, we infer that their primary phases were very similar to the current mineral assemblage. The inferred larger amounts of trapped melt in samples G-6c and G-7m (Table 6) are consistent with their higher abundances of incompatible elements and relatively small Sr and Eu anomalies (Fig. 10b). On the basis of major and trace element compositions, Shiotani & Niida, (1997) also considered that Type II layers formed as cumulates consisting of olivine, plagioclase and clinopyroxene.


Table 6. Estimated amount of trapped melt


Figure 18. Whole-rock normative compositions of Type II layers and plagioclase websterites projected from Cpx onto olivine-CaTs-quartz plane (upper panel) and from CaTs onto (Py-Alm)-Cpx-quartz plane (lower panel) in oxygen units. Liquidus topologies from Suen & Frey, (1987) for pressures ranging between 1·9 and 2·5 GPa. Shaded field is for Type II garnet clinopyroxenites in Beni Bousera peridotite (Kornprobst et al., 1990b; Pearson et al., 1993) and for olivine gabbros in Ronda peridotite [Groups E and F of Suen & Frey, (1987)].


Type V mafic granulite

The Type V sample (FU-14) with a gabbro norite mineralogy has a major element composition similar to garnet: i.e. SiO2 ~42 wt %, Al2O3 ~20 wt %, CaO ~5 wt % and Cr ~12 000 ppm. It is also compositionally similar to garnet produced in high-pressure experiments (Fig. 16). Abundances of trace elements strongly support the interpretation that this lens formed as garnet; i.e. it is highly depleted in LREE, enriched in HREE, relatively depleted in Ti and enriched in Zr and Hf, and has a very high Sc content (131 ppm) (Figs 9 and 10c). If we assume that FU-14 formed as garnet in equilibrium with melt and use garnet-melt partition coefficients (e.g. Hauri et al., 1994; Johnson, 1998), the melt is highly enriched in LREE, as expected for small degrees of melting of garnet peridotite.

Plagioclase websterite

The textures of plagioclase websterite also indicate subsolidus metamorphic reactions. For example, plagioclase often occurs in seams with clinopyroxene, orthopyroxene and spinel. Two pyroxene-spinel symplectites in plagioclase websterite (G-14) may have formed by breakdown of garnet. In the olivine-CaTs-quartz plane and in the clinopyroxene-garnet-quartz plane the plagioclase websterite (G-14) contains a larger amount of garnet component than other websterites (Fig. 18). The higher HREE abundances and the relative depletion of Ti in this sample are also consistent with a cumulate garnet hypothesis for this sample (Fig. 10c). Similar to the Ronda websterites (Suen & Frey, 1987), websterites in the Horoman peridotite may be explained by precipitation of crystal segregates (clinopyroxene + orthopyroxene + olivine + spinel) from high-MgO melts at pressures between 1·9 and 2·5 GPa (Fig. 18). On the other hand, the plagioclase websterite (G-14) originally contained garnet, rather than spinel, and it formed at higher pressure.

Ages and source regions of the mafic layers in the Horoman peridotite

Thin Type I layers and the margins of thick Type I layers have present-day Sr and Nd isotopic ratios similar to MORB (Table 4, Fig. 11). In contrast, the centers of the heterogeneous layers, G-1c and G-3c, are offset from the MORB field to higher 143Nd/144Nd (Fig. 11). An explanation for the very high 143Nd/144Nd in the centers of thick Type I layers is aging, resulting from their very high Sm/Nd (Fig. 10a). For example, two-point Nd isotope isochrons for the margins and centers yield ages of 81·7 ± 13·1 Ma and 81·3 ± 18·0 Ma for G-1c and G-3c, respectively (Fig. 19). In contrast, the 87Sr/86Sr data do not yield consistent ages. Given the low Sr contents in the centers (4-9 ppm; Table 3), the disparate center to margin trends for 87Sr/86Sr (Fig. 11) and the significant effects of acid leaching on 87Sr/86Sr (Table 4), we conclude that the Sr isotopic system was recently disturbed.


Figure 19. 143Nd/144Nd vs 147Sm/144Nd ratio diagram for mafic layers from the Horoman peridotite. Shaded symbols indicate samples collected from the center of thick layers (75-135 cm); tie-lines connect data for centers and margins of thick layers. The 831 Ma reference isochron is based on data for whole-rock peridotites (Yoshikawa & Nakamura, 1999). CHUR is the chondritic uniform reservoir. Shaded field indicates the depleted MORB source mantle.


Based on the Nd isotopic system, we infer that at ~80 Ma the centers of these layers formed as garnet pyroxenites segregated from magmas that were depleted in highly incompatible elements. The geochemical similarities of the Type I layers to aged MORB (Fig. 19) show that their magma source was not enriched mantle formed by deep recycling of crust or in a mantle wedge overlying a subduction zone. Additional evidence for MORB-type magmatism in this region is that Maeda & Kagami, (1996) reported MORB-like Sr and Nd isotope ratios from some gabbroic and dioritic intrusives and basaltic dikes in the Main Zone of the Hidaka metamorphic belt.

A thick Type II layer (G-6 and G-7) has the MORB-like Sr and Nd isotopic signature whereas another thick Type II layer (FU-11) has a higher 143Nd/144Nd ratio of 0·513731 (Fig. 11). It is important that in the Sm-Nd isochron diagram these Type II layers lie close to the whole-rock isochron defined by Horoman peridotites (Fig. 19). This whole-rock isochron defines an age of 831 Ma and an initial 143Nd/144Nd ratio of 0·5119 that is equivalent to MORB source mantle at this time. Yoshikawa & Nakamura, (1999) concluded that partial melting of the peridotite occurred at this time. We infer that Type II layers formed soon after this partial melting event.

The Type V layer (FU-14) has highly radiogenic Pb, 206Pb/204Pb = 21·194, 207Pb/204Pb = 15·781 and 208Pb/204Pb = 39·638; these values are close to the HIMU component recognized in some oceanic basalts (Zindler & Hart, 1986) (Fig. 12). On the basis of trace element abundances and a very high Al2O3/CaO ratio (e.g. Figs 7, 9 and 10c), we infer that this layer formed as a garnet-dominant cumulate. Therefore, these radiogenic Pb ratios may reflect aging of garnet with high U/Pb and Th/Pb ratios (Hauri et al., 1994). Also, experimental results for partitioning of U and Th between garnet and basaltic melt demonstrate that garnet preferentially incorporates U relative to Th; e.g. Beattie, (1993), LaTourrette et al., (1993) and Hauri et al., (1994) found that Dgarnet/melt(U/Th) ranges from 2 to 13. Thus, the high U/Th ratio expected in garnet is consistent with a lower 208Pb/204Pb ratio for FU-14 relative to the NHRL (Fig. 12). The Nd and Pb systematics of this sample, however, do not yield a consistent age inference.

Tectonic implications based on the Horoman mafic layers

The age constraints for the Horoman peridotite range from the whole-rock peridotite Sm-Nd isochron interpreted as reflecting partial melting at 831 Ma (Yoshikawa & Nakamura, 1999) to a 23 Ma Rb-Sr mineral isochron for a phlogopite-bearing peridotite interpreted as reflecting cooling after metasomatism associated with uplift of the Horoman peridotite from the mantle to the crust (Yoshikawa et al., 1993). There is considerable evidence that this ascent originated within the stability field of garnet peridotite (Ozawa & Takahashi, 1995; fig. 8 of Takazawa et al., 1996). The ages and primary mineralogy that we infer for the mafic layers provide further constraints on the tectonic history of the complex. Two important features of Type II mafic layers are: (1) they are relatively old; apparently they formed shortly after the host peridotite experienced partial melting at ~830 Ma; (2) they formed as olivine gabbro cumulates within the plagioclase stability field. This evidence for ancient low-pressure events in a peridotite that was recently emplaced into the crust from the garnet stability field requires a complex pressure-temperature history for the Horoman complex. That is, the Type II mafic layers formed within the plagioclase stability field were subsequently subjected to considerably higher pressure. Evidence for this high-pressure history is (1) the wehrlitic margin of a Type II layer, which contains symplectites that may have formed by breakdown of garnet (Morishita & Arai, 1997), and (2) the recent finding of corundum-bearing gabbroic rocks of Type II lithology in a boulder that was probably derived from a Type II layer (Morishita & Kodera, 1998). The corundum is surrounded by spinel [Cr/(Cr+Al) ~0] and plagioclase (An > 96). Morishita & Kodera, (1998) concluded that the corundum formed when a plagioclase-rich troctolite was metamorphosed at high pressure.

During uplift, garnet in the peridotite and mafic layers reacted to form a lower-pressure mineral assemblage. In the peridotites the cores of clinopyroxene porphyroclasts retain compositions that indicate equilibration with garnet (Takazawa et al., 1996). Also, fertile peridotites typically contain two-pyroxene-spinel symplectites, which are interpreted to be the breakdown products of garnet (Takahashi & Arai, 1989). In both Type I and II mafic layers, however, there is only rare evidence for former garnet. The complete breakdown of garnet may have been facilitated by H2O that now resides in up to 4% modal pargasite. This H2O may have been introduced during metasomatism (Yoshikawa et al., 1993).

MAJOR CONCLUSIONS

(1) Similar to other orogenic lherzolites and mantle xenoliths, the Horoman peridotite contains two dominant types of mafic layers (Type I and Type II mafic granulites). Type I layers have higher Ti and lower Cr contents than Type II layers. Both types consist of plagioclase, clinopyroxene and olivine associated with minor pargasite and opaque minerals. However, these assemblages reflect subsolidus reactions occurring during uplift of the Horoman complex from upper mantle to the lower crust.

(2) The thick Type I layers have significant intralayer compositional heterogeneity. Relative to the margins, the centers have lower mg-number, lower abundances of highly incompatible elements, higher abundances of HREE and higher 143Nd/144Nd ratios. We propose that the centers represent garnet clinopyroxenite cumulates and the margins represent melt in equilibrium with these cumulates. These melts had the isotope characteristics of MORB. On the basis of two-point isochrons and model ages for the Sm-Nd isotopic system, the thick Type I layers formed at ~80 Ma.

(3) Relative to Type I layers, the Type II layers are characterized by lower abundances of incompatible elements and relative enrichments of Sr and Eu in primitive mantle normalized patterns. Type II layers formed as plagioclase-rich cumulates in the upper mantle. Although they formed at lower pressures than Type I layers, the Type II layers have Nd isotopic systematics consistent with an 831 Ma whole-rock peridotite isochron, which may indicate the age of partial melting of the Horoman peridotite (Yoshikawa & Nakamura, 1999).

This evidence for ancient low-pressure events in a peridotite that was recently emplaced into the crust from the garnet stability field (Ozawa & Takahashi, 1995; Takazawa et al., 1996) constrains the pressure-temperature history for the Horoman complex. A possible scenario is (a) ascent of MORB source mantle and partial melting at pressures near the transition from garnet peridotite to spinel peridotite (Takazawa et al., 1999; Yoshikawa & Nakamura, 1999); (b) crystallization of Type II layers in the shallow oceanic upper mantle (this study); (c) subsidence of the complex to pressures within the stability field of garnet peridotite (Morishita & Kodera, 1998; Morishita, 1999; this study); (d) introduction of a second generation of MORB-related melts at ~80 Ma, which led to formation of the Type I layers as garnet clinopyroxenites (this study); (e) recent (~23 Ma, Yoshikawa et al., 1993) uplift from the garnet peridotite stability field into the crust (Ozawa & Takahashi, 1995; Takazawa et al., 1996).

ACKNOWLEDGEMENTS

The manuscript benefited from constructive reviews by Réjean Hébert and Kazuhito Ozawa. We thank Pamela D. Kempton for her editorial advice and patience. E.T. is grateful to Tomoaki Morishita and Kiyoaki Niida for invaluable discussions on the Horoman peridotite and mafic layers. The authors thank Jean-Louis Bodinier, Liliane Savoyant and Simone Pourtales for technical guidance and assistance in ICP-MS analyses at Université de Montpellier; Sam Bowring and Drew Coleman for isotopic analyses at MIT; Stan Hart and Jurek Blusztajn for isotopic analyses at WHOI; and Mike Rhodes and Pete Dawson for XRF analyses at University of Massachusetts (Amherst). Most of this research effort was supported by US NSF Grants EAR 9406105 and 9406914.

REFERENCES

Allègre, C. J. & Turcotte, D. L. (1986). Implications of a two-component marble-cake mantle. Nature 323, 123-127.

Basaltic Volcanism Study Project (1981). Basaltic Volcanism on the Terrestrial Planets. New York: Pergamon Press, 1286 pp.

Beattie, P. (1993). Uranium-thorium disequilibria and partitioning on melting of garnet peridotite. Nature 363, 63-65.

Bodinier, J.-L. (1988). Geochemistry and petrogenesis of the Lanzo peridotite body, western Alps. Tectonophysics 149, 67-88.

Bodinier, J.-L., Guiraud, M., Fabriès, J., Dostal, J. & Dupuy, C. (1987). Petrogenesis of layered pyroxenites from the Lherz, Freychinède and Prades ultramafic bodies (Ariège, French Pyrenees). Geochimica et Cosmochimica Acta 51, 279-290.

Bowring, S. A. & Housh, T. (1995). The earth's early evolution. Science 269, 1535-1540.

Fujimaki, H., Tatsumoto, M. & Aoki, K. (1984). Partition coefficients of Hf, Zr, and REE between phenocrysts and groundmasses. Proceedings of the 14th Lunar Planetary Science Conference, Part 2. Journal of Geophysical Research 89, Supplement, B662-B672.

Garrido, C. J. & Bodinier, J. L. (1999). Diversity of mafic rocks in the Ronda peridotite: evidence for pervasive melt-rock reaction during heating of subcontinental lithosphere by upwelling asthenosphere. Journal of Petrology 40, 729-754.

Hart, S. R. (1984). A large-scale isotope anomaly in the Southern Hemisphere mantle. Nature 309, 753-757.

Hart, S. R. (1988). Heterogeneous mantle domains: signatures, genesis and mixing chronologies. Earth and Planetary Science Letters 90, 273-296.

Hart, S. R. & Brook, C. (1977). Geochemistry and evolution of early Precambrian mantle. Contributions to Mineralogy and Petrology 61, 109-128.

Hauri, E. H., Wagner, T. P. & Grove, T. L. (1994). Experimental and natural partitioning of Th, U, Pb and other trace elements between garnet, clinopyroxene and basaltic melts. Chemical Geology 117, 149-166.

Hirose, K. & Kushiro, I. (1993). Partial melting of dry peridotites at high pressures: determination of compositions of melts segregated from peridotite using aggregates of diamond. Earth and Planetary Science Letters 114, 477-489.

Hirschmann, M. M. & Stolper, E. M. (1996). A possible role of garnet pyroxenite in the origin of the `garnet signature' in MORB. Contributions to Mineralogy and Petrology 124, 185-208.

Housh, T. & Bowring, S. A. (1991). Lead isotopic heterogeneities within alkali feldspars: implications for the determination of initial lead isotopic compositions. Geochimica et Cosmochimica Acta 55, 2309-2316.

Ila, P. & Frey, F. A. (1984). Utilization of neutron activation analysis in the study of geological materials. Atomkernernergie Kerntechnik 44, 710-716.

Ionov, D. A., Savoyant, L. & Dupuy, C. (1992). Application of the ICP-MS technique to trace element analysis of peridotites and their minerals. Geostandards Newsletter 16, 311-315.

Johnson, K. T. M. (1998). Experimental determination of partition coefficients for rare earth and high-field-strength elements between clinopyroxene, garnet, and basaltic melt at high pressures. Contributions to Mineralogy and Petrology 133, 60-68.

Kinzler, R. J. & Grove, T. L. (1992). Primary magmas of mid-ocean ridge basalts, 1. Journal of Geophysical Research 97, 6885-6906.

Komatsu, M. & Nochi, M. (1966). Ultrabasic rocks in Hidaka metamorphic belt, Hokkaido, Japan. I. Mode of occurrence of the Horoman ultrabasic rocks. Earth Sciences 20, 99-108 (in Japanese).

Komatsu, M., Kimura, G. & Kiminami, K. (1981). Tectonics of Hokkaido, with special reference to the Hidaka Metamorphic Belt. In: Hara, I. (ed.) Tectonics of Paired Metamorphic Belts. Hiroshima: Tanishi, pp. 55-59.

Komatsu, M., Miyashita, S., Maeda, J., Osanai, Y., Toyoshima, T., Motoyoshi, Y. & Arita, K. (1982). Petrological constitution of the continental type crust upthrust in the Hidaka belt, Hokkaido. Journal of Japanese Association of Mineralogy, Petrology and Economical Geology, Special Paper 3, 200-230 (in Japanese with English abstract).

Komatsu, M., Miyashita, S., Maeda, J., Osanai, Y. & Toyoshima, T. (1983). Disclosing of a deepest section of continental-type crust up-thrust as the final event of collision of arcs in Hokkaido, North Japan. In: Hashimoto, M. & Uyeda, S. (eds) Accretion Tectonics in the Circum-Pacific Regions. Tokyo: Terra, pp. 149-165.

Komatsu, M., Miyashita, S. & Arita, K. (1986). Composition and structure of the Hidaka metamorphic belt, Hokkaido: historical review and present status. Geology and Tectonics of Hokkaido, Monograph of the Association for the Geological Collaboration in Japan 31, 189-203 (in Japanese with English abstract).

Komatsu, M., Osanai, Y., Toyoshima, T. & Miyashita, S. (1989). Evolution of the Hidaka metamorphic belt, northern Japan. In: Daly, J. S., Cliff, R. A. & Yardley, B. W. D. (eds) Evolution of Metamorphic Belts. Geological Society, London, Special Publication 43, 487-493.

Kornprobst, J. (1969). Le massif ultrabasique de Beni-Bouchera (Rif interne, Maroc): étude des péridotites de haute temperature et de haute pression et des pyroxénites à grenat qui leur sont associées. Contributions to Mineralogy and Petrology 23, 283-332.

Kornprobst, J., Piboule, M., Roden, M. & Tabit, A. (1990a). Corundum-bearing garnet clinopyroxenites at Beni Bousera (Morocco): original plagioclase-rich gabbros recrystallized at depth within the mantle? Journal of Petrology 31, 717-745.

Kornprobst, J., Tabit, A. & Piboule, M. (1990b). Corundum bearing garnet pyroxenites at Beni-Bousera (Rif, Morocco): analytical data on rocks and minerals. Université Blaise Pascal, Note OPGC 106, 1-49.

Kumar, N., Reisberg, L. & Zindler, A. (1996). A major and trace element and strontium, neodymium, and osmium isotopic study of a thick pyroxenite layer from the Beni Bousera Ultramafic Complex of northern Morocco. Geochimica et Cosmochimica Acta 60, 1429-1444.

LaTourrette, T. Z., Kennedy, A. K. & Wasserburg, G. J. (1993). Thorium-uranium fractionation by garnet: evidence for a deep source and rapid rise of oceanic basalts. Science 261, 739-742.

Loubet, M. & Allègre, C. J. (1982). Trace elements in orogenic lherzolites reveal the complex history of the upper mantle. Nature 298, 809-814.

Maeda, J. & Kagami, H. (1996). Interaction of a spreading ridge and an accretionary prism: implications from MORB magmatism in the Hidaka magmatic zone, Hokkaido, Japan. Geology 24, 31-34.

Maeda, J., Suetake, S., Ikeda, Y., Tomura, S., Motoyoshi, Y. & Okamoto, Y. (1986). Tertiary plutonic rocks in the axial zone of Hokkaido-distribution, age, major element chemistry, and tectonics. Geology and Tectonics of Hokkaido, Monograph of the Association for the Geological Collaboration in Japan 31, 223-246 (in Japanese with English abstract).

Manhes, G., Minster, J.-F. & Allègre, C.-J. (1978). Comparative uranium-thorium-lead and rubidium-strontium study of the St. Severin amphoterite: consequences for early solar system chronology. Earth and Planetary Science Letters 39, 14-24.

McDonough, W. F. & Frey, F. A. (1989). Rare earth elements in upper mantle rocks. In: Lipin, B. R. & McKay, G. A. (eds) Geochemistry and Mineralogy of Rare Earth Elements. Mineralogical Society of America, Reviews in Mineralogy 21, 99-145.

Mercier, J.-C. C. & Nicolas, A. (1975). Textures and fabrics of upper mantle peridotites as illustrated by xenoliths from basalts. Journal of Petrology 16, 454-498.

Miyashita, S. (1983). Reconstruction of the ophiolite succession in the western zone of the Hidaka Metamorphic Belt, Hokkaido. Journal of Geological Society of Japan 89, 69-86 (in Japanese with English abstract).

Morishita, T. (1999). Evolution of the Horoman Peridotite Complex and its implications for the origin of heterogeneous mantle. Ph.D. Dissertation, Kanazawa University, 182 pp.

Morishita, T. & Arai, S. (1997). Diversity of occurrence of symplectite in the Horoman peridotite complex of the Hidaka belt, Hokkaido, northern Japan, and its bearing on the P-T history. Memoirs of the Geological Society of Japan 47, 149-162 (in Japanese with English abstract).

Morishita, T. & Kodera, T. (1998). Finding of corundum-bearing boulder possibly derived from the Horoman Peridotite Complex, Hokkaido, northern Japan. Journal of Mineralogy, Petrology and Economic Geology 93, 52-63.

Nicolas, A. & Jackson, M. (1982). High temperature dikes in peridotites: origin by hydraulic fracturing. Journal of Petrology 23, 568-582.

Nicolas, A. & Poirier, J. P. (1976). Crystalline Plasticity and Solid State Flow in Metamorphic Rocks. New York: Wiley-Interscience, 444 pp.

Niida, K. (1974). Structure of the Horoman ultramafic massif of the Hidaka metamorphic belt in Hokkaido, Japan. Journal of the Geological Society of Japan 80, 31-44.

Niida, K. (1975). Textures and olivine fabrics of the Horoman ultramafic rocks, Japan. Journal of Japanese Association of Mineralogy, Petrology and Economical Geology 70, 265-285.

Niida, K. (1984). Petrology of the Horoman ultramafic rocks in the Hidaka metamorphic belt, Hokkaido, Japan. Journal of Faculty of Sciences, Hokkaido University, Series IV 21, 197-250.

Niu, Y. & Batiza, R. (1997). Trace element evidence from seamounts for recycled oceanic crust in the Eastern Pacific mantle. Earth and Planetary Science Letters 148, 471-483.

Nixon, P. H. (1987). Mantle Xenoliths. New York: John Wiley, 844 pp.

Obata, M. (1980). The Ronda peridotite: garnet-, spinel-, and plagioclase-lherzolite facies and the P-T trajectories of a high-temperature mantle intrusion. Journal of Petrology 21, 533-572.

Obata, M. & Dickey, J. S. J. (1976). Phase relations of mafic layers in the Ronda peridotite. Carnegie Institution of Washington Yearbook 75, 562-566.

Obata, M. & Nagahara, N. (1987). Layering of alpine-type peridotite and the segregation of partial melt in the upper mantle. Journal of Geophysical Research 92, 3467-3474.

O'Hara, M. J. & Yoder, H. S. (1967). Formation and fractionation of basic magmas at high pressures. Scottish Journal of Geology 3, 67-117.

Osanai, Y. (1985). Geology and metamorphic zoning of the Main Zone of the Hidaka Metamorphic Belt in the Shizunai River region, Hokkaido. Journal of the Geological Society of Japan 91, 259-278.

Osanai, Y., Komatsu, M. & Owada, M. (1991). Metamorphism and granite genesis in the Hidaka metamorphic belt, Hokkaido, Japan. Journal of Metamorphic Geology 9, 111-124.

Ozawa, K. & Takahashi, N. (1995). P-T history of a mantle diapir: the Horoman peridotite complex, Hokkaido, northern Japan. Contributions to Mineralogy and Petrology 120, 223-248.

Owada, M. (1989). Geology and chemical composition of granitic rocks in the southern part of the Hidaka metamorphic belt, with special reference to cordierite-bearing granitic rocks. Journal of the Geological Society of Japan 95, 227-240.

Pearson, D. G., Davies, G. R. & Nixon, P. H. (1993). Geochemical constraints on the petrogenesis of diamond facies pyroxenites from the Beni Bousera peridotite massif, North Morocco. Journal of Petrology 34, 125-172.

Pin, C. & Bassin, C. (1992). Evaluation of a strontium-specific extraction chromatographic method for isotopic analysis in geological materials. Analytica Chimica Acta 269, 249-255.

Polvé, M. & Allègre, C. J. (1980). Orogenic lherzolite complexes studied by 87Rb-87Sr: a clue to understand the mantle convection processes? Earth and Planetary Science Letters 51, 71-93.

Rapp, R. P. & Watson, E. B. (1995). Dehydration melting of metabasalt at 8-32 kbar: implications for continental growth and crust-mantle recycling. Journal of Petrology 36, 891-931.

Remaïdi, M. (1993). Etude pétrologique et géochimique d'une association péridotites réfractaires-pyroxenites dans le massif de Ronda (Espagne): implications pour les mécanismes de circulation des magmas dans le manteau supérieur. Ph.D. Thesis, Université de Montpellier II, 437 pp.

Rhodes, J. M. (1983). Homogeneity of lavas flows; chemical data for historic Mauna Loa eruption. Proceeding of the 13th Lunar and Planetary Science Conference, Part 2. Journal of Geophysical Research 88, Supplement, A869-A879.

Sawaguchi, T. & Takagi, H. (1997). Inverted ductile shear movement of the Horoman peridotite complex in the Hidaka metamorphic belt, Hokkaido, Japan. Memoirs of the Geological Society of Japan 47, 193-208.

Shimura, T., Komatsu, M. & Iiyama, J. T. (1992). Genesis of the lower crustal garnet-orthopyroxene tonalites (S-type) of the Hidaka Metamorphic Belt, northern Japan. Transactions of the Royal Society of Edinburgh: Earth Sciences 83, 259-268.

Shiotani, Y. & Niida, K. (1997). Origin of mafic layers in the Horoman peridotite complex. Memoirs of the Geological Society of Japan 47, 123-137 (in Japanese with English abstract).

Sobolev, A. V. & Shimizu, N. (1993). Ultra-depleted primary melt included in an olivine from the Mid-Atlantic Ridge. Nature 363, 151-154.

Staudigel, H., Zindler, A., Hart, S. R., Leslie, T., Chen, C.-Y. & Clague, D. (1984). The isotope systematics of a juvenile intraplate volcano: Pb, Nd, and Sr isotope ratios of basalts from Loihi Seamount, Hawaii. Earth and Planetary Science Letters 69, 13-29.

Suen, C. J. & Frey, F. A. (1987). Origins of the mafic and ultramafic rocks in the Ronda peridotite. Earth and Planetary Science Letters 85, 183-202.

Sun, S.-S. & McDonough, W. F. (1989). Chemical and isotopic systematics of oceanic basalts: implications for mantle composition and processes. In: Saunders, A. D. & Norry, M. J. (eds) Magmatism in the Ocean Basins. Geological Society, London, Special Publication 42, 313-345.

Takahashi, E., Nakajima, K. & Wright, T. L. (1998). Origin of the Columbia River basalts: melting model of a heterogeneous plume head. Earth and Planetary Science Letters 162, 63-80.

Takahashi, N. (1991). Origin of three peridotite suites from Horoman peridotite complex, Hokkaido, Japan; melting, melt segregation and solidification processes in the upper mantle. Journal of Mineralogy, Petrology and Economical Geology 86, 199-215.

Takahashi, N. & Arai, S. (1989). Textural and chemical features of chromian spinel-pyroxene symplectites in the Horoman peridotites, Hokkaido, Japan. Science Reports of the Institute of Geoscience, University of Tsukuba, Section B 10, 45-55.

Takazawa, E. (1989). Petrological study of the Horoman ultramafic massif in the Hidaka Metamorphic Belt, Hokkaido, Japan. M.Sc. Thesis, Hokkaido University, 150 pp.

Takazawa, E. (1996). Geodynamic evolution of the Horoman peridotite, Japan: geochemical study of asthenospheric and lithospheric processes. Ph.D. Dissertation, Massachusetts Institute of Technology, Cambridge, MA, 562 pp.

Takazawa, E., Frey, F. A., Shimizu, N., Obata, M. & Bodinier, J.-L. (1992). Geochemical evidence for melt migration and reaction in the upper mantle. Nature 359, 55-58.

Takazawa, E., Frey, F. A., Shimizu, N. & Obata, M. (1996). Evolution of the Horoman peridotite (Hokkaido, Japan): implications from pyroxene compositions. Chemical Geology (Special Issue: Melt Processes and Exhumation of Garnet, Spinel and Plagioclase Facies Mantle; 2nd Orogenic Lherzolites Conference, Granada) 134, 3-26.

Takazawa, E., Frey, F. A., Shimizu, N. & Obata, M. (1999). Whole-rock compositional variations in an upper mantle peridotite (Horoman, Hokkaido, Japan): implications for melt segregation, migration and reaction. Geochimica et Cosmochimica Acta (submitted).

Toyoshima, T. (1991). Tectonic evolution of the Hidaka Metamorphic Belt and its implication in Late Cretaceous-Middle Tertiary tectonics of Hokkaido, Japan. Science Reports of Niigata University, Series E (Geology and Mineralogy) 8, 1-107.

Tsuchiya, N., Shibata, T., Koide, Y., Owada, M., Takazawa, E., Goto, Y., Choi, J. H., Terada, S. & Hariya, Y. (1989). Major element analysis of rock samples by X-ray fluorescence spectrometry, using scandium anode tube. Journal of Faculty of Sciences, Hokkaido University, Series IV 22, 489-502.

Tsuruta, K. & Takahashi, E. (1998). Melting study of an alkali basalt JB-1 up to 12·5 GPa: behavior of potassium in the deep mantle. Physics of the Earth and Planetary Interiors 107, 119-130.

Varfalvy, V., Hébert, R. & Bédard, J. H. (1996). Interactions between melt and upper-mantle peridotites in the North Arm Mountain massif, Bay of Islands ophiolite, Newfoundland, Canada: implications for the genesis of boninitic and related magmas. Chemical Geology 129, 71-90.

Wilshire, H. G. & Shervais, J. W. (1975). Al-augite and Cr-diopside ultramafic xenoliths in basaltic rocks from western United States. Physics and Chemistry of the Earth 9, 257-272.

Yasuda, A., Fujii, T. & Kurita, K. (1994). Melting phase relations of an anhydrous mid-ocean ridge basalt from 3 to 20 GPa: implications for the behavior of subducted oceanic crust in the mantle. Journal of Geophysical Research 99, 9401-9414.

Yoshikawa, M. & Nakamura, E. (1999). Geochemical evolution of the Horoman Peridotite Complex: implications for melt extraction, metasomatism and compositional layering in the mantle. Journal of Geophysical Research (submitted).

Yoshikawa, M., Nakamura, E. & Takahashi, N. (1993). Rb-Sr isotope systematics in a phlogopite-bearing spinel lherzolite and its implications for age and origin of metasomatism in the Horoman peridotite complex, Hokkaido, Japan. Journal of Mineralogy, Petrology and Economical Geology 88, 121-130.

Zindler, A., Hart, S. R., Frey, F. A. & Jacobsson, S. P. (1979). Nd and Sr isotope ratios and rare earth element abundances in Reykjanes peninsula basalts: evidence for mantle heterogeneity beneath Iceland. Earth and Planetary Science Letters 45, 249-262.

Zindler, A. & Hart, S. R. (1986). Chemical geodynamics. Annual Review of Earth and Planetary Sciences 14, 493-571.


Top
*Corresponding author. Present address: Department of Geology, Faculty of Science, Niigata University, Niigata, 950-2181, Japan.Telephone: +81-25-262-6114. Fax: +81-25-262-6194.e-mail: takazawa@sc.niigata-u.ac.jp
This page is run by Oxford University Press, Great Clarendon Street, Oxford OX2 6DP, as part of the OUP Journals World Wide Web service.
Comments and feedback: jnl.info@oup.co.uk
Last modification: November 1999
Copyright© Oxford University Press, 1999.