Anda di halaman 1dari 36

International Geology Review, Vol. 47, 2005, p. 65100. Copyright 2005 by V. H. Winston & Son, Inc.

. All rights reserved.

Tehama-Colusa Serpentinite Mlange: A Remnant of Franciscan Jurassic Oceanic Lithosphere, Northern California
CLIFFORD A. HOPSON1
Department of Geological Sciences, University of California, Santa Barbara, California 93106-9630

EMILE A. PESSAGNO, JR.


Department of Geosciences, University of Texas at Dallas, P.O. Box 830688, Richardson, Texas 75083-0688

Abstract
The Coast Range ultramafic belt that borders the Sacramento Valley is largely a serpentinitematrix mlange, called the Tehama-Colusa serpentinite mlange (TCSM) after the counties it spans. It is bordered on the west across the Coast Range fault by exhumed high-P/T metamorphic rocks of the Franciscan subduction complex, and on the east, across the steep Stony Creek fault, by the JuraCretaceous terriginous clastic strata of the Great Valley Group (GVG). Remnants of the Coast Range ophiolite (CRO) that lie stratigraphically beneath the basal GVG are exposed only at the northern end of the Tehama-Colusa serpentinite mlange belt. The serpentinite protolith was peridotite tectonite (harzburgite > dunite), comprising oceanic upper mantle. Non-native mlange blocks in the serpentinite are chiefly basaltic submarine lava and sparse radiolarian ribbon chert (Middle and early Late Jurassic), plus rare plutonic rocks and slaty argillite. The serpentinite mlange protolith was basaltic oceanic crust above uppermost mantle, overlain by radiolarian chert and thin argillite (deep-sea clay?). An open-ocean setting, far from a continent margin or volcanic arc, is inferred from the absence of terriginous clastic or volcaniclastic sediments. The TCSM wraps around the Stonyford volcanic complex, a basaltic seamount that grew atop the basaltic oceanic crust during the late Middle and Late Jurassic. A widespread deepsea tectonic event in the Late Jurassic disrupted the proto-TCSM upper mantle and crust. Pervasive hydration turned peridotite into serpentinite, which invaded and mixed with the basalt/chert oceanic crust, creating mlange. Possible causes are discussed. The TCSM basaltic lavas and radiolarian chert closely resemble Franciscan oceanic crustal (not seamount) rocks, and are within the Franciscan basalt/chert age span; they are unlike the CRO igneous assemblage and its Late Jurassic tuffaceous chert. The TCSM was a segment of Franciscan oceanic lithosphere that records only Jurassic open-ocean depositional and tectonic history. It escaped latest JurassicCretaceous continent-margin and trench sedimentation, subduction, subduction accretion, high-P/T metamorphism, and exhumation of deeply subducted rocks that characterize the Franciscan Complex farther west in the Coast Ranges. The TCSM and CRO terranes are remnants of two different mid Jurassic paleoequatorial oceanic plates, separated now by the Stony CreekBeehive Flat fault system. TCSM evolution involves four successive stages of Mesozoic tectonic history: (1) Late Jurassic, when the CRO oceanic plate was being drawn NNE toward the Great Valley subduction zone in front of the oceanic Nevadan arc, trailed by TCSM (eastern Franciscan) lithosphere; (2) the latest Jurassicearliest Cretaceous, when (a) N-S dextral transform faulting replaced earlier dextral oblique subduction, bringing the Jurassic TCSM/Franciscan oceanic plate northward alongside the unsubducted CRO plate remnant, and (b) terriginous sediments from the new Nevadan orogen spread progressively seaward across the CRO and adjacent TCSM/Franciscan Jurassic ocean floor; (3) resumption of eastward subduction in the Early Cretaceous, from a new trench lying farther west, caused (a) disruption, mlanging, and metamorphism of the TithonianValanginian clastic strata and underlying Jurassic (Franciscan) oceanic crust, and (b) underthrusting, uplift, and eastward tilting of the unbroken TCSM/CRO oceanic basement, forming a new forearc ridge; and (4) a post-Valanginian Cretaceous era when Franciscan subduction operated west of the TCSM/CRO submarine ridge, and GVG forearc basin sedimentation
1Corresponding

author; email: hopson@geol.ucsb.edu

0020-6814/05/781/65-36 $25.00

65

66

HOPSON AND PESSAGNO

progressed behind the ridge on the east. The Stony CreekBeehive Flat composite fault sytem between the TCSM and CRO was a dextral transform fault during stage 2 (TithonianValanginian) and a W-vergent reverse fault during stage 3 (Valanginian), during underthrusting and tilting of TCSM/CRO basement. The extensional Coast Range fault, bounding TCSM on the west, brought up deeply subducted Franciscan Cretaceous high-P/T metamorphic rocks in the late Cretaceous (Jayko et al., 1987).

A Question of Identity
THE GEOLOGIC MAP of California (California Division of Mines and Geology, 1966, scale 1:2,500,000; Jennings, 1977, scale 1:750,000) and the Ukiah Sheet of the state geologic map (Jennings and Strand, 1960, scale 1:250,000) show a prominent N-Strending purple map unit along the eastern flank of the Coast Range, west of Sacramento Valley. This is shown as Ultramafic Rocks in the map legends and described as mostly serpentinite. But it also encompasses mafic rocks including remnants of a Jurassic ophiolite in the Digger CreekElder Creek area near Paskenta (Bailey et al., 1970; Hopson et al., 1981; Robertson, 1990; Blake et al., 1992; Shervais, 2001; Shervais et al., 2004), and a thick pile of mostly basaltic submarine lavas near Stonyford (Brown, 1964a, 1964b; Bailey et al, 1970; Shervais and Kimbrough, 1985b; Shervais and Hanan, 1989; Shervais et al., 2002 and in press, b) plus similar basalt with red chert near Wilbur Springs (Hopson et al., 1981; McLaughlin et al., 1989). Other submarine basalt-diabase-gabbroultramafic successions farther south in the Coast Ranges were also recognized as Jurassic ophiolites, were inferred to be remnants of former oceanic crust and upper mantle, and were collectively named the Coast Range ophiolite (Bailey et al., 1970). The remaining, more extensive serpentinite, and also the Stonyford and Wilbur Springs lavas, were then commonly portrayed as part of the Coast Range ophiolite (CRO). The serpentinite is commonly assumed to be the lower, relict mantle portion of the ophiolite because of its peridotite tectonite protolith and because it was situated at the apparent base of the tilted east-facing Digger CreekElder Creek CRO remnant and its Great Valley Group (Ingersoll, 1990) sedimentary cover. Moreover, the serpentinitic beltlike the ophioliteis unmetamorphosed except for low-P/T hydrous alteration; apparently it was never subducted. In contrast, it is faulted on the west against high-P/T metamorphic rocks of the South Fork Mountain Schist (SFMS), part of the widespread Franciscan Complex that comprises much of the

northern Coast Ranges (e.g., Blake et al., 1985, 1988, 1992). The SFMS protolithUpper Jurassic/ Lower Cretaceous (i.e., upper Tithonian to Valanginian) graywacke and mudstone, plus slices of basaltic lava and chertwere subducted to depths of >25 km, deformed, and metamorphosed at blueschistfacies P/T conditions (Jayko et al., 1986; Blake et al., 1988) during the Early Cretaceous (Suppe and Armstrong, 1972; Lanphere et al., 1978; McDowell et al., 1984), then uplifted tectonically and exhumed by extensional unroofing (Jayko et al., 1987; Harms et al., 1992) toward the rear of the seaward-growing Franciscan accretionary wedge (Platt, 1986). This contrast of metamorphic grades seemed further reason to link the serpentinite belt with the CRO, rather than with the local Franciscan (i.e., the South Fork Mountain schist). Thus, the regional geologic sketch maps within many CRO papers since 1970 show the entire width of the purple belt, including the Stonyford submarine lavas, as part of the California Coast Range ophiolite (or Jurassic ophiolites). This practice began with Bailey et al. (1970), Bailey and Blake (1974), Hopson et al. (1981), and Crerar et al. (1982), all of whom lumped the ultramafic purple belt with the Coast Range ophiolite. However, a contrary note was sounded by Shervais and Kimbrough (1985a, 1985b, 1987; also Hanan et al., 1991): This serpentinite-matrix mlange, which forms a belt that extends from Stonyford in the south to Paskenta on the north, has been interpreted as CRO by most recent investigators (e.g., Bailey et al., 1970; Bailey and Blake, 1974; Hopson et al., 1981). We suggest that correlation with the Franciscan complex may be more likely, based on the geochemical similarity between volcanic rocks of the mlange and Franciscan volcanic rocks. (Shervais and Kimbrough, 1985b, p. 828829) A prescient suggestion indeed! But most subsequent workers remained content to lump this huge ultramafic belt (Fig. 1) as part of the Coast Range

TEHAMA-COLUSA SERPENTINITE MLANGE

67

FIG. 1. Generalized geologic map of the northern Coast Range bordering the Sacramento Valley, showing the TehamaColusa serpentinite mlange and the Mesozoic terranes that border it on each side. Geology modified from Jennings and Strand (1960) with improvements from MacPherson (1983), McLaughlin et al. (1989), Blake et al. (1992), and Shervais et al. (in press, a).

68

HOPSON AND PESSAGNO

ophiolite on small regional maps that depict CRO distribution. These include MacPherson and Phipps (1985); Lagabrielle et al. (1986); Blake et al., (1987); McLaughlin et al. (1988); Seiders, (1988); Robertson (1989, 1990); Shervais (1990, 2001); Stern and Bloomer (1992); Cowan and Bruhn, 1992; Dickinson et al. (1996); Giaramita et al. (1998); Coleman (2000); Godfrey and Dilek (2000); Ingersoll (2000); Pessagno et al. (2000), Constenius et al. (2000), and Huot and Maury (2002). Even the Napa County serpentinite massifs south of Figure 1 are described as part of the Coast Range ophiolite (Phipps, 1984). Problems with this interpretation arose, however. First, the serpentinitic belt was not originally the base of the Coast Range ophiolite (Paskenta and Elder Creek remnants) but is separated from it by the Beehive Flat fault, which diagonally truncates the CRO members (Fig. 2). Second and more telling, much of the serpentinitic belt is a serpentinitematrix tectonic mlange (e.g., Hopson et al., 1981; Shervais and Kimbrough, 1985a, 1987; Jayko and Blake, 1986; Jayko et al., 1987; Robertson, 1990; Huot and Maury, 2002; Shervais et al., in press, b), not simply a serpentinized peridotite-tectonite basal member of the ophiolite succession. Up to 3040% of the serpentinitic belt is comprised of tectonic mlange blocks enclosed in a sheared, pulverized serpentinite matrix. Most of the blocks are altered submarine basaltic lava, including pillow lavas and lava capped by radiolarian ribbon chert. Native blocks of serpentinized harzburgite and dunite tectonitegrading into the enclosing sheared serpentinite matrixand sparse blocks of ribbon chert are widespread but minor. Diabase and plutonic blocks are rare. Jayko and coworkers mapped the serpentinite mlange separately from the ophiolite. They inferred that it lies structurally below the CRO and originated differently (Jayko and Blake, 1986; Jayko et al., 1987). But the two units are now faulted together and occasionally treated as a cohesive tectonic unit: The Coast Range ophiolite consists of a serpentinite-matrix mlange structurally overlain by a disrupted ophiolitic fragment. (Jayko et al., 1987, p. 1057) The present study takes another look at the serpentinite belt to assess its lithology, structure, age, and origin, and also the question: Is it part of the Coast Range ophiolite (CRO)? Or the Franciscan Complex? Or some different, perhaps

unrelated, terrane? How did it originate, and how was it emplaced in its present position? Our findings have important implications for the Mesozoic tectonic evolution of the region. The ultramafic belt is here called the TehamaColusa serpentinite mlange (TCSM) to better denote its geologic character and its geographic location spanning Tehama, Glenn, and Colusa counties N-S along the Coast Range front west of the Sacramento Valley.

Tehama-Colusa Serpentinite Mlange


The ultramafic belt (Fig. 1)the Tehama-Colusa serpentinite mlange (TCSM)is conveniently described in two segments: (1) a northern segment, informally called the Tehama serpentinite mlange (known also as the Round Mountain serpentinite mlange); and (2) a broad southern segment, the Colusa serpentinite mlange. They are described below, following a brief description of the serpentinite that is common to both segments.

The serpentinite The TCSM belt, whose protolith was peridotite, is pervasively serpentinized throughout its >90 km length. Massive serpentinite with relict peridotite textures is widely overprinted by rock with spaced shear surfaces (slickentite), schistose serpentinite, and locally serpentinite gouge, recording various degrees of internal deformation. Hard blocks of massive, serpentinized peridotite weather out with topographic relief, forming native mlange blocks that grade into surrounding, low-relief, sheared serpentinite or gouge. The serpentinite shear surfaces and schistosity are variable locally and warp around mlange blocks, but in general are subvertical and strike roughly N-S, subparallel to the outcrop belt. The serpentinite terrane has a characteristic vegetation marked by sparse although ubiquitous digger pines, leather oak, and abundant manzanita brush on the mountainsides, and also by cypress trees along some stream canyons. The vegetation cover varies from relatively open, with local bare rocky ground, to heavy brush cover that is difficult to penetrate on foot. In most cases, however, the large proportion of manzanita to other brush types is indicative of serpentinite bedrock. The serpentinite mineral assemblage is chiefly chrysotile-lizardite-brucite-magnetite chlorite. Antigorite has been reported (Huot and Maury, 2002) but seems not to be a major phase. Submicro-

TEHAMA-COLUSA SERPENTINITE MLANGE

69

FIG. 2. Generalized geologic map showing the northern (Tehama) segment of the Tehama-Colusa serpentinite mlange and adjacent Mesozoic map units. Geology modified from Blake et al. (1992) and Jayko and Blake (1986). Identification of CRO peridotite as chiefly dunite is from (Shervais et al., in press, a).

scopic serpentine pseudomorphs the tectonically deformed (warped, kink-banded) enstatite crystals, which stand out as bastite on weathered rock

surfaces. Original olivine crystals are replaced by microscopic serpentine minerals, brucite, and dusty magnetite. Sparse relict clinopyroxenes comprise

70

HOPSON AND PESSAGNO

several percent of some serpentinites and remain largely fresh. Cr-Al spinel (~23 %) is largely altered to opaque oxides. Carbonate minerals including magnesite locally fill fractures. Mafic dikes that cut the original peridotite are altered to whitish rodingite composed chiefly of hydrous CaAl silicate minerals. A detailed discussion of the serpentinite mineralogy, chemistry, and the serpentinization process in the Stonyford area is given by Shervais et al. (in press, a). The chrysotile-lizardite-brucite serpentinite is inferred to be the product of low-temperature hydrous alteration (Coleman, 1971). Antigorite serpentinite, the product of higher-temperature metamorphism, has not been reported from the Tehama-Colusa ultramafic belt.

west of Paskenta, and in a stream traverse through the spectacular gorge of Thomes Creek (i.e., the Paskenta, Bennett Creek, and Thomes Creek remnants described by Hopson et al., 1981). The mlange is quite similar in all three traverses, but the description of the Bennett Creek road section captures its essence (Hopson et al., 1981, p. 459): The ophiolite [i.e., serpentinite mlange] belt is well exposed along the Covalo road at Bennett Creek, 11 miles southwest of Paskenta. The belt is 1.7 km wide at this point and composed entirely of ophiolitic mlange. Steeply dipping faults bound the mlange on both sides, separating it from Tithonian strata of the Great Valley Sequence on the east and from metagraywacke and phyllite of the South Fork Mountain Schist (Franciscan) on the west. The ophiolite [serpentinite] mlange consists of tectonic blocks of pillow lava, massive lava (greenstone), and rare blocks of radiolarian chert, and remnants (native blocks) of serpentinized peridotite enclosed in a matrix of serpentinite. The blocks of volcanic rock range from less than 1 m up to 500 m in diameter, with blocks 100 to 200 m being commonplace. The larger blocks form resistant knobs and hillocks amid hummocky mlange terrrane. Volcanic blocks comprise approximately one-half to two-thirds of the mlange surface area, but the proportion of serpentinite is locally much higher. The mlange matrix consists of intensely sheared, slickensided serpentinite. Shear surfaces dip steeply and trend sub-parallel to the ophiolite [serpentinite] belt, but in detail these surfaces conform to the enclosed tectonic blocks. Commonly, the serpentinite cuts up through some of the larger blocks in dike-like sheets; here the serpentinite has clearly intruded or been plastically squeezed in along faults or shear zones. Locally, two or more subparallel serpentinite intrusions isolate slices of volcanic rock between them, and those slices become progressively disrupted into boudin-like blocks. Finally, zones of serpentinite 5 to 10 m wide contain trains of disrupted volcanic blocks or only isolated blocks. It is evident that a once-intact volcanic formation has been broken by faulting and shearing, and then pervasively invaded and

Tehama (Round Mountain) serpentinite mlange


The mlange character of the serpentinitic belt (Fig. 3A) and its cross-cutting fault contact against the Digger CreekElder Creek CRO remnant were recognized by Maxwell (1974), Fritz (1975), Hopson et al. (1981), Shervais and Kimbrough (1985a, 1987), Jayko and Blake (1986), and Jayko et al. (1987). Jayko and Blake (1986) provided an informative small-scale map of the serpentinite mlange showing distribution of the larger mlange blocks. Hopson et al. (1981) named this and similar rocks to the south the Tehama County Ophiolite mlange belt, from which the present name is modified. But they left it as part of the CRO for want of definitive evidence to the contrary. However, Jayko et al. (1987, p. 480) noted that In the Paskenta area a small but complete remnant of the Coast Range ophiolite is structurally underlain by a serpentinite mlange, which they depicted separately from CRO on a geologic map (their Fig. 6; see also Jayko and Blake, 1986; Blake et al., 1992). They described and interpreted this unit, which they called the Round Mountain serpentinite-matrix mlange, (see also Huot and Maury, 2002). Robertson (1990) also showed this belt as tectonic mlange faulted against the CRO units. Shervais and Kimbrough (1985b, 1987) and Shervais and Hanan (1989) recognized that serpentinite-matrix mlange comprised the ultramafic belt in the Stonyford area, since mapped by Shervais and Zoglman-Schuman (Shervais et al., in press, a, b). The Tehama (Round Mountain) serpentinite mlange (Fig. 2) is well exposed (less now than formerly, due to roadcut deterioration) along road traverses through the ultramafic belt west and south-

TEHAMA-COLUSA SERPENTINITE MLANGE

71

FIG. 3. Outcrop photographs of typical rocks within the northern Tehama-Colusa (Round Mountain) serpentinite mlange. A. Typical serpentinite mlange southwest of Paskenta. Dark rock is serpentinite matrix; light-hued rocks are weathered basaltic lava mlange blocks. B. Serpentinite mlange at Black Diamond Glade northwest of Stonyford, showing the characteristic sparse vegetation with abundant manzanita; low rounded brushy hill in the center is a basaltic mlange block. C. Pillow lava within a basaltic mlange block; Tehama (Round Mountain) serpentinite mlange west of Crowfoot Point. The pillows, resting on dark basaltic mud, are upside down. D. Pillow lava within a basaltic mlange block; Tehama (Round Mountain) serpentinite mlange northwest of Bennett Creek. E. Radiolarian ribbon chert within a composite basalt-chert mlange megablock; Tehama serpentinite mlange west of Crowfoot Point. F. Part of a weathered basaltic mlange block (lower left) capped by radiolarian ribbon chert (dark rock, partly behind tree); Tehama mlange along Road M2 near Round Mountain.

internally disrupted by tectonically mobile serpentinite, creating a mlange. The serpentinite is composed of chrysotilelizardite serpentine and brucite, indicative of formation at low temperature (Coleman,

1971b). It originated mainly from harzburgite, for serpentinized harzburgite is the ubiquitous type of native mlange block. Pillow lava is abundant among the volcanic mlange blocks. One large block contains 7 m

72

HOPSON AND PESSAGNO

of thin-bedded radiolarian chert resting in depositional contact upon pillows. Petrographically, these lavas are spilites derived chiefly from aphyric and plagioclase-microphyric basalt. Primary igneous textures, including variolitic quench textures, are well preserved, but alteration has produced mainly secondary assemblages of albitechlorite-epidote-sphene-calcite quartz. Clinopyroxene is the main surviving volcanic phase. The tectonic block assemblage within the mlange and also the blocks themselves (Fig. 3) resemble Franciscan basaltic lava and chert, and differ significantly from rocks in the CRO. Comparisons are deferred to later sections. Red Mountain, at the southern end of the Tehama serpentinite mlange belt west of Chrome (Fig. 2), exposes massive serpentinized harzburgite without mlange blocks. Jayko and Blake (1986, Fig. 3) and Jayko et al. (1987, Fig. 6) draw a contact that separates this harzburgitic area from the adjacent serpentinite mlange and show it as CRO. But we regard it instead as a change from serpentinized mantle peridotite mixed with blocks of oceanic crusti.e., the mlangeto more massive serpentinized peridotite without blocks. This may provide a glimpse of once-deeper mantle below the level of admixed crustal blocks. Such a transition from serpentinite mlange to massive serpentinized peridotite (without mlange blocks) is common farther south, in the Colusa mlange. The TCSM belt in Glenn County (Fig. 1) consists of a narrow (0 to 250 m) strip of serpentinite, bounded on both sides by steep faults (Jennings and Strand, 1960; Blake et al., 1992). Its appearence was described by Hopson et al. (1981, p. 458): A typical section is well exposed where this belt is crossed by the south fork of Elk Creek. Here the ophiolite is represented by only 20 to 40 m of strongly sheared and pulverized serpentinite. The eastern contact is a vertical fault (Stony Creek fault) that truncated the bedding of steeply dipping Tithonian mudstones belonging to the lower Great Valley Sequence. The western contact is another vertical fault that sets the serpentinite against dark slaty argillite and metagraywacke of the South Fork Mountain Schist (Franciscan Complex). This fault is sharp, lacks imbrica-

tion, and does not noticeably disturb the adjacent bedded metasedimentary sequence. Basaltic mlange blocks are locally present (Blake et al., 1992) but inconspicuous in this narrow serpentinitic septum, where weathering and poor exposure obscures details. We concur with Jayko and Blake (Jayko and Blake, 1986; Jayko et al., 1987; Blake et al., 1992) that this narrow strip is a continuation of the Round Mountain (Tehama) serpentinite mlange.

Colusa serpentinite mlange The name change from Tehama to Colusa segment of the TCSM belt is arbitrarily placed near the Colusa County line, where the belt turns westward to curve around the Stonyford volcanic complex (Fig. 1). The belt southward from here crops out in Colusa County. This southern segment of the ultramafic belt exposes serpentinite-matrix mlange similar to the Tehama mlange, but with patchy distribution of mlange blocks and large areas of serpentinized peridotite devoid of blocks. Our findings come from three widely spaced traverses across the belt: (1) the Dry CreekBlack Diamond Ridge road north of Stonyford: (2) the Goat Mountain Road through Little Stony Creek canyon; and (3) the Bartlett Springs Road from Bear Valley to Little Indian Valley Reservoir. The Dry CreekBlack Diamond Ridge section exposes chiefly sheared serpentinite at lower elevations with blocks increasing in frequency up the eastern flank of Black Diamond Ridge. Several basaltic blocks up to 250 m or more across occur in the Black Diamond Glade area. One of these (Fig. 3B) is rimmed (overlain) by >20 m of dark red ribbon chert, then by dark argillite soil. Three other slabs of dark-grey slaty argillite, each tens of meters long, crop out lower along this traverse. This mlange changes westward into more massive, less serpentinized peridotite tectonite along Black Diamond Ridge itself. The Little Stony Creek Canyon traverse (Goat Mountain Road) through the TCSM reveals a similar pattern. Here too, crustal mlange blocks in sheared serpentinite are found only along the eastern side of the belt, changing to blocky serpentininized peridotitelocally sheared but devoid of mlange blocksthrough the interior and western side. Perhaps the oceanic mantle is more deeply exposed here, beneath the level of disrupted oceanic crust. The geologic map of the Stonyford area (in Shervais et al., in press, a, b), spanning the N-S interval

TEHAMA-COLUSA SERPENTINITE MLANGE

73

from Black Diamond Ridge to Little Stony Creek, also shows an eastern belt of sheared serpentinite with mlange blocks and a western zone of massive, partly serpentinized harzburgite. Our observations accord with theirs. Also, two megablocks of volcanogenic sandstone are shown within the serpentinite south of Stonyford, on the Shervais map. But we regard them instead as remnants of ophiolitic sandstone (from former basal GVG thrust over the serpentinite) that were let down as erosional lag blocks upon the TCSM surface. We find a similar sandstone block resting upon (not within) the sheared serpentinite nearby at Little Stony Creek. The mlange is best developed on the south, along the Bartlett Springs Road, partly within the 1:24,000-scale geologic map of the Little Indian ValleyWilbur Springs area (McLaughlin et al., 1989). Here sheared to massive serpentinite encloses tectonic blocks up to 200 m or more that comprise approximately 20% (locally up to 40%) of the bedrock. Most of the mlange blocks are basaltic submarine lava. The lavas are strongly altered, and commonly sheared internally with chlorite coating the slip surfaces. Pillow structure is preserved locally. Sedimentary blocks are rare; slaty argillite is exposed at one roadside locality. The complex structural relations of the serpentinite belt from here southward to its termination near Wilbur Springs are well shown on the 1:24,000-scale geologic map and cross sections of McLaughlin et al. (1989). Although their 1989 geologic map shows the Colusa serpentinite mlange as serpentinized peridotite (Jsp) of the Coast Range ophiolite, and a 3 km long slab of basalt faulted against Jsp as also part of the CRO (following Bailey et al., 1970, and Hopson et al., 1981), our reassessment of red radiolarian chert within that basalt (Table 1, and discussion in the section on mlange blocks) now identify the chert-bearing basalt as a TCSM mlange block. Our investigation of the TCSM belt ends at Wilbur Springs (Fig. 1), but its apparent continuation (i.e., direct alignment) with the large serpentinite masses farther south in Lake and Napa counties (to Lake Berryessa) requires comment. Moiseyev (1966, 1970) pointed out that those serpentinites are of two types: intact masses derived from peridotite, and detrital serpentinitesthe latter described later in a classic paper on Ophiolitic olistostromes in the basal Great Valley sequence, Napa County by S. P. Phipps (1984). He noted that a thick (up to 1 km), laterally extensive serpentinous chaotic

unit lies directly above the serpentinite that represents the Coast Range ophiolite and below the well bedded Great Valley sequence of Upper Jurassic and Cretaceous age (p. 103). And also: Throughout the study area the chaotic unit lies directly above serpentinite that represents the Coast Range ophiolite (p. 121). No mafic section ever intervenes (p. 121). From this and other evidence discussed later, we infer that the large serpentinite masses below the chaotic serpentinous olistostrome unit is not the CRO but rather a southward extension of the TCSM belt. The serpentinite was derived from harzburgite tectonite (Phipps, 1984) but evidently without mlange blocks. Perhaps the shallow serpentinite-mlange component was never present this far south. Or, the mlange carapace may have slid off and its remnants mixed within the ophiolitic olistostrome unit. The significance of the olistostrome unit is discussed later.

Summary
The entire length of the Tehama-Colusa ultramafic belt (TCSM) is serpentinite-matrix mlange, at least in part, with mlange blocks composed chiefly of oceanic basalt, minor radiolarian ribbon chert, plus rare plutonic rocks and dark slaty argillite. The serpentinite host was derived from peridotite tectonite (harzburgite > dunite). The mlange is best developed on the north (Tehama serpentinite mlange) but is recognizable (at least locally) all along the ultramafic belt. From Chrome southward, the mlange blocks amid sheared serpentinite are more common along the eastern side of the TCSM belt, and large areas of massive, less-serpentinized harzburgite without mlange blocks crop out on the western side. This is seen west of Chrome (Fig. 2; also Jayko and Blake, 1986) and across the TCSM belt at Black Diamond Ridge and in Little Stony Creek Canyon (Shervais et al., in press, a, b). If the block-bearing mlange is indicative of strongly disrupted and pervasively serpentinized oceanic crust and uppermost mantle, and the massive, less serpentinized harzburgite once lay deeper, then the TCSM is a warped oceanic block tilted eastward, exposing deeper mantle levels to the west.

Metagraywacke-slate-greenstone slices along the Coast Range fault


The high-angle Coast Range fault brought up the once-deeply buried South Fork Mountain schist (SFMS) on the west against the Coast Range

74

HOPSON AND PESSAGNO

TABLE 1. Biostratigraphic Data from TCSM Radiolarian Cherts


Sample PAS-03-B2: dark grey, non-tuffaceous radiolarian chert Geologic occurrence: Near base of 14 m ribbon chert succession capping weathered basaltic lava, composite mlange block in northern TCSM belt (Round Mountain serpentinite mlange). Sample location: Forest Road M2 (Toomes Camp Rd.), 4.3 miles (road distance) beyond Crowfoot Point, SE 1/4 of Sect. 28, T24N, R7W. GPS = N3954' 20.4"; W12239' 3.7". Collected by: Emile Pessagno Jr., June 2003. Processed August 2004. Radiolarian assemblage: Ristola procera Pessagno, Praecaneta turpicula Pessagno, Praecaneta decora Pessagno, Pantanellium josephinense Pessagno, Blome, and Hull, Pantanellium meraceibaense Group Pessagno, Bernoullius sp., Archaeospongoprunum sp., Spongocapsulum sp. Biostratigraphic determination: Superzone 1, Zone 1I, and possibly also Zone 2, Subzone 2. Chronostratigraphic assignment: Middle Jurassic (upper Callovian) to Upper Jurassic (lower Oxfordian to possibly middle Oxfordian). Paleobiogeographic determination: Central Tethyan. Fauna contains abundant members of the Pantanellium meraceibaense Pessagno Group. Members of this group are abundant in Tethyan strata of east-central Mexico and central Mexico. Moreover, they are quite abundant in the volcanopelagic succession that overlies the Josephine ophiolite. Virtually no specimens of Praeparvicingula (with horn) are present in this assemblage. Clearly Central Tethyan. Sample PK 14-2. Red, non-tuffaceous radiolarian chert Geologic occurrence: Near top of 14 m ribbon chert succession capping weathered basaltic lava, composite mlange block in northern TCSM (Round Mountain serpentinite mlange). Sample location: Forest Road M2 (Toomes Camp Rd.), 4.3 miles (road distance) beyond Crowfoot Point, SE 1/4 of Sect. 28, T24N, R7W. GPS = N3954' 20.4"; W12239' 3.7". Collected by: Cliff Hopson, Sept. 1976. Processed 1977. Re-examined April 2003. Radiolarian assemblage: Ristola procera Pessagno, Praecaneta turpicula Pessagno and Whalen, Praecaneta decora Pessagno, Hsuum brevicostata Ozvoldova. Biostratigraphic determination: Superzone 1, Zone 1I (based on occurrence of R. procera and P turpicula). Chronostratigraphic assignment: Middle Jurassic (upper Callovian) to Late Jurassic (lower Oxfordian). Paleobiogeographic determination: Central Tethyan (criteria of Pessagno and Blome, 1986; Pessagno et al., 1993). Sample DF-PK-1. Red, nontuffaceous radiolarian chert Geologic occurrence: Ribbon chert mlange block in northern TCSM belt (Round Mountain serpentinite mlange). Sample location: Forest Road M2 (Toomes Camp Rd.) between Crowfoot Point and Round Mountain saddle (exact location not recorded). Collected by: Debra Fritz (M.A. thesis, University of Texas, Austin, Texas, 1974). Re-examined by E. Pessagno, April 2003. Radiolarian assemblage: Praecaneta decora Pessagno and Whelan. The only diagnostic taxon identified. The fauna lacks E. pyctum and Mirifusus. These latter taxa are both easy to recognize and the abundance and preservation are good enough to determine that both fauna are clearly absent. Biostratigraphic determination: Superzone 1, Zone 1F to Zone 1I (lower part). Chronostratigraphic assignment: Middle Jurassic (upper Bathonian to upper Callovian). Paleobiogeographic determination: Central Tethyan (criteria of Pessagno and Blome, 1986; Pessagno et al., 1993). Table continues

TEHAMA-COLUSA SERPENTINITE MLANGE

75

TABLE 1. (Continued)
Louvion-Trellu, 1986 determinations Data source: Louvion-Trellu, 1986: DEA thesis, Universite de Bretagne Occidentale, Brest, France. Quoted in Huot and Maury, 2002. Geologic occurrence: Ribbon chert mlange blocks in Round Mountain serpentinite mlange, northern TCSB belt. Sample locations: Along (and near?) Forest Road M2 (Toomes Camp Rd.) between Crowfoot Point and Round Mountain saddle (exact locations not given in Huot and Maury, 2002). Biostratigraphy: Faunal assemblages and zonal data not given in Huot and Maury (2002). Chronostratigraphic assignment: Middle Jurassic (Callovian) and Late Jurassic (upper Oxfordian). Sample NCF 997A, B, C. Red, manganiferous (non-tuffaceous) radiolarian chert Geologic occurrence: Red manganiferous chert interbedded with vesicular [amygdaloidal] basalts in 3 km long mass of submarine basalt faulted against serpentinite (southernmost TCSM belt) on the north and juxtaposed discordantly (faulted?) against terriginous clastic strata of Great Valley Group on the south. Sample location: Outcrop on ridge west of Sulphur Creek east fork, 0.58 km S69W of Eagle Rock, USGS Wilbur Springs 7.5' quadrangle; Section 18, T14N, R5W. Collected by: Emile Pessagno, 1977. Processed in 1977 and initial results reported in Micropaleontology, no. 1, p. 104, 1977. Results reported below are from a re-examination in Sept. 2004. Radiolarian assemblage: Ristola altissima (Ruest), Mirifusus baileyi (Pessagno, Mirifusus guadalupensis Pessagno, Mirifusis mediodilata (Ruest), Parvicingula turrita (Ruest), Caneta hsui (Pessagno), Podobursa sp., Saitoum pagei Pessagno, Eucryrtidium ptyctum (Riedel and Sanfilippo), Archaeodictyomitra rigida Pessagno, Pantanellium sp., Acanthocircus variabilis Squinabol. Biostratigraphic determination: Zone 3, Subzone 3. (Concurrence of Mirifusus baileyi, Caneta hsui, and M. guadalupensis pin the age of this sample). Chronostratigraphic assignment: Upper Jurassic (uppermost Kimmeridgian to lower Tithonian). Paleobiogeography: Tethyan Realm, Northern Tethyan Province.

ophiolite and the serpentinite mlange on the east, during extensional exhumation of the Eastern Franciscan belt in the Late Cretaceous (Jayko et al., 1987; Blake et al., 1988). About ten fault-bounded slabs of slate, metagraywacke, and greenstone crop out along a 50 km segment of the Coast Range fault from near Yolla Bolly junction (southern end of the western Klamath Mountains) on the north to near Elk Creek on the south (Jayko and Blake, 1986, Fig. 3). The seven southernmost slabs occur along the SFMSserpentinite mlange contact (Fig. 2), in most cases occuring slightly within the serpentinite where they might locally be mistaken for mlange blocks. A detailed study by Jaydo and Blake (1986) clearly demonstrated, however, that these slices are exotic, i.e., low-grade (prehnite-pumpellyite facies) metamorphic rocks whose metamorphic mineralogy, fabric, sandstone petrology, and K/Ar age (~158 Ma) are quite different from the SFMS, the TCSM

mlange blocks, the CRO rocks, and the GVG strata that bound them on both sides. Their remarkable resemblance to rocks of the Late Jurassic (Oxfordian) Galice Formation of the western and southern Klamath Mountains is pointed out, along with possible emplacement scenarios along the Coast Range fault (Jayko and Blake, 1986). The main point made here is that these slate/ graywacke greenstone slices are not mlange blocks within the TCSM ultramafic belt, even though some of them occur within serpentinite close to the present Coast Range fault. That is, they were not part of the oceanic crustal assemblagenow disruptedthat once lay atop the TCSM upper mantle peridotite. The composition of these slivers, their low metamorphic grade, and especially their occurence along 50 km of the Coast Range fault, suggests to us that these were Franciscan rocks that once lay structurally above the higher-grade South

76

HOPSON AND PESSAGNO

Fork Mountain schist, and have since been displaced against it. Extensional exhumation of the SFMS in the Late Cretaceous (Jayko et al., 1987) might have left slivers of those less deeply buried roof rocks clinging to the serpentinite as the SFMS block was brought up against it. This explanation accords best with the tectonic history of the TCSM terrane, proposed in a later section.

have mineralogical and textural features of abyssal peridotites including interstitial plagioclase (pl) associated with amygdaloidal [poikilitic?] clinopyroxenes. These phases are interpreted as evidence of impregnation left by magma perculation (Huot and Maury, 2002). And further: Less serpentinized clinopyroxene harzburgites contain partly preserved olivine, orthopyroxene, clinopyroxene, and spinel, the proportions of which are close to the modal mineral contents of abyssal plagioclase-free peridotites. (Huot and Maury, 2002, p. 113) Mineral compositions are given for the olivine, orthopyroxene (enstatite), clinopyroxene, and spinel, and they (Huot and Maury, 2002, p. 114) concluded that: Spinels from serpentinized clinopyroxene harzburgites have Cr# values [100 Cr/Cr + Al] similar to those of spinels from abyssal peridotites reported in Dick and Bullen (1984). The harzburgite sensu stricto (s.s.) and opx dunite have higher Cr# values of 65 and 7274, respectively, which fall within the wide range of alpinetype [including ophiolite] peridotites and also forearc peridotite (Huot and Maury, 2002, Fig. 5). The presence of orthopyroxene (enstatite) dunite among the peridotite blocks is significant. In wellpreserved ophiolites these typically occur at the base of the Moho transition zone (Boudier and Nicolas, 1995; Nicolas and Boudier, 2000), where dunite s.s. at the top of oceanic mantle sections grades down through enstatite dunite to impregnated cpx-pl harzburgite to harzburgite s.s. (Pallister and Hopson, 1981; Nicolas and Prinzhofer, 1983; Benn et al., 1988; Nicolas, 1988; Juteau et al., 1988). Also, the same succession from impregnated dunites and impregnated harzburgite to harzburgite s.s. was recovered in the crust-mantle transition zone (MTZ) of 1 Ma East Pacifc Rise oceanic lithosphere, exposed in Hess Deep (Hekinian et al., 1993; Girardeau and Francheteau, 1993). This petrologic sequence is due to the partial to complete dissolution of enstatite in picritic melt expelled from upwelling upper mantle harzburgite at low pressure (Quick, 1981; Pallister and Gregory, 1983; Kelemen, 1990; Kelemen and Dick, 1995) beneath oceanic spreading centers (Ceuleneer and Rabinowicz, 1992). The resulting peridotite sequence (top downward) in uppermost oceanic mantle (recorded

Mlange Blocks
Tectonic blocks within the Tehama-Colusa serpentinite mlange belt include native blocks of serpentinized peridotite and oceanic crustal blocks of submarine basaltic lava, radiolarian ribbon chert, andin the Colusa sector of TCSMslabs of slaty argillite, all enclosed in a matrix of variably deformed serpentinite. The occurrence of tectonic blocks of garnet amphibolite and their possible relationship to the TCSM belt are also discussed.

Serpentinized peridotite Resistant blocks of massive serpentinized peridotite weather out in relief above softer serpentinite host rock, which is commonly deformed to slickentite (spaced shear surfaces), schistose serpentinite, or soft gouge. These are native mlange blocks, not externally derived but standing out because of resistance to weathering and erosion. The original character of the serpentinized peridotite is best shown in the relatively massive blocks. Protolithic harzburgite is distinguished from dunite by the presence of serpentinized pyroxenes (bastite), conspicuous in outcrop and thin sections. But closer scrutiny (Huot and Maury, 2002) reveals further important details. Their work in the Round Mountain area (north end of Tehama mlange) found that:
Despite their extreme serpentinization (>65%) and commonly highly tectonized character, the studied examples of ultramafic rocks can still be identified as clinopyroxene harzburgite, harzburgite senso stricto (s.s.), and orthopyroxene dunite. In most samples, olivine, orthopyroxene (1025 modal %), and rare clinopyroxene (<5%) are totally transformed into various serpentine polymorphs and magnetite. (Huot and Maury, 2002, p. 112113). They deduced that the clinopyroxene (cpx) harzburgites, which are common in the melange,

TEHAMA-COLUSA SERPENTINITE MLANGE

77

in ophiolites) is dunite s.s., resorbed-enstatite dunite, cpx (pl) harzburgite (abyssal peridotite), and harzburgite s.s. (depleted upper mantle), respectively. Those rocks above the harzburgite s.s. are commonly streaked with gabbroic (pl + cpx + ol) impregnations from entrapped picritic melts expelled from deeper mantle (Nicolas and Prinzhofer, 1983; Ceuleneer and Rabinowicz, 1992; Hekinian et al., 1993; Boudier and Nicolas, 1995; Hopson et al., in prep.). One of us has observed well-exposed gradational sequences from dunite (a melt-extraction cumulate) downward through resorbed-enstatite dunite to cpx ( pl) harzburgite tectonite and then harzburgite s.s. in the Wadi Gideah East traverse across the mantle-crust transition zone (MTZ) of the Oman ophiolite, and also the same succession with abundant gabbroic impregnations (pl + cpx + ol) in four traverses across the Oman ophiolite mantle-crust transition zone, especially in the bare mountainside east of Khaffifah (Pallister and Hopson, 1981). Analyzed clinopyroxene harzburgite mlange blocks in the Tehama (Round Mountain) serpentinite mlange also correspond geochemically to abyssal peridotite from the Mid-Atlantic Ridge and from Hess Deep near the East Pacific Rise (Huot and Maury, 2002, Fig. 7). The common occurence of abyssal peridotite (including cpx-pl-impregnated harzburgite) and enstatite dunite in the serpentinite mlange block assemblage is strongly suggestive of uppermost oceanic mantle, including the dunitic Moho transition zone (MTZ) occurring just beneath basaltic/gabbroic oceanic crust. Such a peridotite assemblage is unlikely in deeper sections of the oceanic mantle, including the hanging walls of subduction zones.

Basaltic lava
Basaltic mlange blocks range in size from kilometer-scale down to meter scale (Fig. 3A). Jayko and Blake (1986) and Blake et al. (1992) mapped basaltic mlange blocks up to 4 km long, with several having kilometer-scale dimensions (Fig. 2). The Round Mountain basaltic block in the Tehama mlange west of Paskenta measures 2.5 km by 1 km, and a lesser but still immense basaltic slab at Thomes Creek gorge is at least 1200 m by 500 m by >200 m (Hopson et al., 1981). Most blocks, however, are less than 250 m across. The basaltic blocks commonly weather out in topographic relief above the less resistant serpentinite. These are readily apparent in the moderately vegetated Tehama

mlange but less obvious in the brushier Colusa mlange. Traverses through the latter commonly find the basaltic blocks standing out as craggy pinnacles. But many basaltic blocksweathered to soilform only low mounds densely covered with brush. Yet, these mounds stand out from the enclosing manzanita-cloaked serpentinite by their distinctive vegetation cover, i.e., dense brush with abundant dark green chamise (Adenostoma fasciculatum) (Fig. 3B) whose tan-hued blooms give an overall orange appearence when viewed from nearby. Pillow lava, massive lava, and basaltic breccias (including submarine talus breccia) are found in the mlange blocks (Figs. 3C and 3D). Pillow and breccia structure, however, are commonly deformed or obliterated by shearing and/or weathering. The deformation and low-grade alteration partly mask the original composition, textures, and structures of the mafic lavas. Most of the lavas are aphyric megascopically and intersertal microscopically. Microphyric basalts have plagioclase or clinopyroxene-plagioclase microphenocrysts; phyric olivinealtered to green phyllosilicates or carbonateis less easily recognized. Groundmass glass is ubiquitously transformed into variable proportions of authigenic clays, chlorite, and pumpellyite (Huot and Maury, 2002, p. 114). Internal shear surfaces, along which the rocks break, are coated with chlorite or other green phyllosilicates. Immobile trace element geochemistry, however, sees through the mask of rock alteration and deformation. Shervais and Kimbrough (1985a, 1987) analyzed basaltic lavas from three Paskenta mlange blocks (i.e., Tehama serpentinite mlange west of Paskenta) and compared them with volcanic rocks from the Coast Range ophiolite, the Stonyford seamont lavas, and some Franciscan lavas. They reported, In contrast [to CRO lavas], volcanic rocks from the Stonyford seamount and the Paskenta mlange blocks are transitional subalkaline basalts that are compositionally similar to enriched MORB or ocean-island tholeiites (Shervais and Kimbrough, 1985a, p. 37, 1987, Fig. 5, p. 176). Jayko et al. (1987, p. 481) concluded that This difference in chemistry between the ophiolite remnant and blocks within the Round Mountain (Tehama) serpentinite mlange implies that the mlange blocks were not derived from the ophiolite. Huot and Maury (2002), with a larger data base of 14 analyzed samples of basalt and diabase from

78

HOPSON AND PESSAGNO

the mlange blocks, and employing the immobile trace element classification of Winchester and Floyd (1977), found that all samples have subalkaline affinities and their compositions range from basalt to basaltic andesite. Furthermore, The majority of the mafic blocks define a trend marked by a TiO2 enrichment along the liquid line of descent. A detailed discussion of the trace-element geochemistry led to the following conclusion: Only one main magmatic series of subalkaline affinity is represented in the Round Mountain mafic blocks. This conclusion is essentially supported by TiO2, Zr, REE abundances, and LREE-depleted patterns similar to N-MORB. (Huot and Maury, 2002, p. 118) However: Sample 38, with a LREE-enriched pattern and the highest (Zr/Y) ratio of our set, is thought to be the only sampled transitional oceanic tholeiite. Such transitional MORBs are commonly reported in mafic lithologies from the Franciscan assemblages (MacPherson, 1983; Jayko, 1984; Shervais and Kimbrough, 1987; Shervais and Hanan, 1989; Shervais, 1990). (Huot and Maury, 2002, p. 118) Two of the three mlange-block lavas studied by Shervais and Kimbrough (1985a, 1987) are of this latter type, leading to their conclusion that the mlange mafic lavas are like those at Franciscan localities. Plotting all 14 of their basaltic samples on a TiO2 vs 100 Mg/(Mg +Fe2+) diagram, Huot and Maury (2002) showed four in the Franciscan metabasalt field (which is weighted toward highTiO2 seamount lavas) and most of the others clustered just beneath it, partly overlapping the CRO lavas, which have lower TiO2. We conclude tentatively that the Round Mountain mlange-block lavas (northern TCSM) are similar to Franciscan spreading-center lavas but not Franciscan seamounts. Perhaps an even more telling basis for comparing the crustal mlange blocks with the Franciscan versus the CRO is the contrast between their subvolcanic and plutonic assemblages, noted next.

Jayko et al. (1987) both mentioned the rarity of diabase and gabbro blocks in the mlange. Huot and Maury (2002) found rare diabase but only one block of gabbro. Diorite, clinopyroxenite, and wehrlite mlange blocks are reported from one local area (Shervais et al., in press, b). The appearence of minor wehrlite/clinopyroxenite in the mlange is not surprising, inasmuch as they commonly occur with cpx-impregnated dunite in the transition zone (MTZ) at the top of oceanic upper mantle (Nicolas and Prinzhofer, 1983; Juteau et al., 1988; Boudier and Nicolas, 1995). Related cpx-impregnated uppermost mantle (MTZ) rocks have already been described elsewhere in the TCSM belt (see serpentinized peridotite, above). Franciscan igneous assemblages throughout the California Coast Ranges comprise chiefly basaltic lavas (basaltic greenstone) and serpentinized peridotite; diabase and gabbro are rare (e.g., Bailey et al., 1964; Blake et al., 1985). In contrast, diabasic dike and sill complexes, isotropic and layered (cumulus) gabbros, and clinopyroxenite/wehrlite are prominent components of well-preserved CRO crustal successions (e.g., Bailey et al., 1970; Bailey and Blake, 1974; Hopson et al., 1981, 1996, and in prep.; Hopson, 2002; Shervais, 2001; Shervais et al., 2004). Thus, igneous/meta-igneous blocks in the Tehama-Colusa serpentinite mlange match the Franciscan but are quite unlike the Coast Range ophiolite igneous assemblage.

Radiolarian ribbon chert


Radiolarian chert comprises only a small proportion (<5 %) of the mlange blocks, but they provide invaluable information about the age of the original oceanic crust, its tectonic setting, and comparison with both the Franciscan radiolarian cherts and those associated with the Coast Range ophiolite. The chert in the mlange blocks are typical ribbon cherts, with thin shaly partings between the discrete chert beds (Fig. 3E). The Tehama/Colusa cherts are mostly brick red, brown, dark grey, and black, veined by fracture-filling quartz. The cherts are chiefly siliceous and non-tuffaceous; thin sections reveal abundant radiolaria remnants but are without volcanic microlites, shards, or pumice lapilli, except the rare exception (hornblende-plagioclase microlites) noted by Huot and Maury (2002). Thus, the Tehama/Colusa mlange-block cherts strongly resemble those of the Franciscan Complex (Bailey et al., 1964, Figs. 26, 28, 29), and are quite unlike the greenish grey (white-weathering) tuffaceous cherts

Plutonic and subvolcanic rocks Despite the remarkable abundance and size of basaltic lava blocks in the Tehama/Colusa serpentinite mlange, blocks of diabase and especially plutonic rocks are rare. Hopson et al. (1981) and

TEHAMA-COLUSA SERPENTINITE MLANGE

79

and tuffaceous radiolarian mudstones that overlie the CRO (Hopson et al., 1981, 1996, in prep.; Pessagno et al., 2000; Robertson, 1989). The mlange cherts occur both alone (as deformed slabs of ribbon chert) and as part of composite blocks with chert beds attached to basaltic lava (Fig. 3F). Such composite blocks are inferred to have been pelagic radiolarite beds deposited atop basaltic oceanic crust. Most of the chert remnants are only as few meters thick, with parts evidently sheared off. A 25 m ribbon chert succession resting on basaltic lava along the Toomes Camp road (SE1/ 4 of Sect. 28, T24N, R7W), is the thickest continuous section of chert observed. Work in progress indicates that the mlangeblock cherts are late Middle Jurassic in age, extending into the Late Jurassic (Table 1). The radiolarian assemblage once considered to be OxfordianKimmeridgian (Pessagno, in Hopson et al., 1981) is now reassigned to the Callovian to upper Oxfordian. Dark gray ribbon chert (Fig. 3E) capping a basaltic block, and red ribbon chert from higher in that 14 m succession, both have Zone 1I radiolarian assemblages, assigned to the Callovian to lower Oxfordian (Table 1). The radiolarians from another chert remnant nearby span Zones 1F to lower 1I (Middle Jurassic: upper Bathonian to Callovian). LouvionTrellu (1986, cited in Huot and Maury, 2002) also reported Callovian to early Oxfordian radiolarian ages for some of the mlange-block cherts. Based on these data, the radiolarites deposited atop TCSM basaltic oceanic crust range in age from late Middle Jurassic (upper Bathonian and Callovian) to earliest Late Jurassic (lower Oxfordian). The underlying oceanic basalt, therefore, is Middle Jurassic. Younger late Jurassic radiolarian chert occurs interbedded with oceanic basalt, inferred to be volcanic seamounts that grew atop the TCSM oceanic lithosphere. The most complete such remnant is the Stonyford volcanic complex (Shervais et al., in press a, and references cited therein), described in a later section. Another example is the red, non-tuffaceous radiolarian chert interbedded in a thick 3 km long remnant of submarine basalt cropping out 34 km northwest of Wilbur Spring at the southern end of the TCSM belt (Fig. 1). This basalt, formerly mistaken for a Coast Range ophiolite remnant (Bailey et al., 1970; Hopson et al., 1981; McLaughlin et al., 1989), is regarded now as a TCSM mlange megablock within the southern TCSM belt. This reassessment is based on the red, nontuffaceous, Franciscan-like appearence of the chert and its

occurrence within the lava massas at SFVC instead of resting on top. Thus, it is unlike the greenish-gray tuffaceous volcanopelagic (VP) chert strata that overlie the CRO basalts. The abundant, wellpreserved radiolarians belong to Zone 3, Subzone 3, assigned to the uppermost Kimmeridgian to lower Tithonian stages of the Upper Jurassic (Table 1, sample NSF 997A, B, C). The basaltic megablock is faulted against (originally erupted upon?) the TCSM serpentinized peridotite (shown as CRO serpentinite by McLaughlin et al., 1989), and juxtaposed discordantly (faulted?) against Great Valley Group strata. The full span of TCSM radiolarite deposition Middle and Late Jurassiclies within the Early Jurassic through Early Cretaceous age span of Franciscan radiolarian cherts (Murchey, 1984; Sedlock and Isozaki, 1990), allowing their correlation. The TCSM chert only partly overlaps the depositional span of the tuffaceous radiolarian cherts (VP succession) that overlie the Coast Range ophiolite. The VP cherts are chiefly Late Jurassic, i.e., early Oxfordian to late Tithonian (Hull et al., 1993; Pessagno, Hull, Munoz, and Blome in Hopson et al., 1996), although basal chert at two CRO-VP remnants may be late Callovian (Hopson et al., in prep.). In conclusion, the TCSM mlange-block cherts physically match the nontuffaceous Franciscan ribbon cherts and fall within their temporal range. In contrast, they differ physically and are partly older than the tuffaceous radiolarian cherts and mudstones that lie atop the Coast Range ophiolite. Present limited evidence (Table 1) suggests that the TCSB radiolarians were warm-water dwellers from the Central Tethyan (paleoequatorial) Province of the Middle Jurassic ocean, using the criteria of Pessagno and Blome (1986, 1990), Pessagno et al. (1991, 1993, 1999), Hull (1995, 1997), Hull et al., (1997), and Pessagno and Martin (2003). The radiolarian cherts are therefore exotic to their present location near 3940N (within the Southern Boreal Province)i.e., they were transported northward from an original paleoequatorial setting. They may be similar in this regard to the Jurassic Franciscan radiolarian cherts, whose oceanic basaltic substrate evidently originated south of the paleoequator (Alvarez et al., 1980; Murchey and Jones, 1984). Both the TCSM and Franciscan ribbon cherts record slow pelagic deposition on the ocean floor at depths greater than the calcium carbonate compensation depth. The depositional conditions and diagenetic development of ribbon chert from radiolarite at ODP

80

HOPSON AND PESSAGNO

sites 800 and 801 in the equatorial Pacific (Ogg et al., 1992) seem applicable here.

Argillite
Dark slaty argillite, enclosed by serpentinite, is the only other sedimentary rock type present in the Tehama-Colusa serpentinite mlange belt. We have observed these only in the Colusa mlange, where they occur as small slabs (<15 m thick and >30 m long) composed wholly of dark argillite. The lack of silty or sandy interbeds suggests that the original deposit may have been abyssal clay (Kennett, 1982; Berger and Winterer, 1974). Original stratigraphic relations come from a single locality northwest of Stonyford: a huge chamisecovered basaltic block crossed by the road at Black Diamond Glade, just east of the spring (Fig. 3B). Here poorly exposed ribbon chert >10 m thick directly abuts the weathered basalt, presumably deposited on its surface. Bordering the opposite side of the chert is a narrow zone of dark soil with argillite fragments. Evidently argillite overlay the chert that rested on basalt. It appears that radiolarite deposition on basaltic sea floor was followed by abyssal clay, a succession common on the deep-sea floor due to progressive subsidence beneath the silica compensation depth with sea-floor spreading away from a ridge (Berger and Winterer, 1974; Ingersoll, 1988, Fig. 4). Alternatively, the argillite may have been terriginous mud beyond reach of sandy turbidites.

Garnet amphibolite?
Rare blocks of amphibolite [and] garnet amphibolite in the serpentinite mlange are also reported (Jayko and Blake, 1986, p. 1059). No significance is attached to the amphibolite, which is a common foliated variant of uralitic gabbro in oceanic crustal assemblages. But garnet amphibolite is a mafic metamorphic rock that forms at high pressure. The single block of garnet amphibolite known to us in the TCSM belt occurs where Thomes Creek crosses the eastern margin of the serpentinite mlange (Hopson et al., 1981). How the block got there is enigmatic, but it is not visibly embedded in the serpentinite. John Shervais reported another block of garnet amphibolite (eclogite?) on the west flank of Black Diamond Ridge and perhaps others littering the slope nearby (pers. commun., August, 2004). We doubt that the garnet amphibolites found along the TCSM were blocks within the serpentinite

mlange, because of their apparent surficial occurence. Small serpentinite bodies (diapirs?, tectonic slabs?, serpentinitic debris-flow remnants?) within (or lying on?) GVG Upper Jurassic mudstone near the south fork of Elder Creek (Fig. 2) contain small blocks of greenstone, chert, amphibolite, and a garnet amphibolite (Blake et al., 1987). The origin of those serpentinite bodies is speculative, and their postulated connection with the serpentinite mlange belt (Jayko and Blake, 1986)several kilometers away, beyond the intervening PaskentaElder Creek CRO remnantis tenuous. The amphibolite and garnet amphibolite blocks on the TCSM would seem to have some connection with the mid to Late Jurassic (~165140 Ma) amphibolite, garnet amphibolite, and other high-P/T metamorphic rocks (blueschist, eclogite) that occur widely as blocks and slabs in the Franciscan Complex (e.g., Bailey et al., 1964; Coleman and Lanphere, 1971; Suppe and Foland, 1978; Blake et al., 1988; Cowan and Bruhn, 1992; Wakabayashi, 1990, 1999), a topic beyond the scope of this paper. In the context of their rare occurence on the TCSM belt, however, suffice it to say that we believe that these are from CRO oceanic crust metamophosed in the mid to Late Jurassic Great Valley subduction zone (see later section on Mesozoic Tectonic History). Fragments from exhumed slabs of once deeply subducted crust were released into the coarse clastic proxymal facies of uppermost Jurassic and lowermost Cretaceous Great Valley Group. Then, following upheaval beneath the eastern Great Valley zone in the late Early Cretaceous, they were transported westward into the Franciscan depositional realm as submarine debris flows and slide blocks. Such transport westward by subaqueous mass flowage, followed later by erosional removal of their host sediments, left heavy amphibolitic and other high-grade metamorphic lag blocks resting on the surface of the TCSM and adjacent terranes, where some of them still remain. During uplift of the TCSM/CRO forearc ridge in the Early Cretaceous (see later section on Mesozoic Tectonic History) many of the metamorphic lag blocks were carried farther westward into the Franciscan Central belt; others moved eastward a short distance into the new forearc basin, carried by serpentinitic debris flows that interfinger with distal terriginous strata of the Great Valley Group near Wilbur Springs (Carlson, 1981a, 1981b).

TEHAMA-COLUSA SERPENTINITE MLANGE

81

Origin of the Mlange


The protolith of the serpentinite mlange was basaltic oceanic crust overlying upper mantle peridotite and transition-zone dunite. Radiolarian ribbon chert up to 25 m thick lay atop seafloor basalt, and thin argillaceous mudstone or clay locally overlay the chert. Terriginous sandstones and/or volcaniclastic strata were lacking, indicating an open-ocean setting, far from a continent margin or active volcanic arc. This oceanic crust-mantle succession was pervasively disrupted, hydrated, and physically mixed by a large-scale tectonic event beneath the deep ocean floor, forming areally extensive submarine serpentinite-matrix mlange. It escaped the later trench sedimentation, subduction, and high P/T metamorphism that characterize the Franciscan rocks farther west in the Coast Ranges. Insight into the disruption mechanism comes from the following description of basaltic block-serpentinite structural relationships, viewed along the Toomes Camp Road (Rd. M2) west of Crowfoot Point (Paskenta section of Tehama serpentinite mlange): The transition from large intact masses of volcanic rock into serpentinite-matrix mlange shows relationships that reveal the melanging mechanism. Beginning in the volcanic unit, massive or pillowed lavas are locally broken by narrow fault zones that are occupied by sheared, pulverized serpentinite. As the adjacent mlange is approached, the serpentinite fault slices within the volcanics become wider and more abundant, until they gradually isolate masses of volcanic rock between them. This grades into mlange as the proportion of serpentinite (host) to volcanic rock (isolated masses and blocks) increases. Finally, mlange grades into barren serpentinite as the volcanic blocks become fewer and then disppear altogether. Thus, a complete transition occurs, from volcanic rock through serpentinite-matrix mlange to serpentinite. The melanging evidently resulted from (1) faulting and tectonic disruption of the volcanic unit, (2) the intrusion of incompetent, plastically flowing serpentinite into the fault zones and disrupted brittle masses of volcanic rock, and (3) further disruption and mixing by pervasive shearing and plastic flowage. (Hopson et al., 1981, p. 461462)

The timing of serpentinization relative to oceanic crust-mantle disruption and mixing is not well constrained. The outcrop relations described in the quotation above suggest mobility of the ultramafic material at low temperature and shallow depths beneath the sea floor. Thus, the invasive ultramafic material had already been converted to incompetent serpentinite or was being converted to serpentinite. The latter possibility seems more attractive: the pervasive disruption of oceanic crust and underlying peridotite would have facilitated large-scale entry of seawater, launching serpentinization. Thus, we infer that deformation and serpentinization proceeded together. What was the setting and cause of the TehamaColusa oceanic lithosphere disruption? An oceanic fracture zone (Shervais and Kimbrough, 1987) is an attractive possibility. Saleeby (1984), reviewing the occurrence of oceanic serpentinites, noted that The majority of these ocean floor serpentinites were recovered from fracture zones. The relationship to serpentinite mlange is explained: Tectonic disruption of newly formed oceanic crust along transform segments of fracture zones in conjuction with, and followed by, the solid state injection of serpentinitic rocks leads to the development of serpentinitematrix mlange as a primary constituent of oceanic crust. (Saleeby, 1984, p. 155) The process described here corresponds nicely with our interpretation of structural relationships in mlange exposed in traverses through the northern TCSM belt, such as the one west of Crowfoot Point described above. The large scale of oceanic fracture zones also seems adequate to match the TCSM melanging. Quoting again from Saleeby (1984): the total length of fracture zone crust on the ocean floors far exceeds the length of subduction and collision zones, and fracture zones are commonly on the order of 50 km wide This explanation requires that alignment of the FZ coincide approximately with the N-S alignment (present coordinates) of the ~100 km long Tehama-Colusa serpentinite mlange belt (Fig. 1). Close similarities are noted with the Upper Triassic serpentinite mlange that comprises the disrupted oceanic basement of the Rattlesnake Creek terrane in the southwestern Klamath Mountains. Here Wright and Wyld (1994, p. 1033) concluded that:

82

HOPSON AND PESSAGNO

the basement mlange was derived by disruption of oceanic crust and upper mantle in a setting far removed from either an arc or terriginous landmass and are most consistent with an interpretation that mlange formation occurred in an oceanic fracture zone. Their detailed discussion of this and other possibilities provide helpful insights and data sources. Other early Mesozoic and Paleozoic serpentinite mlanges in the western Sierra Nevada (Saleeby, 1979, 1981; Sharp, 1988; Edelman et al., 1989) are complicated by multistage histories and/or by orogenic deformation, metamorphism, and plutonism, veiling their primary origin and setting. Another possible cause of a broad swath of pervasive lithosphere disruption in a deep-sea, openocean setting is a migrating deformation zone (i.e., migrating transform fault zone) between propagating and failing rift tips where ocean-ridge jumping occurred. Here spreading center abandonment, paired with a new spreading center that propagates through slightly older oceanic crust alongside the failing rift, leaves disrupted crust between them. A modern example along the western Galapagos spreading center was described by Kleinrock and Hey (1989). This process could accord with the pervasive disruption and serpentinization of oceanic lithosphere moving progressively through a very large area, like the TCSM. Arguing against this scenario, however, is the thick cover of ribbon chert (up to 25 m), perhaps capped by abyssal clay (i.e., the argillite mlange slabs). Such a sediment cover suggests old ocean crust, perhaps carried by spreading to abyssal depths, away from expected sites of rift propagation. But whatever the cause of TCSM mlanging, it evidently occurred in an open-ocean setting, away from a source of clastic detritus. A different, deep-seated origin of the serpentinite mlange proposed by Jayko et al. (1987) involves its formation in a structural setting below the mafic igneous rocks of the Coast Range ophiolite and above a subducting plate. They explain: We infer that serpentinite mlange forms within a zone below the oceanic crust of the upper plate [i.e., the CRO] and above the subducting oceanic lithosphere. This thick zone of tectonized harzburgite is partially serpentinized by dewatering of the subduction crust (Jayko et al., 1987, p. 481). Further:

It is also an environment in which one might expect to find a lithologic association dominated by the upper part (chert and basalt) and the lower part (harzburgite) of ocean crust without intervening plutonic and cumulate layers. We infer that the alkalic volcanic rocks and chert were derived from the subducting oceanic plate rather than from the upper plate (Coast Range ophiolite) and that the serpentinized harzburgite composing the mlange matrix is primarily derived from the hanging wall of the subduction zone. Volcanic rocks that occur as blocks in the Franciscan Complex tend to be alkalic like the volcanic rocks found in the Round Mountain serpentinite mlange (Bailey et al., 1964; MacPherson, 1983; Jayko, 1984; Shervais and Kimbrough, 1985). We infer that the ophiolite remnant was juxtaposed against the Round Mountain mlange along a detachment fault (the Beehive Flat fault, Fig. 6). (Jayko et al., 1987, p. 482) This attractive hypothesis explains several features of the mlange. It is rejected, however, for the want of an adequate water source (i.e., the basaltchert subducting crust) for large-scale serpentinization of hanging-wall peridotite, and especially for lack of evidence for subduction. The latter includes: (1) the basaltic and chert mlange blocks show no hint of high P/T metamorphism; (2) the serpentinite lacks coarse metamorphic antigorite; and (3) no sandy trench sediments occur as mlange blocks. Instead, the serpentinite is mostly the low-T chrysotile-lizardite-brucite variety; the basaltic blocks show mainly low-grade, subgreenschist-facies alteration; the cherts are unrecrystallized with good preservation of radiolarians; and the absence of terriginous clastic or volcaniclastic sandy mlange blocks all favor development of the mlange by shallow disruption and hydration of oceanic crust and uppermost mantle in a nonsubduction setting. So too does the abundance of abyssal peridotite (cpx pl harzburgite) and enstatite dunite in some serpentinite mlange blocksrocks indicative of uppermost oceanic mantle and crust-mantle transition zone (MTZ), not the deep underside of a mantle wedge. Huot and Maury (2002) followed Jayko et al. (1987) in proposing an origin for the serpentinite mlange involving subduction of Franciscan oceanic lithosphere beneath the CRO, inferred by them to be backarc crust. They too believe that

TEHAMA-COLUSA SERPENTINITE MLANGE

83

melanging occurred in a subduction setting with tectonic accretion of the uppermost part of the downgoing slab to the lowest part of the overriding plate Specifically: We propose, in agreement with Shervais and Kimbrough (1987) and Jayko et al. (1987), that such a mechanism of accretion is responsible for the formation of the Round Mountain serpentinite mlange. The Franciscan supracrustal rocks (represented by the volcano-sedimentary lithologies) would thus have been tectonically underplated beneath the overriding plate (represented by the sheared serpentinite-matrix and the ultramafic blocks). (Huot and Maury, 2002, p. 121) They continue: Accreted crustal slivers [i.e., the chert, basalt, and diabase] are devoid of high-pressure, low-temperature minerals, a feature that suggests a shallow accretion of the units of the Round Mountain serpentinite mlange. (Huot and Maury, 2002, p. 121) But this interpretation remains confronted with formidable problems. One is the source of the vast volumes of water necessary to achieve serpentinization of peridotite throughout the huge TCSM belt (Fig. 1). A subducting TCSM slab capped only by basaltic lava and radiolarian chert (plus minor argillite) is inadequate as a water source, as mentioned above. Also, the style of mlange mixing, described earlier, seems incompatible with tectonic slicing, duplexing, and/or ductile deformation where a downgoing slab and its roof interacted dynamically. The absence of tectonite fabrics in the cherts and basalts is especially telling. Still another problem is the absence of terriginous sediment (graywacke and mudstone) on the downgoing oceanic slab (inferred to be Franciscan by those authors) when mlanging occurred. Edgar H. Bailey, a prime authority on the northern Coast Range Franciscan (e.g., Bailey et al., 1964) forcefully told one of us (CAH, 1977) that more than 90 percent of the Franciscan is graywacke [including metagraywacke + mudstone]. Its huge abundance is borne out by much work in the northern Coast Ranges (e.g., Bailey et al., 1964; Suppe, 1973; Maxwell, 1974; McLaughlin, 1978; Worrall, 1981; Jayko, 1984; McLaughlin and Ohlin, 1984; Blake et al., 1982, 1985, 1988, 1992; McLaughlin

et al., 1989). This terriginous sediment component of the Franciscan, deposited initially as a clastic apron (Blake and Jones, 1974; McLaughlin and Ohlin, 1984) and then as trench sediment along the North American continent margin (e.g., Page, 1981; Dickinson et al., 1982), began in the latest Jurassic (Jones et al., 1969; Blake and Jones, 1974). But, the serpentinite mlanging disrupted Middle Jurassic and earliest Late Jurassic oceanic lithosphere, either prior to burial beneath those terriginous clastics or beyond their reach. We return to the question of where and when mlanging took place in the later section on Mesozoic Tectonic History. In conclusion, the TCSM mlanging is explained here as the widespread tectonic disruption and serpentinization of basaltic oceanic crust and uppermost mantle beneath the sea floor, perhaps along one or more oceanic fracture zones. Serpentinization marks the entry of enormous volumes of seawater during pervasive deformation of the upper mantle peridotite. Swelling, plastically flowing serpentinite invaded, wedged apart, and engulfed blocks of the disrupted crustal rocks, forming mlange. An open ocean setting, away from continent-margin sedimentation or active arc volcanism, is inferred from the absence of terriginous clastic or volcaniclastic sediment cover when crustal disruption took place. Radiolarian ribbon chert, perhaps capped by abyssal clay, was the chief sediment cover involved in the crustal disruption. Also involved locally was a long-lived submarine volcano (Jurassic seamount) growing atop the basaltic oceanic crust during radiolarite deposition, considered next.

Stonyford Volcanic Complex


An important feature closely associated with the TCSM belt occurs near Stonyford, where the serpentinite mlange wraps around a large mass of intact submarine lavas, hyaloclastites, and minor interstratified radiolarian cherts. This assemblage, cropping out as an ovoid mass approximately 8 5 km in plan [with] estimated stratigraphic thickness at least 1 km and possibly as much as 2.5 km (Brown, 1964a), was once called the Stonyford seamount (Hopson et al., 1981; Shervais and Kimbrough, 1985a) and the Stonyford volcanics (MacPherson and Phipps, 1985), but was later renamed the Stonyford volcanic complex (Shervais and Hanan, 1989). Its chief features were summarized by Zoglman and Shervais (1991, p. p. A395):

84

HOPSON AND PESSAGNO

The Stonyford volcanic complex (SFVC) is a thick sequence of submarine volcanic rocks structurally juxtaposed between high P/T metamorphic rocks of the Franciscan assemblage and unmetamorphosed sediments of the Great Valley sequence, in the northern Coast Ranges of California. The SFVC consists primarily of pillow lava, with subordinate massive lava flows, diabase dikes, fragmental eruptive breccias, and hyaloclastite breccias. The hyaloclastite breccias, interpreted as submarine fire fountain deposits, contain unaltered volcanic glass which has been dated by 39Ar/40Ar at 164 1 m.y. (Hanan et al., 1991). Radiolarian chert and minor limestone are found intercalated with pillow lava throughout the sequence; chert forms bedded sequences up to 10 m thick as well as interpillow fillings. The lavas and hyaloclastites include OIB tholeiites, low-Ti basalts, andhigh in the pilealkali basalts (Shervais, 1993). A small brecciated microplagiogranite body crops out at the northern margin of the volcanic pile, evidently an SFVC silicic intrusive sheet (Hopson observation, June 2004). Further details of the geology, geochemistry, Pb isotope composition, and age of the SFVC are given by Shervais and Kimbrough (1987), Shervais and Hanan (1989), Hanan et al. (1991), Zoglman and Shervais (1991), and Shervais et al. (2002, and in press, b). The SFVC is regarded as a relict seamount (Shervais and Kimbrough, 1987; Zoglman and Shervais, 1991; Shervais, 1993; Shervais et al., in press, b; also Hopson et al., 1981). The thickness of this local submarine lava pile, the chemistry of its lavas, and the absence of cogenetic plutonic rocks beneath it all argue for its origin as a seamount, built on proto-TCSM oceanic crust and mantle. Three lines of evidence connect the SFVC with oceanic crustal rocks (now mlange blocks) of the serpentinite-matrix mlange: (1) the partly matching immobile trace-element chemistry of their basaltic lavas (Shervais and Kimbrough, 1985a, 1987); (2) their physically matching radiolarian ribbon cherts; and (3) the approximate corresponding age of those cherts. Typical SFVC chert has been described as: Reddish-brown manganiferous (nontuffaceous) radiolarian ribbon chert [that] is locally interbedded within the pillow lavas. Along Stony Creek near the campground

(Sec. 35) chert units 15 and 30 m thick are interstratified within the basalt. (Hopson et al., 1981, p. 457). And also: Chert forms a prominent horizon about 10 m thick intercalated with basalt in the western part of the complex; lenses of Mn-rich umber up to 2 m thick grade laterally into red jasperoid breccias. Radiolaria are common in the upper 45 m of the banded ribbon chert that overlies the umbers and jasperoid breccias. (Shervais and Hanan, 1989, p. 510511) The radiolarian cherts within the SFVC seamount and those occurring as mlange blocks in the Tehama-Colusa serpentite mlange belt are twins, physically. Both are nontuffaceous ribbon cherts, typically in dark red-brown hues but dark grey locally. Both are indistinguishable physically from the Franciscan radiolarian ribbon cherts (Bailey et al., 1964, Figs. 26, 28, and 29). The age of the TCSM oceanic crustal cherts and the lenses of similar chert occuring low stratigraphically in the SFVC appear to span about the same range, whereas the higher SFVC cherts are somewhat younger. Specifically, the ages of TCSM mlange-block ribbon cherts measured thus far (Table 1) range from Middle Jurassic (upper Bathonian through Callovian) to early Late Jurassic (lower to possibly middle Oxfordian), using the radiolarian zonal system of Pessagno et al. (1987, 1993). By comparison, radiolarians from thick chert lenses intercalated with basalt in the SFVC range from Middle Jurassic (Bathonian, corresponding to Unitary Association Zone 6-6 of Baumgartner et al., 1995) near the base of the complex to Late Callovian to early Kimmeridgian (UAZ 8-10) in the upper part (B.L. Murchey, in Shervais et al., in press, b). Comparison of the TCSM chert ages in Table 1 with the SFVC chert ages listed in Shervais et al. (in press, b) is not precise, since the former correspond to the radiolarian zonation scheme of Pessagno et al. (1987, 1993) and the latter to the UAZ zonal system of Baumgartner et al. (1995), which assign older ages for comparable taxa. The differences, discussed by Pessagno and Hull (1996), Hull et al., (1997), Pessagno et al. (1999, 2003), and by Murchey (1997, and in Shervais et al., in press, b), are beyond the scope of this paper. Despite unresolved chert age differences it seems clear that the Stonyford submarine volcano (SFVC) experienced its early growth (Middle Juras-

TEHAMA-COLUSA SERPENTINITE MLANGE

85

sic) upon TCSM basaltic ocean crust during radiolarite deposition on that surface, and that its growth lasted well into the Late Jurassic. A corresponding TCSM basaltic megamlange block with interbedded Upper Jurassic red radiolarian chert (Table 1) occurs ~25 km farther south, near Wilbur Springs (Fig. 1). This too is probably the remnant of a seamount that grew upon TCSM oceanic lithosphere. Why then did not TCSM radiolarite deposition on the deep-ocean floor also continue into the mid-Late Jurassic? Perhaps it did: the uppermost TCSM cherts above the basalt have not yet been identified within the mlange. Or, speculatively, seafloor spreading may have carried the TCSM ocean crust to depths below the silica compensation depth (SCD)i.e., where radiolarite deposition ceased and only abyssal clay slowly accumulatedwhereas the seamounts rose above the SCD. The thin slabs of argillite in the Colusa serpentinite mlange could be remnants of that abyssal clay. Paleobiogeographically, the TCSM sea-floor cherts studied by Pessagno are Central Tethyan (Table 1), indicating radiolarian growth in paleoequatorial warm waters. In the Stonyford seamount, B. L. Murchey (in Shervais et al., in press, b) notes that A major faunal change occurs within the Locality B section, wherein the relatively smallsized, polytaxitic radiolarian faunas in the lower part of the section give way to very robust, oligotaxitic nassilarian-dominated faunas that include Praeparvicingula spp. A similar change occurs in the upper volcanipelagic succession atop the Josephine ophiolite in the Smith River section (western Klamath Mountains), where Pessagno et al. (1993) noted the disappearence upsection of pantanellids and influx of Praeparvicingula, inferred to indicate transport from Central Tethyan to Southern Boreal paleolatitudes during Jurassic sea-floor spreading (Pessagno and Blome, 1986; Pessagno et al., 1993, 2000). The similar faunal succession at Stonyford (reported in Shervais et al., in press, b) therefore suggests that initial growth of the long-lived submarine volcano began in warm paleoequatorial watersas did the TCSM oceanic crustand completed its growth farther north in cooler, higher-latitude waters. This holds also for the Wilbur Springs seamount(?) basalt that contains Upper Jurassic red chert with a Northern Tethyan fauna (Table 1). These observations favor a model in which during the late Middle Jurassicthe Stonyford volcanic complex was growing atop basaltic oceanic

crust and peridotite upper mantle during the period of radiolarite deposition. BathonianCallovian radiolarite (transformed authigenically to ribbon chert as described by Ogg et al., 1992) was accumulating as a thin blanket atop the basaltic sea floor, andat about the same timealso as intercalations within the growing SFVC basaltic seamount. Both were affected by the subsequent widespread tectonic event, but to different degrees. The Tehama-Colusa belt of oceanic crust and mantle was pervasively disrupted and hydrated, forming serpentinite-matrix mlange. The overlying thick pile of seamount lavas was doubtless also deformed but not mlanged, being separated from the underlying mobile serpentinite. Faulting and other internal deformation of the Stonyford seamount doubtless occurred, e.g., note the brecciation of the plagiogranite body mentioned above, and especially the invasion of basaltic SFVC lavas by the underlying serpentinite, mapped by Shervais et al. (in press a, b). But, most of the Stonyford seamount survived oceanic deformation, remaining largely intact (Shervais et al., in press, b and references cited therein). In conclusion, we concur with Shervais and Kimbrough (1985b, 1987) that the serpentinite mlange terrane is a Franciscan feature, and also with Shervais and others (Shervais and Kimbrough, 1985a, 1987; Shervais and Hanan, 1989) that the SFVC was a Franciscan Jurassic seamount. This holds also for the basaltic megablock near Wilbur Springs (Fig. 1) with Franciscan-like radiolarian chert. We reject the more recent interpretation of the SFVC as part of the Coast Range ophiolite (Shervais et al., 2002, and in press, b), which is strikingly different lithologically, pseudostratigraphically (Hopson et al., 1981, and in prep.), geochemically (Shervais and Kimbrough, 1985a, 1985b), and the overlying tuffaceous radiolarian cherts (volcanopelagic succession) of which are younger (mostly Late Jurassic) than the TCSM cherts and reflect a quite different depositional setting (Hopson et al., 1996, in prep.; Pessagno et al., 2000). The proto-TCSM oceanic mantle and basaltic crust (Bathonian or perhaps older) overlaps and partly predates the CRO remnants (161168 Ma; J. M. Mattinson, pers. commun., 2004), it lacks the abundant plutonic rocks, and key trace elements in its lava are different. Moreover, the major tectonic event that mlanged the TCSM belt left the CRO uneffected. The Stonyford seamount evidently began its growth at more or less the same time as the CRO oceanic crust (late Middle Jurassic; Hopson et al., in

86

HOPSON AND PESSAGNO

prep.) but in a different, distant tectonic setting, as reflected by their contrasting radiolarian cherts and other differences. Plate tectonic movements then brought them together along the Stony CreekBeehive Flat fault system, a topic addressed next.

Mesozoic Tectonic History: Tehama-Colusa Serpentinite Mlange and Adjacent Terranes


The importance of the Tehama-Colusa serpentinite mlange (TCSM), tectonically, is the role that it and the adjacent Coast Range ophiolite (CRO) played in development of the forearc ridge that separated the inboard Great Valley forearc basin from the outboard Franciscan subduction/accretion complex during the Cretaceous. This section deals with the Jurassic prehistory of the forearc ridge, its abrupt rise in the Early Cretaceous, and related subsequent Cretaceous events. The TCSM and CRO are juxtaposed along the Beehive Flat fault, a northern continuation of the Stony Creek fault (Fig. 2). This subvertical fault system separates the TCSM on the west from the CRO and conformably overlying Upper Jurassic basal strata of the Great Valley Group (GVG) on the east. The Stony Creek fault s.s., extending from Wilbur Springs north to Mill Creek (west of Paskenta), separates the serpentinite mlange from GVG Upper Jurassic strata (Fig. 1), presumably underlain by unexposed CRO. The small, NWtrending Mill Creek fault west of Paskenta dextrally offsets the Stony Creek fault and lifts up the rocks on the east side, exposing CRO (its Digger Creek Elder Creek remnant) beneath the GVG strata. Beyond this small offset the Stony Creek fault continues as the Beehive Flat fault (Fig. 2), which places the serpentinite mlange visibly against CRO and diagonally truncates its map units. The Beehive Flat fault is cut off by the late Cretaceous Coast Range fault ~13 km farther north (Jayko and Blake, 1986; Blake et al., 1992). Brown (1964a, 1964b), working in the Stonyford area, described the Stony Creek fault (Fig. 1) as a thrust fault. He interpreted the Stonyford volcanic complex (SFVC) as a klippe above a low-dipping thrust fault that flattens westward from the otherwise steep Stony Creek fault (Brown, 1964b, Fig. 123.2). Shervais and Schuman mapped the SFVC basal contact as complexly faulted but not as a low-angle thrust (Shervais et al., in press, a, b). For most of its length, however, we accept Browns interpretation of

the Stony Creek fault as a steep reverse fault that brings up the rocks on the east. McLaughlin et. al. (1984, 1989) also showed the Stony Creek fault as a thrust fault in the Wilbur Springs area. It seems likely, however, that the original boundary between the TCSM and CRO terranesnow the Stony CreekBeehive Flat fault systemwas a relict transform fault with very large strike-slip displacement. The evidence comes from the contrasting Jurassic tectonic histories of the TCSM and the CRO, the two Jurassic oceanic terranes that it separates. Neither terrane was ever deeply buried, and no amount of dip-slip displacement can account for their different Jurassic tectonic histories. Detailed comparison of their contrasting Jurassic histories is beyond the scope of this paper, but a brief summary of some important differences is offered. The Coast Range ophiolite (CRO)1 is a composite body encompassing remnants of Middle Jurassic oceanic lithosphere, an Upper Jurassic abyssal volcanopelagic sediment cover, an ophiolitic breccia (OB) at northern CRO remnants, Late Jurassic intrusive sheets, and an ophiolite remnant that invades and displaces the VP and underlying CRO. The mid-Jurassic ophiolite (now dismembered) originally comprised an axial crust-mantle sequence (tholeiitic submarine lava, sheeted diabase, diorite/ gabbro, cumulus gabbro, and impregnated harzburgite tectonite members), and a coeval offaxis sequence of upper submarine lava (from basaltic/ankaramitic small volcanoes), a dunite-wehrliteclinopyroxenite intrusive complex (Moho Transition Zone, MTZ), and a wehrlitic-gabbroic-noritic dike swarmrooted in the MTZthat fed the upper lava (Hopson, 2002; Hopson et al., in prep.). This midJurassic ophiolite with well-preserved primary structures and textures, and with minimal internal deformation, implies a robust thermal budget and thus a fast spreading center, similar to the modern East Pacific Rise (Dilek et al., 1998). In the Late Jurassic (Kimeridgian/Tithonian), an unstable oceanic ridge system (with ridge-jumping) propagated through the mid-Jurassic oceanic lithosphere, sending intrusive sheets into the still-accumulating
1Many of our conclusions concerning the CRO come from a large manuscript entitled California Coast Range ophiolite: Composite Middle and Late Jurassic oceanic lithosphere, whose preparation is nearly complete. This manuscript, on which we draw heavily and believe to be the best current source of information on the ophiolite, is cited here and elsewhere in this paper as Hopson et al. (in prep.).

TEHAMA-COLUSA SERPENTINITE MLANGE

87

VP soft sediments, locally displaced them with newly formed mafic crust, and generated fragmental detritus (ophiolitic breccia and mafic clastic sediment sheets) at northern Coast Range CRO localities (Hopson et al., in prep.). In contrast, the Tehama-Colusa serpentinite mlange belt comprised only thin basaltic crust composed largely of submarine lavas with sparse diabase and only rare plutonic remnants (diorite, gabbro, pyroxenite/wehrlite). That succession reflects a low thermal budget with conspicuous deformation of primary features (Dilek et al., 1998; Thy and Dilek, 2000; Robinson et al., 2000; also Sinton and Detrick, 1992), perhaps including ridgeaxis core complexing that brought up serpentinized mantle peridotite while stripping off basaltic crust, analogous to the slow-spreading Mid-Atlantic Ridge (Cannat, 1993; Blackman et al., 1998; Cann et al., 1998). The Jurassic CRO and TCSM oceanic lithosphere formed at about the same timelate Middle Jurassic and Late Jurassicat paleoequatorial midocean ridges (Hopson et al., 1996, in prep.), but those ridges were widely separated, with contrasting thermal budgets and spreading rates. Other important differences between the CRO and TCSM oceanic lithosphere involve the contrasts in their Jurassic sedimentary histories. The CRO-VP plate stratigraphy reveals an upward succession from intralava pelagic limestone within the midJurassic ophiolite, to a CRO-VP disconformity (39 m.y. hiatus), to a thin succession of tuffaceous radiolarian chert (VP distal facies), to a thicker assemblage (at only two of 22 CRO remnants) of radiolarian-bearing tuff, volcaniclastic turbidites, and submarine volcanic debris-flow deposits (VP proxymal facies). This upward succession, spanning 1822 m.y., is incompatible with an arc-related setting in which the oceanic basement remains close to the arc. Instead, this stratigraphy reflects CRO plate formation in an open-ocean setting far from an arc, followed by a progressive approach to an active arc through zones of sub-CCD calcareous sediment starvation, then radiolarian ooze/fallout-tuff deposition (distal facies), and finally proxymal-facies deposition (forearc volcaniclastic apron lobes, reaching two CRO segments), together spanning most of the Late Jurassic (Hopson et al., 1996, in prep.; Pessagno et al., 2000). The CRO oceanic lithosphere was being drawn progressively closer to an east-dipping subduction zone in front of the mid- to late Jurassic (Rogue-Logtown Ridge) continentfringing intraoceanic arc. VP sedimentation

declined and was overwhelmed by terriginous muds and distal turbidites (basal Great Valley Group) in the latest Jurassic (Hopson et al., 1981, 1996), following the inboard Nevadan orogeny (Harper et al., 1994; Sharp, 1988) and erosion of the new contractional mountain belt. Mafic sands, grits, and gravels spread locally around high-standing sea-floor remnants of the CRO ophiolitic breccia unit, interfingering with Late Jurassic GVG terriginous sediments from the distant Nevadan orogen. The results of paleomagnetic and biostratigraphic studies document CRO formation within a few degrees of the mid-Jurassic paleoequator (Luyendyk and Hornafias, 1982; Beebe and Luyendyk, 1983; Beebe, 1986) and a progression from paleoequatorial warm-water to higher-latitude coolerwater radiolarian faunas during VP sedimentation (Pessagno et al., 1993, 1999, 2000; Hull, 1995, 1997; Hopson et al., 1996; Hull et al., 1997; Pessagno and Martin, 2003). Together these record a northward component of CRO plate motion during the Late Jurassic. The quite different interpretion of CRO petrogenesis and Jurassic history by Shervais et al (2004) is recognized but not accepted (Hopson et al., in prep.). In contrast, the TCSM oceanic lithosphere remained in an abyssal deep-sea environment in which radiolarian ooze and clay accumulated below the calcium carbonate compensation depth (i.e., forming radiolarite that changed diagenetically to ribbon chert) mainly during the late Middle Jurassic and early Late Jurassic (Table 1). Greater thickness of the Callovianupper Oxfordian ribbon cherts may reflect a higher accumulation rate of radiolarian detritus while seafloor spreading carried the basaltic substrate across the equatorial zone of pelagic high productivity (Berger and Winterer, 1974; Kennett, 1982; Ingersoll, 1988). The TCSM ribbon cherts, unlike the VP succession atop the CRO, are largely Franciscan-like ribbon cherts devoid of volcaniclastic components such as tuff (from airborne dust and ash), shards, and calc-alkaline volcanic microlites (Huot and Maury, 2002 note a single exception). A paleoequatorial origin for the TCSM cherts, like that of Early Jurassic Franciscan radiolarian cherts (Murchey, 1984; Murchey and Jones, 1984), seems likely in view of their Central Tethyan radiolaria (Table 1). Differences between the VP tuffaceous cherts (atop CRO) and the nontuffaceous TCSM (and Franciscan) ribbon cherts reflects their different Jurassic histories. Although both have paleoequatorial

88

HOPSON AND PESSAGNO

(Central Tethyan) radiolarian faunas in their lower portions, their transport northward followed different paths or had different timing. The VP tuffaceous cherts accumulated while being drawn obliquely north-northeastward toward the subduction zone in front of the active Jurassic arcfollowing the fallout zone of the volcanic dust and ash carried south and southwest by the trade winds (Hopson et al., 1996, Fig. 3). Thus, radiolarian ooze was mixed with airborne tephra that settled to the abyssal ocean floor throughout most of the Late Jurassic, producing the VP tuffaceous radiolarian mudstones and cherts (Hopson et al., 1996, in prep.; Pessagno et al., 2000). In contrast, the TCSM (and Franciscan) radiolarites accumulated either outside the Late Jurassic tephra fallout zone ormore likelyafter arc volcanism diminished and ended during a late stage of the Nevadan orogeny (Sharp, 1988; Harper et al., 1994). The latter possibility seems more likely inasmuch as the VP and TCSM cherts probably followed similar paths to end up alongside each other (Figs. 1 and 2). The tectonic significance and timing of the ancestral Stony CreekBeehive Flat fault system is inferred as follows. The plate boundary between western North America (at the California paleolatitude) and the offshore oceanic plate (CRO), prior to the Nevadan orogeny, had been the Great Valley subduction zone (Fig. 4A) with northeastward dextral oblique subduction of the oceanic plate moving up from the southwest (Hopson et al., 1996, in prep.). When subduction ended during the Nevadan orogeny the plate boundary changed to a dextral transform fault (Fig. 4B), which continued to accommodate the oceanic plates northward component of motion. The transform boundary may have initially been farther east within the CRO plate (i.e., a segment now hidden beneath the Great Central Valley) but perhaps shifted position with timesimilar to strike-slip faults within the Cenozoic San Andreas fault systemlater bringing up more southerly oceanic lithosphere whose remnants now comprise the Franciscan peridotites, submarine lavas, and radiolarian ribbon cherts including the TCSM terrane (Fig. 4B). Thus, the ancestral Stony CreekBeehive Flat fault system, separating the TCSM and CRO oceanic lithosphere, was probably part of the postNevadan transform fault system. Portions of this reconstruction (Fig. 4A) disagree in part with recent interpretations of lithology and structure of basement rocks hidden beneath Cretaceous and Cenozoic deposits of the Sacramento

Valley (e.g., Godfrey and Klemperor, 1998; Godfrey and Dilek, 2000; Coleman, 2000). The geophysical constraints are important but can fit different geologic reconstructions beneath the Great Valley, e.g., compare Coleman (2000) with Godfrey and Dilek (2000). Reconciliation of the geophysical data with our model (Fig. 4) is possible, we believe, but beyond the scope of this article. In summary, the large differences between the TCSM and CRO oceanic crustal rocks, and between TCSM and VP (atop CRO) radiolarian cherts, probably signify different original settings and different Jurassic travel histories. Lacking evidence of subduction between them, it is inferred that large-scale strike-slip displacement along an oceanic transform fault probably brought the two terranes together. Murchey and Blake (1993) also inferred a transform boundary between the oceanic and North American plates here during the latest Jurassic and earliest Cretaceous. But, the Stony CreekBeehive Flat fault system that now separates the TCSM and CRO (Fig. 2) entails a further history that involved thrust faulting and also truncation of uppermost Jurassiclowermost Cretaceous (TithonianValanginian) terriginous mudstones and turbidites that cover the oceanic CRO/VP/OB succession. The next stage of our historical reconstruction explains these and other facets of CRO/VP/OB/basal GVG and Franciscan tectonics. The unsubducted portion of the Jurassic CRO/ VP/OB oceanic plate stopped short of the Great Valley trench and east-dipping subduction zone (Fig. 4A), which became inactive during the Late Jurassic Nevadan orogeny. The newly inactive trench was filled by the growing apron of uppermost Jurassic (Late Tithonian) volcaniclastic turbidites and submarine debris flows (VP proxymal facies), which locally bridged the newly filled trench at the Llanada and Del Puerto Canyon CRO/VP localities (Hopson et al., 1996, in prep.). The rising, newly mountainous Nevadan orogen (Harper et al., 1994, Fig. 16) then shed clastic sediments, i.e., the uppermost Jurassic and lowermost Cretaceous (late TithonianValanginian) continental slope and basinplain submarine fan systems of terriginous muds and turbidites that now comprise the lower GVG (e.g., Page, 1966; Dickinson and Rich, 1972; Moxon, 1988, 1990; Constenius et al., 2000). The sandy proximal portion of the basal GVG sediment sheet also contained blocks of amphibolite, eclogite, and antigorite serpentinite derived from newly exhumed

TEHAMA-COLUSA SERPENTINITE MLANGE

89

portions of the Jurassic subduction slab (Fig. 4B). The finer distal sediments advanced out over the adjacent deep ocean floor (i.e., the stalled CRO plate with VP/OB sediment cover) andstill farther westover the Jurassic Franciscan oceanic plate (Fig. 4B), which was moving up from the south along the ancestral Stony CreekBeehive Flat transform fault. The Tithonian-Valanginian distal terriginous strata within the eastern Franciscan belt are widespread and well documented (Blake and Jones, 1974; Blake et al., 1982; McLaughlin and Ohlin, 1984; McLaughlin et al., 1989). They were derived from the western Klamath Mountains and western Sierra Nevada terranes (Dickinson et al., 1982). Those clastic strata must have also covered the TCSM belt that was already in place between the Jurassic Franciscan and CRO oceanic teranes (Fig. 4B). This puts a mid Late Jurassic upper age limit on the TCSM oceanic mlanging, since the basal GVG strata (upper Kimmeridgian or lower Tithonian near Paskenta) are not part of the mlange. Thus, the serpentinite mlanging took place in the Late Jurassic after TCSM chert lithification but before latest Jurassic terriginous clastic sedimentation. But those clastic strata were deformed and mlanged when eastward subduction, beginning in the late Valanginian (Blake and Jones, 1974), replaced the dextral transform faulting (including the ancestral Stony CreekBeehive Flat fault system) that prevailed from late Tithonian into Valanginian time. The Early Cretaceous eastward subduction (Fig. 4C) created the eastern Franciscan collage of underthrust strata and mlange units of TithonianValanginian depositional age (Maxwell, 1974; Blake and Jones, 1974; Worrall, 1981; Blake et al., 1982; Jayko, 1984; McLaughlin and Ohlin, 1984; McLaughlin et al., 1989) and the early Cretaceous high P/T metamorphic assemblages (e.g., Ernst, 1970, 1971, 1977, 1984; Suppe, 1973; Jayko et al., 1986; Blake et al., 1988). Also, compressional arching inboard of the newly developing Franciscan subduction zone steepened the westward paleoslope, launching submarine slides and debris flows from the GVG proximal apron that carried coarse sand and metamorphic blocks westward over the GVG distal strata and the CRO, TCSM, and Franciscan oceanic basements (Fig. 4C). The start of Franciscan subduction in the Early Cretaceous, from a new trench lying farther west, also uplifted and tilted eastward the TCSM and CRO/VP/OB/basal GVG terranes as newly

subducted rocks were wedged beneath them. Tilting and uplift created the forearc basement ridge that bounded a new, post-Valanginian, Great Valley forearc basin. The Stony CreekBeehive Flat thrust fault evidently formed during the eastward tilting (with progressive steepening) of the TCSM and CRO basement units (Fig. 4C), cutting across and displacing the earlier transform fault between them. GVG sediments that initially covered the TCSM basement were also removed (by submarine slides?) at that time. Remnants of incompletely removed GVG sediments still lying on the TCSM include: (1) the volcanogenic sandstones south of Stonyford (mapped as blocks within serpentinite mlange by Shervais et al., in press, a, b); and (2) Napa County localities (south of Fig. 1) where submarine fan rocks of the Great Valley sequence [renamed Great Valley Group by Ingersoll, 1990] directly overlie serpentinized harzburgite tectonite (Phipps, 1984, p. 104), evidently a southern extension of the TCSM (see earlier section on Tehama-Colusa Serpentinite Mlange). Farther west, continued Cretaceous (postValanginian) trench sedimentation (Dickinson et al., 1982), subduction accretion, and subduction-zone metamorphism (e.g., Blake et al., 1982, 1985, 1988; Ernst, 1984; McLaughlin and Ohlin, 1984; Jayko, 1984; Jayko et al., 1986; McLaughlin et al., 1989; Cowan and Bruhn, 1992) built and assembled the eastern Franciscan terranes. It also seems likely that the detrital serpentinites within Lower Cretaceous GVG strata in the Wilbur Springs area (Moiseyev, 1966, 1970; Carlson, 1981a, 1981b; McLaughlin and Ohlin, 1984; McLaughlin et al., 1989) were shed from the new TCSM-CRO basement ridge during its tilting and uplift in the late Valanginian. Serpentinite turbidites and submarine debris flows from the TCSM belt, including sparse amphibolite and blueschist erosional lag blocks from its surface (Fig. 4C), were shed into the adjacent forearc basin and interstratified with the early Cretaceous terriginous muds and sands of the Great Valley Group. The timing in the early Cretaceous is right, and this tectonic scenario seems more credible than the subsea serpentinitediapir hypothesis (Carlson, 1981a, 1981b) in which the high-grade metamorphic blocks were brought up from deep levels of an underlying Franciscan subduction zone. The associated Grizzly Creek Mlange of the Wilbur Springs area, comprising mixtures of TithonianValanginian sediments and volcanic/plutonic blocks from the Coast Range ophiolite (McLoughlin and Ohlin, 1984;

90

HOPSON AND PESSAGNO

TEHAMA-COLUSA SERPENTINITE MLANGE

91

FIG. 4 (facing page). Diagrammatic depiction of five stages in the tectonic evolution of the Tehama-Colusa serpentinite mlange and adjacent terranes, during the late Mesozoic. A. Late Jurassic: CRO oceanic lithosphere subducted obliquely beneath the oceanic Nevadan (RogueLogtown Ridge) volcanic arc that fringed western California. Arc volcanism shed volcaniclastic detritus into the trench, and downwind fallout dust and ash mixed with radiolarian ooze (VP distal facies) on the CRO deep ocean floor. B. Latest Jurassicearliest Cretaceous (TithonianValanginian) exhumation of Jurassic subduction slab shed mafic metamorphic blocks (amphibolite, eclogite) into the adjacent apron of coarse clastic sediment (proximal GVG). Also, the Jurassic Franciscan oceanic plate including its disrupted segment (TCSM) moved northward and was faulted against the CRO/VP oceanic plate remnant along a dextral transform fault (ancestral Stony Creek fault, SCF1). Distal terriginous clastic marine sediment (lower Great Valley Group) from the uplifted Nevadan orogen covered deeply submerged Franciscan and CRO oceanic lithosphere. C. Subduction resumed in the Early Cretaceous (Valanginian) outboard of the earlier, now inactive Jurassic subduction zone. Associated compressional arching between those zones steepened the westward paleoslope, launching submarine slides and debris flows from the GVG proximal apron that carried coarse sand and metamorphic blocks westward across GVG distal strata and their oceanic basements. D. With continued Franciscan subduction, underthrust oceanic crust and disrupted, scraped-off sediment wedges lifted up the TCSM, CRO, and VP/basal GVG strata and tilted them progressively eastward, creating a new forearc ridge. The Stony Creek thrusting of CRO/GVG over TCSM (offsetting the earlier transform) occurred during tilting. Cretaceous Franciscan subduction, accretion, and high-P/T subduction metamorphism proceeded outboard of the ridge, and forearc sedimentation (Cretaceous GVG) took place behind it. E. Late Cretaceous extensional uplift at the rear of the Franciscan accretionary wedge brought up subducted Early Cretaceous metamorphic rocks along the Coast Range fault, placing them against the E-tilted remnant of unsubducted Franciscan Jurassic oceanic lithosphere (TCSM). Forearc basin sedimentation (Upper Cretaceous GVG) continued behind the TCSM-CRO forearc ridge.

McLaughlin et al., 1989), plus serpentinite of possible TCSM derivation, may also be related to the TCSM/CRO tilting and uplift in the Early Cretaceous. Indeed, the descriptions of the Grizzley Creek mlange strikingly resemble those of the serpentinous chaotic unit between basement serpentinite (probably TCSM) and basal GVG strata that crop out extensively farther south (beyond Fig. 1), in Napa County. The ophiolitic olistostromes in the basal Great Valley Group at Napa County localities are described by Phipps (1984, p. 120) as an ophiolitic chaotic unit up to 1 km thick [that] lies above the Coast Range serpentinite [TCSM] and below the bedded, arc-derived sub-sea fan rocks of the Great Valley sequence [GVG]. Phipps (1984, p. p. 103) convincingly demonstrates that the chaotic unit in Napa County is not a tectonic mlange [but] . . . an amalgam of olistostromes. Descriptively, the unit is characterized by ophiolitic blocks in a mudstone, serpentinous mudstone, and serpentinite matrix (p. 120); the blocks include bedded cherts, amorphous unbedded chert; pillow basalts and other greenstones; unsorted aphanitic and diabasic mafic breccias; rare gabbro and plagiogranite breccias and massive gabbro; metamorphic rocks; and serpentinite (p. 120). This block assemblage can readily be construed as a mixture of serpentinitic debris (and basalt-chert mlange blocks?) from the TCSM, plus blocks of disrupted CRO including its capping ophiolitic (mainly basaltic-diabasic) breccia unit,

plus basal GVG mudstone and sandstone, plus Jurassic metamorphic (including amphibolite) lag blocks from the TCSM-CRO surface, all of which slid together from a rising submarine ridge of TCSM-CRO-basal GVG rocks into the adjacent new early Cretaceous forearc basin. Still to be resolved, however, is the age of this large-scale tectonic event in Napa County, i.e., whether it was Tithonian (Phipps, 1984), or early Cretaceous (Valanginian), with Tithonian fossiliferous sediments admixed within and locally carried atop Cretaceous submarine slide masses. Extensional exhumation of the eastern Franciscan Complex during the late Cretaceous (Jayko et al., 1987; Harms et al., 1992) brought up the deep metamorphic rocksearly Cretaceous pelitic schist and metagraywacke (from Tithonian-Valanginian protoliths) and metabasalt/metachert (from Jurassic protoliths)along the Coast Range fault (Fig. 4D). That left the east-tilted, unsubducted strip of Franciscan-like TCSM oceanic lithosphere isolated between the extensional Coast Range fault and the Stony CreekBeehive Flat reverse fault. Ring and Brandon (1994) propose an alterative interpretation in which the Coast Range fault is not extensional but instead consists of low-dipping out-of-sequence thrust faults, signifying crustal shortening. But their analysis is based on the incorrect assumption that the Coast Range fault spans the entire width of the serpentinite belt (TCSM), so that it juxtaposes Franciscan Complex in the footwall

92

HOPSON AND PESSAGNO

against a forearc massif in the hanging wall, which comprises Coast Range ophiolite and the Great Valley forearc basin. Ring and Brandon were evidently unaware (i.e., no references cited) that this 34 km wide beltwhich they describe as a fault zone decorated by serpentinite-rich shear zonesis instead a much older oceanic serpentinite mlange. Because of this and other untenable assumptions, we discount their conclusions. The extensional uplift of deeply subducted eastern Franciscan metamorphic rocks (Platt, 1986; Jayko et al., 1987; Harms et al., 1992) remains the most viable explanation of the Coast Range fault where it borders the TCSM. In conclusion, the Stony CreekBeehive Flat fault marks a composite fault system: an Early Cretaceous steep reverse fault (labeled SCF2 on Fig. 4C) that now separates the TCSM and CRO-VP-GVG terranes, following and offsetting an ancestral latest Jurassicearliest Cretaceous transform fault (SCF1 on Fig. 4B) between two broadly contemporaneous but different Jurassic oceanic terranes (plates). The Stony CreekBeehive Flat reverse faulting accomodated eastward tilting of the stranded TCSM and CRO/basal GVG terranes during the inception of outboard Franciscan subduction in the Early Cretaceous (Valanginian). Earlier proposals that eastward tilting of the CRO/GVG strata formed in response to later Cretaceous and/or Cenozoic eastward wedging of the Franciscan Complex (Wentworth et al., 1984; Unruh et al., 1991, 1995; Jachens et al., 1995; Wakabayashi and Unruh, 1995; Godfrey et al., 1997) has recently been repudiated by Constenius et al. (2000). We concur with the latter authors, and propose that eastward tilting of the TCSM, CRO/VP/ OB, and basal GVGcreating a forearc ridge with a new forearc basin behind itwas caused by the onset of eastern Franciscan outboard subduction in the early Cretaceous. The Stony CreekBeehive Flat composite fault system is much modified by post-Early Cretaceous deformations. The Elder Creek to Wilbur Springs segment of that fault is locally warped (Figs. 1, 2), especially near Chrome and Mill Creek. It is truncated north of Elder Creek by the Late Cretaceous Coast Range fault (Jayko and Blake, 1986; Blake et al., 1992). The latter bounds the TCSM belt on the west and marks the eastern limit of Late Cretaceous extensional faulting responsible for exhumation of the deeply subducted eastern Franciscan rocks (Platt, 1986; Jayko et al., 1987). Near Wilbur Springs, the Stony Creek thrust and the Coast Range fault are folded (McLaughlin and Ohlin, 1984;

McLaughlin et al., 1989), and dextrally offset large distances by NW-trending Tertiary faults (McLaughlin et al., 1988, 1989). The TCSM belt probably continues into Napa County at least as far south as Lake Berryessa.

Conclusions
1. The Tehama-Colusa ultramafic belt is a serpentinite-matrix mlange (TCSM), best developed in its northern (Tehama) segment and patchily developed in its broad southern (Colusa) segment, especially along the eastern side. Mlange blocks range from kilometer scale down to meter scale. 2. The serpentinite protolith was peridotite (harzburgite > dunite) tectonite. Massive serpentinite with relict peridotite mineralogy and textures is widely overprinted by deformed (especially sheared) serpentinite. Areas of less-altered peridotite (without mlange blocks) occur mainly along the western side of the TCSM belt. 3. Non-native mlange blocks within the serpentinite consist chiefly of subalkaline basaltic submarine lava, plus less common diabase and rare plutonic rocks. Pillow structure in the lava is preserved locally but widely obscured or erased by deformation and weathering. Radiolarian chert mlange blocks, some capping basalt, are sparse but widespread. Rare argillite slabs, apparently capping chert in one case, are found in the Colusa mlange. 4. The peridotite is altered to chrysotile-lizardite-brucite serpentinite, signifying low-temperature hydrous alteration.The protolith lavas are partially altered to subgreenschist-facies mineral assemblages and weathering products, including clays. 5. The basaltic mlange-block lavas chiefly resemble Franciscan subalkaline lavas geochemically (Huot and Maury, 2002); those resembling Franciscan alkaline and transitional tholeiite seamount lavas are also present (Shervais and Kimbrough, 1987) but minor. The basalts are chemically distinct from most CRO lavas. 6. The TCSM basaltic lava plus peridotite protolith assemblage resembles Franciscan igneous/ meta-igneous rock assemblages. Diabase, gabbro, and clinopyroxenite/wehrlitecommon in the Coast Range ophiolite but rare in the Franciscanare rare also in the Tehama-Colusa serpentinite mlange. 7. The TCSM radiolarian cherts are nontuffaceous ribbon cherts that closely resemble unmetamorphosed Franciscan cherts. They differ from the

TEHAMA-COLUSA SERPENTINITE MLANGE

93

greenish-gray (white weathering) tuffaceous cherts (and tuff) that overlie the CRO. Their age, mainly late Middle Jurassic (BathonianCallovian) to early Late Jurassic (Oxfordian), is within the age span of Franciscan radiolarian cherts but older in part than the chiefly Late Jurassic (Oxfordian through Tithonian) tuffaceous cherts that overlie the CRO. 8. The protolith of the serpentinite mlange was basaltic oceanic crust overlying uppermost mantle including abyssal peridotite and opx dunite from the Moho transition zone. Radiolarian ribbon chert lay atop the basalt, capped by argillaceous mudstone (deep-sea clay?). Terriginous and/or volcaniclastic sandstones are absent, indicating an open-ocean setting away from a continent margin or volcanic arc. 9. A strong, widespread, Late Jurassic tectonic event disrupted the TCSM oceanic crust and upper mantle. Pervasive hydration from admixed seawater converted peridotite to serpentinite, which invaded and mixed with the disrupted basaltic crust. The deformation probably occurred along an intra-oceanic fracture zone but other nonsubduction origins are possible. 10. The Stonyford volcanic complex (SFVC) was a long-lived basaltic seamount that rose during the Middle and Late Jurassic (Shervais and Kimbrough, 1985a; Shervais and Hanan, 1989; Shervais et al., 2002, in press, b). It grew upon proto-TCSM basalt/ peridotite oceanic crust and mantle during the period of radiolarite deposition, chiefly prior to the Late Jurassic lithospheric disruption, serpentinization, and mlanging. 11. The TCSM belt, part of the Franciscan Jurassic oceanic lithosphere, differs strikingly from the Coast Range ophiolite, although both comprised segments of Jurassic oceanic lithosphere that originated in the Pacific paleoequatorial region. The TCSM/Franciscan and CRO oceanic plates had different tectonic histories during their Late Jurassic transport to higher paleolatitudes. The TCSM probably followed the CRO northward (after cessation of tephra clouds that drifted downwind from the Nevadan arc), and was emplaced against it along the ancestral Stony CreekBeehive Flat dextral transform fault during the latest Jurassic and earliest Cretaceous (TithonianValanginian). 12. Latest Jurassic to earliest Cretaceous (TithonianValanginian) terrigenous sediments, derived from the rising Nevadan orogen, spread seaward over the deep-ocean floor covering CRO and TCSM/

Franciscan oceanic basement while dextral transform displacement was in progress. 13. The juxtaposed TCSM and CRO/VP/OB oceanic terranes were underthrust and tilted eastward when Franciscan subduction began farther west during the Early Cretaceous (Valanginian). The Stony CreekBeehive Flat thrust fault was active during tilting, following the trace of the earlier transform fault. Tilting and uplift of the TCSM/CRO terranes formed an early Cretaceous submarine forearc ridge that bounded a new forearc basin that hosted Cretaceous GVG terriginous clastic sedimentation. 14. Serpentinous and ophiolitic detritus were shed from the rising TCSM/CRO forearc ridge, accumulating nearby beneath and within early Cretaceous GVG strata. 15. Geologic maps should distinquish between the broad TCSM belt, comprising unsubducted Jurassic eastern Franciscan terrane, and the Coast Range ophiolite remnant that is faulted against the TCSM at its northern end. No other CRO remnant is exposed west of Sacramento Valley from Thomes Creek (west of Paskenta) to Wilbur Springs (Fig. 1).

Dedication and Acknowledgments


This paper is dedicated to Robert G. Coleman, an authority on Coast Range geology, whose leadership especially in the study of serpentinites and oceanic crust-mantle rocks (ophiolites) worldwide is recognized internationally. Professor Coleman has been a longtime teacher to one of us (CAH), in California and abroad. It is a pleasure to salute this outstanding geologist and friend. We acknowledge with pleasure the guidance, writings, and discussions with Edgar Bailey, Clark Blake, Chuck Blome, Bob Coleman, Bill Dickinson, Gary Ernst, Donna M. Hull, Porter Irwin, Angela Jayko, Bruce Luyendyk, Jim Mattinson, Glenn MacPherson, Bob McLaughlin, Ben Page, Steve Phipps, John Shervais, and others who have inspired and influenced our work in the Coast Ranges. The expert collaboration of Mattinson and Luyendyk in our related work on the Coast Range ophiolite merits special recognition, as does the sustained encouragement of Gary Ernst. Ken Macdonald, Rachel Haymon, Tanya Atwater, and John Sinton have enhanced our understanding of MOR-generated oceanic lithosphere and its deformation. We especially acknowledge the incisive manuscript reviews by Bob McLaughlin, John Shervais, and

94

HOPSON AND PESSAGNO

John Wakabayashi, whose critical comments have enlightened us and improved the manuscript. Special thanks also to John Shervais, who provided manuscripts (in press) describing his work with others on some of the rocks described in this paper. Pessagnos investigations were supported by grants from the National Science Foundation (DES-7201528, DES-72 01528-A01, GA35094, EAR8615790, EAR-8816601, EAR-94181-94, EAR9304459), and Hopson appreciates the financial assistance (early 1970s) from the Academic Senate, UC Santa Barbara.

REFERENCES
Alvarez, W., Kent, D. V., Premoli-Silva, I., and Schweikert, R. A., 1980, Franciscan Complex limestone deposited at 17 South Latitude: Geological Society of America Bulletin, v. 91, p. 476484. Bailey, E. H., and Blake, M. C., Jr., 1974, Major chemical characteristics of Mesozoic Coast Range ophiolite in California: U.S. Geological Survey Journal of Research, v. 2, p. 637656. Bailey, E. H., Blake, M. C., and Jones, D. L., 1970, Onland Mesozoic oceanic crust in California Coast Ranges: U.S. Geological Survey Professional Paper 700-C, p. C70C81. Bailey, E. H., Irwin, W. P., and Jones, D. L., 1964, Franciscan and related rocks, and their significance in the geology of western California: California Division of Mines and Geology Bulletin 183, 177 p. Baumgartner, P. O., 1995, Towards a Mesozoic radiolarian databaseUpdates of work 1984-1990, in Baumgartner, P. O., ODogherty, L., Gorican, S., Urquhart, E., Pillevuit, A., and De Wever, P., eds., Middle Jurassic to Lower Cretaceous radiolaria of Tethys: Occurrences, systematics, biochronology: Memoires de Geologie (Lausanne): Lausanne, Switzerland, p. 689700 and Appendix, p. 10911106, 11441150. Beebe, W. J., 1986, A paleomagnetic study of the southern Coast Range ophiolite, California, and tectonic implications: Unpubl. M.A. thesis, University of California, Santa Barbara, California, 148 p. Beebe, W. J., and Luyendyk, B. P., 1983, Preliminary paleomagnetic results from the Llanada remnant of the Coast Range ophiolite, San Benito County, California (abstract): EOS (Transactions of the American Geophysical Union), v. 64, no. 45, p. 686. Benn, K., Nicolas, A., and Rueber, I., 1988, Mantle-crust transition zone and origin of wehrlitic magmas: Evidence from the Oman ophiolite, in Boudier, F., and Nicolas, A., eds., The ophiolites of Oman: Tectonophysics, v. 151, p. 7586. Berger, W. H., and Winterer, E. L., 1974, Plate stratigraphy and the flucuating carbonate line, in Hsu, K. J.,

and Jenkyns, H. C., eds., Pelagic sediments, on land and under the sea: Oxford, UK, Blackwell Scientific Publications, Ltd., International Association of Sedimentologists, Special Publication no. 1, p. 1148. Blackman, D. K., Cann, J. R., Janssen, B., Smith, D. K., 1998, Origin of extensional core complexes: Evidence from the Mid-Atlantic Ridge at Atlantis II fracture zone: Journal of Geophysical Research, v. 103, p. 21,31521,333. Blake, M. C., Jr., Helley, E. J., Jayko, A. S., Jones, D. L., and Ohlin, H. N., 1992, Geologic map of the Willows 1:100,000 quadrangle, California: U.S. Geological Survey Open-File Report 92-271, 38 p. Blake, M. C., Jr., Jayko, A. S., and Howell, D. G., 1982, Sedimentation, metamorphism, and tectonic accretion of the Franciscan assemblage in northern California, in Leggett, J. K., ed., Trench and forearc sedimentation in modern and ancient subduction zones: Geological Society of London Special Publication 10, p. 433438. Blake, M. C., Jr., Jayko, A. S., Jones, D. L., and Rogers, B. W., 1987, Unconformity between Coast Range ophiolite and part of the lower Great Valley sequence, South Fork of Elder Creek, Tehama County, California, in Hill, M. L., ed., Cordilleran Section of the Geological Society of America, Centennial Field Guide: Boulder, CO, Geological Society of America, p. 279282. Blake, M. C., Jr., Jayko, A. S., and McLaughlin, R. J., 1985, Tectonostratigraphic terranes of the northern Coast Ranges, California, in Howell, D. G., ed., Tectonostratigraphic terranes of the Circum-Pacific region: Houston, TX, Circum-Pacific Council for Energy and Mineral Resources, Earth Science Series, no. 1, p. 159171. Blake, M. C., Jr., Jayko, A. S., McLaughlin, R. J., and Underwood, M. B., 1988, Metamorphic and tectonic evolution of the Franciscan Complex, northern California, in Ernst, W.G., ed., Metamorphism and crustal evolution of the western United States (Rubey volume VII): Englewoods Cliffs, NJ, Prentice-Hall, p. 1035 1060. Blake, M. C., Jr., and Jones, D. L., 1974, Origin of Franciscan mlanges in northern California: Society of Economic Paleontologists and Mineralogists Special Paper 19, p. 255263. Boudier, F., and Nicolas, A., 1995, Nature of the Moho transition zone in the Oman ophiolite: Journal of Petrology, v. 36, p. 777796. Brown, R. D., 1964a, Geologic map of the Stonyford quadrangle, Glenn, Colusa, and Lake counties, California, Mineral Investigations Field Study Map MF-279: Scale 1:48000, U.S. Geological Survey. Brown, R. D., 1964b, Thrust-fault relations in the northern Coast Ranges, California: U.S. Geological Survey Professional Paper 475-D, p. D7D13. California Division of Mines and Geology, 1966, Geologic Map of California, scale 1:2,500,000: Sacramento, CA, California Division of Mines and Geology.

TEHAMA-COLUSA SERPENTINITE MLANGE

95

Cann, J. R., Blackman, D. K., Smith, D. K, McAllister, E., Janssen, B., Mello, S., Avgerinos, E., Pascoe, A. R., and Escartin, J., 1998, Corregated slip surfaces formed at ridge-transform intersections on the Mid-Atlantic Ridge: Nature, v. 385, p. 329332. Cannat, M., 1993, Emplacement of mantle rocks in the seafloor at mid-ocean ridges: Journal of Geophysical Research, v. 98, p. 41634172. Carlson, C., 1981a, Sedimentary serpentinites of the Wilbur Springs areaA possible early Cretaceous structural and stratigraphic link between the Franciscan Complex and the Great Valley sequence: Unpubl. M.S. thesis, Stanford University, 105 p. Carlson, C., 1981b, Upwardly mobile mlanges, serpentinite protrusions, and transport of tectonic blocks in accretionary prisms [abs.]: Geological Society of America Abstracts with Programs, v. 13, no. 2, p. 48. Ceuleneer, G., and Rabinowicz, M., 1992, Mantle flow and melt migration beneath oceanic ridges: Models derived from observations in ophiolites, in Morgan, J. P., Blackman, D. K., and Sinton, J. M., eds., Mantle flow and melt generation at mid-ocean ridges: American Geophysical Union Geophysical Monograph 71, p. 123154. Coleman, R. G., 1971, Petrologic and geophysical nature of serpentinites: Geological Society of America Bulletin, v. 82, p. 897918. Coleman, R. G., 2000, Prospecting for ophiolites along the California continental margin, in Dilek, Y., Moores, E. M., Elthon, D., and Nicolas, A., eds., Ophiolites and oceanic crust: New insights from field studies and the Ocean Drilling Program: Geological Society of America Special Paper 349, p. 351364. Coleman, R. G., and Lanphere, M. A., 1971, Distribution and ages of high-grade blueschists, associated eclogites, and amphibolites from Oregon and California: Geological Society of America Bulletin, v. 82, p. 23972412. Constenius, K. N., Johnson, R. A., Dickinson, W. R., and Williams, T. A., 2000, Tectonic evolution of the JurassicCretaceous Great Valley forearc, California: Implications for the Franciscan thrust-wedge hypothesis: Geological Society of America Bulletin, v. 112, p. 17031723. Cowan, D. S., and Bruhn, R. L., 1992, Late Jurassic to early Late Cretaceous geology of the U.S. Cordillera, in Burchfiel, B. C., Lipman, P. W., and Zoback, M. L., eds., The Cordilleran Orogen: Conterminous U.S.: Boulder, CO, Geological Society of America, The Geology of North America, v. G-3, p. 169203. Crerar, D. A., Namson, J., Chyi, M. S., Williams, L., and Feigenson, M. D., 1982, Manganiferous cherts of the Franciscan assemblage: I. General geology, ancient and modern analogues, and implications for hydrothermal convection at oceanic spreading centers: Economic Geology, v. 77, p. 519540.

Dick, H. J. B., and Bullen, T., 1984, Chromian spinel as a petrogenetic indicator in abyssal and alpine-type peridotites and spatially associated lavas: Contributions to Mineralogy and Petrology, v. 86, p. 5476. Dickinson, W. R., Hopson, C. A., and Saleeby, J. B., 1996, Alternate origins of the Coast Range ophiolite (California): Introduction and implications: GSA Today, v. 6, no. 2, p. 110. Dickinson, W. R., Ingersoll, R. V., Cowan, D. S., Helmold, K. P., and Suczek, C. A., 1982, Provenance of Franciscan graywackes in Coastal California: Geological Society of America Bulletin, v. 93, p. 95107. Dickinson, W. R., and Rich, E. I., 1972, Petrologic intervals and and petrofacies in the Great Valley sequence, Sacramento Valley, California: Geological Society of America Bulletin, v. 83, p. 30073024. Dilek, Y., Moores, E. M., and Furnes, H., 1998, Structure of modern oceanic crust and ophiolites and implications for faulting and magmatism, at oceanic spreading centers, in Buck, R., Karson, J., Delany, P., and Lagabrielle, Y., eds., Faulting and magmatism at mid-ocean ridges: American Geophysical Union Mongraph 106, p. 219265. Edelman, S. H., Day, H. W., Moores, E. M., Zigan, S. M., Murphey, T. P., and Hacker, B. R., 1989, Structure across a Mesozoic ocean-continent suture zone in the northern Sierra Nevada, California: Geological Society of America Special Paper 224, 53 p. Ernst, W. G., 1970, Tectonic contact between Franciscan mlange and the Great Valley SequenceCrustal expression of a late Mesozoic Benioff zone: Journal of Geophysical Research, v. 75, p. 886901. Ernst, W. G., 1971, Metamorphic zonations on presumably subducted lithospheric plates from Japan, California, and the Alps: Contributions to Mineralogy and Petrology, v. 34, p. 4359. Ernst, W. G., 1977, Tectonics and prograde versus retrograde P-T trajectories of high-pressure metamorphic belts: Rendiconti Societa Italiana di Mineralogia e Petrologia, v. 33, p. 191220. Ernst, W. G., 1984, Phanerozoic continental accretion and the metamorphic evolution of northern and central California: Tectonophysics, v. 100, p. 287320. Fritz, D. M., 1975, Ophiolite belt west of Paskenta, northern California Coast Ranges, California: Unpubl. M.A. thesis, University of Texas, Austin, Texas. Giaramita, M., MacPherson, G. J., and Phipps, S. P., 1998, Petrologically diverse basalts from a fossil oceanic forearc in California: The Llanada and Black Mountain remnants of the Coast Range ophiolite: Geological Society of America Bulletin, v. 110, p. 553571. Girardeau, J., and Francheteau, J., 1993, Plagioclasewehrlites and peridotites on the East Pacific Rise (Hess Deep) and the Mid-Atlantic Ridge (DSDP Site 334): Evidence for magma perculation in the oceanic upper mantle: Earth and Planetary Science Letters, v. 115, p. 137149.

96

HOPSON AND PESSAGNO

Godfrey, N. J., Beaudoin, B. C., Klemperer, S. L., and Mendocino Working Group, 1997, Ophiolitic basement to the Great Valley forearc basement, California, from seismic and gravity data: Implications for crustal growth at the North American continental margin: Geological Society of America Bulletin, v. 108, p. 15361562. Godfrey, N. J., and Dilek, Y., 2000, Mesozoic assimilation of oceanic crust and island arc into the North American continental margin in California and Nevada: Insights from geophysical data, in Dilek, Y., Moores, E. M., Elthon, D., and Nicolas, A., eds., Ophiolites and oceanic crust: New insights from field studies and the Ocean Drilling Program: Geological Society of America Special Paper 349, p. 365382. Godfrey, N. J., and Klemperer, S. L., 1998, Ophiolitic basement to a forearc basin and implications for continental growth: The Coast Range/Great Valley ophiolite, California: Tectonics, v. 17, p. 558570. Hanan, B. B., Kimbrough, D. L., Renne, P. R., Shervais, J. W., 1991, Stonyford volcanic complex: Pb isotopes and 40Ar/39Ar ages of volcanic glasses from a Jurassic seamount in the northern California Coast Ranges, Geological Society of America Abstracts with Programs, v. 23, no. 5, p. A395. Harms, T. A., Jayko, A. S., and Blake, M. C., Jr., 1992, Kinematic evidence for extensional unroofing of the Franciscan Complex along the Coast Range fault, northern Diablo Range, California: Tectonics, v. 11, p. 228v241. Harper, G. D., Saleeby, J. B., and Heizler, M., 1994, Formation and emplacement of the Josephine ophiolite and the Nevadan orogeny in the Klamath Mountains, California-Oregon: U-Pb zircon and 40Ar/39Ar geochronology: Journal of Geophysical Research, v. 99, p. 42934321. Hekinian, R., Bideau, D., Francheteau, J., Cheminee, J. L., Armijo, R., Lonsdale, P., and Blum, P., 1993, Petrology of the East Pacific Rise crust and upper mantle exposed in Hess Deep (Eastern Equatorial Pacific): Journal of Geophysical Research, v. 98, p. 80698094. Hopson, C. A., 2002, Wehrlitic magma series and upper mantle transition zone, California Coast Range ophiolite: Ocean ridge off-axis magmatism [abs.]: Geological Society of America Abstracts with Programs, v. 34, no. 3, p. A-23. Hopson, C. A., Mattinson, J. M., and Pessagno, E. A., Jr., 1981, Coast Range ophiolite, western California, in Ernst, W. G., ed., The geotectonic development of California (Rubey Volume 1): Englewood Cliffs, NJ, Prentice-Hall, Inc., p. 418510. Hopson, C. A., Mattinson, J. M., Pessagno, E. A., Jr., and Luyendyk, B. P., in prep., California Coast Range ophiolite: Composite Middle and Late Jurassic oceanic lithosphere.

Hopson, C., Pessagno, E., Mattinson, J., Luyendyk, B., Beebe, W., Hull, D., Munoz, E., and Blome, C., 1996, Coast Range ophiolite as paleoequatorial mid-ocean lithosphere, p. 46 in Dickinson, W. R., Hopson, C. A., and Saleeby, J. B., eds., Alternate origins of the Coast Range ophiolite (California): Introduction and implications: GSA Today, v. 6, p. 110. Hull, D. M., 1995, Morphologic diversity and paleogeographic significance of the Family Parvicingulidae (Radiolaria): Micropaleontology, v. 41, p. 148. Hull, D. M., 1997, Upper Jurassic Tethyan and Southern Boreal radiolarians from western North America: Micropaleontology, v. 43, suppl. 2, p. 1202. Hull, D. M., Blome, C. D., and Pessagno, E. A., Jr., 1997, Paleomagnetism of Jurassic radiolarian chert above the Coast Range ophiolite at Stanley Mountain, California, and its implications for its paleogeographic origins: Discussion and reply: Geological Society of America Bulletin, v. 109, p. 16331639. Hull, D. M., Pessagno, A. E., Jr., Hopson, C. A., Blome, C. D., and Munoz, I. M., 1993, Chronostratigraphic assignment of volcanopelagic strata above the Coast Range ophiolite, in Dunne, G., and McDougall, K., eds., Mesozoic paleogeography of the Western United StatesII: Pacific Section, Society of Economic Paleontologists and Mineralogists, Book 71, p. 151170. Huot, F., and Maury, R. C., 2002, the Round Mountain serpentinite mlange, northern Coast Ranges of California: An association of backarc and arc-related tectonic units: Geological Society of America Bulletin, v. 114, p. 109123. Ingersoll, R. V., 1988, Tectonics of sedimentary basins: Geological Society of America Bulletin, v. 100, p. 17041719. Ingersoll, R. V., 1990, Nomenclature of Upper Mesozoic strata of the Sacramento Valley of California: Review and recommendations, in Ingersoll, R. V., and Nilsen, T. H., eds., Sacramento Valley symposium and guidebook: Pacific Section, Society of Economic Paleontologists and Mineralogists, v. 65, p. 13. Ingersoll, R. V., 2000, Models for origin and emplacement of Jurassic ophiolites in northern California, in Dilek, Y., Moores, E. M., Elthon, D., and Nicolas, A., eds, Ophiolites and oceanic crust: New insights from field studies and the Ocean Drilling Program: Geological Society of America Special Paper 349, p. 395402. Jachens, R. C., Griscom, A., and Roberts, C. W., 1995, Regional extent of Great Valley basement west of the Great Valley, California: Implications for extensive tectonic wedging in the California Coast Ranges: Journal of Geophysical Research, v. 100, p. 1276912790. Jayko, A. S., 1984, Structure and deformation in the eastern Franciscan belt, northern California: Unpubl. Ph.D. dissertation, University of California, Santa Cruz, 218 p. Jayko, A. S., and Blake, M. C., Jr., 1986, Significance of Klamath rocks between the Franciscan Complex and

TEHAMA-COLUSA SERPENTINITE MLANGE

97

Coast Range ophiolite, northern California: Tectonics, v. 5, p. 10551071. Jayko, A. S., Blake, M. C., Jr., and Harms, T., 1987, Attenuation of the Coast Range ophiolite by extensional faulting, and nature of the Coast Range thrust, California: Tectonics, v. 6, p. 475488. Jayko, A. S., Blake, M. C., Jr., and Brothers, R. N., 1986, Blueschist metamorphism of the eastern Franciscan belt, northern California, in Evans, B. W., and Brown, E. H., eds., Blueschists and eclogites: Geological Society of America Memoir 164, p. 107123. Jennings, C. W., 1977, Geologic Map of California, scale 1:750,000: Sacramento, CA, California Division of Mines and Geology. Jennings, C. W., and Strand, R. G., 1960, Geologic Map of California, Ukiah sheet, scale 1:250,000: Sacramento, CA, California Division of Mines. Jones, D. L., Bailey, E. H., and Imlay, R. W., 1969, Structural and stratigraphic significance of the Buchia zones in the Colyear SpringPaskenta area, California: U.S. Geological Survey Professional Paper 647-A, p. 124. Juteau, T., Ernewein, M., Reuber, I., Whitechurch, H., and Dahl, R., 1988, Duality of magmatism in the plutonic sequence of the Sumail nappe, Oman, in Boudier, F., and Nicolas, A., eds., The ophiolites of Oman: Tectonophysics, v. 151, p. 107135. Kelemen, P. B., 1990, Reaction between ultramafic rock and fractionating basaltic magma; phase relations, and the origin of calc-alkaline magma series, and the formation of discordant dunite: Journal of Petrology, v. 31, p. 5198. Kelemen, P. B., and Dick, H. J. B., 1995, Focused melt flow and localized deformation in the upper mantle: Juxtaposition of replacive dunite and ductile shear zones in the Josephine peridotite, SW Oregon: Journal of Geophysical Research, v. 100, no. B1, p. 423438. Kennett, J. P., 1982, Marine geology: Englewood Cliffs, NJ, Prentice Hall, 795 p. Kleinrock, M. C., and Hey, R. N., 1989, Migrating transform zone and lithosphere transfer at the Galapagos 95.5W propagator: Journal of Geophysical Research, v. 94, p. 13,85913,878. Lagabrielle, Y., Roure, F., Coutelle, A., Maury, R. C., Joron, J.-L., and Thonon, P., 1986, The Coast Range ophiolites (northern California): Possible arc and back-arc basin remnants: Their relations with the Nevada orogeny, United States, in United States: Basin and Range and metamorphic core complexes, Coast Ranges, ophiolites and arc series, Mexico, Geol. Cordillera of North America: Paris, France, Geological Society of France, p. 981999. Lanphere, M. A., Blake, M. C., Jr., and Irwin, W. P., 1978, Early Cretaceous metamorphic age of the South Fork Mountain schist in the northern Coast Ranges of California: American Journal of Science, v. 278, p. 798 815.

Louvion-Trellu, C., 1986, Les radiolarites des Coast Ranges, Californie: Etude biochronologique, sedimentologique et geochimique: DEA thesis, Universite de Bretagne Occidentale, Brest, France, 46 p. Luyendyk, B. P., and Hornafias, S. J., 1982, Paleolatitude of the Point Sal ophiolite [abs.]: Geological Society of America Abstracts with Programs (Cordilleran Section), v. 14, p. 182. MacPherson, G. J., 1983, The Snow Mountain Complex: An on-land seamount in the Franciscan terrain, California: Journal of Geology, v. 91, p. 7392. MacPherson, G. J., and Phipps, S. P., 1985, Comment, in Comment and Reply on Geochemical evidence for the tectonic setting of the Coast Range ophiolite: A composite island arc-oceanic crust terrane in western California: Geology, v. 13, p. 827828. Maxwell, J. C., 1974, Anatomy of an orogen: Geological Society of America Bulletin, v. 85, p. 11951204. McDowell, F. W., Lehman, D. H., Gucwa, P. R., Fritz, D., and Maxwell, J. C., 1984, Glaucophane schists and ophiolites of the northern California Coast Ranges: Ages, mineralogy, and their tectonic implications: Geological Society of America Bulletin, v. 95, p. 13731382. McLaughlin, R. J., 1978, Preliminary geologic map and structural sections of the central Mayacamas Mountains and the Geysers steam field, Sonoma, Lake, and Mendocino counties, California: U.S. Geological Survey Open-File Map 78-389, scale 1:24,000. McLaughlin, R. J., Blake, M. C., Jr., Griscom, C. D., Blome, C. D., and Murchey, B. L., 1988, Tectonics of formation, translation, and dispersal of the Coast Range ophiolite of California: Tectonics, v. 7, p. 1033 1056. McLaughlin, R. J., and Ohlin, H. N., 1984, Tectonic framework of The GeysersClear Lake region, California, in Blake, M. C., Jr., ed., Franciscan geology of Northern California: Society of Economic Paleontologists and Mineralogists, Pacific Section Book Series, v. 43, p. 221254. McLaughlin, R. J., Ohlin, H. N., Thormalen, D. J., Jones, D. L., Miller, J. W., and Blome, C. D., 1989, Geologic map and structure sections of the Little Indian Valley Wilbur Springs geothermal area, northern Coast Ranges, California, scale 1:24,000: U.S. Geological Survey Miscellaneous Investigations Map I-1706, Sheets 1 and 2. Moiseyev, A. N., 1966, Geology and geochemistry of the Wilbur Springs quicksilver deposits, Lake and Colusa Counties, California: Unpubl. Ph.D. dissertation, Stanford University, Stanford, California. Moiseyev, A. N., 1970, Late serpentinite movements in the California Coast Ranges: New evidence and its indications: Geological Society of America Bulletin, v. 81, p. 17211732. Moxon, I. W., 1988, Sequence stratigraphy of the Great Valley basin in the context of convergent margin

98

HOPSON AND PESSAGNO

tectonics, in Graham, S. A., ed., Studies of the geology of the San Joaquin basin: Pacific Section, Society of Economic Paleontologists and Mineralogists Book 60, p. 328. Moxon, I. W., 1990, Stratigraphy and structure of Upper JurassicLower Cretaceous strata, Sacramento Valley, in Ingersoll, R. V., and Nilsen, T. H., eds., Sacramento Valley symposium and guidebook: Pacific Section, Society of Economic Paleontologists and Mineralogists Book 65, p. 529. Murchey, B. M., 1984, Biostratigraphy and lithostratigraphy of chert in the Franciscan Complex, Marin Headlands block, California, in Blake, M. C., Jr., ed., Franciscan geology of Northern California: Society of Economic Paleontologists and Mineralogists, Pacific Section Book Series, v. 43, p. 5170. Murchey, B. L., and Blake, M. C., Jr., 1993, Evidence for subduction of a major ocean plate along the California margin during the Middle to early Late Jurassic, in Dunn, G., and McDougall, K., eds., Mesozoic paleogeography of the Western United States II: Society of Economic Paleontologists and Mineralogists, Pacific Section, Book 71, p. 118. Murchey, B. M., and Hagstrum, J. T., 1997, Paleomagnetism of Jurassic radiolarian chert above the Coast Range ophiolite at Stanley Mountain, California, and implications for its paleogeographic origins: Reply: Geological Society of America Bulletin, v. 109, p. 16331639. Murchey, B. M., and Jones, D. L., 1984, Age and significance of chert in the Franciscan Complex in the San Francisco Bay region, in Blake, M. C., ed., Franciscan geology of Northern California: Society of Economic Paleontologists and Mineralogists, Pacific Section Book Series, v. 43, p. 2330. Nicolas, A., 1988, Melt extraction model based on structural studies in mantle peridotites: Journal of Petrology, v. 27, p. 9991022. Nicolas, A., and Boudier, A., 2000, Large mantle upwellings and related variations in crustal thickness in the Oman ophiolite, in Dilek, Y., Moores, E. M., Elthon, D., and Nicolas, A., eds., Ophiolites and oceanic crust: New insights from field studies and the Ocean Drilling Program: Boulder, CO, Geological Society of America Special Paper 349, p. 6773. Nicolas, A., and Prinzhofer, A., 1983, Cumulative or residual origin of the transition zone in ophiolites: Structural evidence: Journal of Petrology, v. 24, p. 188206. Ogg, J. G., Karl, S. M., and Behl, R. J., 1992, Jurassic through early Cretaceous sedimentation history of the central equatorial Pacific and of Sites 800 and 801, in Larson, R., Lancelot, Y., et al., eds., Proceedings of the Ocean Drilling Program, Scientific Results, v. 129, p. 571613. Page, B. M., 1966, Geology of the Coast Ranges of California, in Bailey, E. H., ed., Geology of Northern Califor-

nia: California Division of Mines and Geology Bulletin 190, p. 255276. Page, B. M., 1981, The southern Coast Ranges, in Ernst, W. G., ed., The geotectonic development of California (Rubey Volume I): Englewood Cliffs, NJ, PrenticeHall, p. 329417. Pallister, J. S., and Gregory, R. T, 1983, Composition of the Samail ocean crust: Geology, v. 11, p. 638642. Pallister, J. S., and Hopson, C. A., 1981, Samail ophiolite plutonic suite: Field relations, phase variation, cryptic variation and layering, and a model of a spreading ridge magma chamber: Journal of Geophysical Research, v. 86. p. 25932644. Pessagno, E. A., Jr., and Blome, C. D., 1986, Faunal affinities and tectonogenesis of Mesozoic rocks in the Blue Mountains Province of eastern Oregon and western Idaho, in Vallier, T. L., and Brooks, H. C., eds., Geology of the Blue Mountains region of Oregon, Idaho, and Washington: Geologic Implications of Paleozoic and Mesozoic paleontology and biostratigraphy: U.S. Geological Survey Professional Paper 1435, p. 6578. Pessagno, E. A., Jr., and Blome, C. D., 1990, Implications of new Jurassic stratigraphic, geochronometric, and paleolatitudinal data from the Western Klamath terrane (Smith River and Rogue Valley subterranes): Geology, v. 18, p. 665668. Pessagno, E. A., Jr., Blome, C. D., Carter, E. S., McLeod, N., Whelan, P. A., and Yeh, K. Y., 1987, Studies of North American Jurassic Radiolaria; Part II, Preliminary radiolarian zonation for the Jurassic of North America: Special Publications, Cushman Foundation for Foraminiferal Research, v. 23, p. 118. Pessagno, E. A., Jr., Blome, C. D., Hull, D. M., and Six, W. M., Jr., 1993, Middle and Upper Jurassic radiolaria from the Western Klamath terrane, Smith River subterrane, northwestern California: Their biostratigraphic, chronostratigraphic, geochronologic, and paleolatitudinal significance: Micropaleontology, v. 39, p. 93 166. Pessagno, E. A., Jr., Cantu-Chapa, A., Hull, D. M., Kelldorf, M., Longoria, J. F., Martin, C., Meng, X., Montgomery, H., Fucugauchi, U., and Ogg, J. G., 1999, Stratigraphic evidence for northwest to southeast tectonic transport of Jurassic terranes in central Mexico and the Caribbean (Cuba), in Mann, P., ed., Caribbean basins: Sedimentary basins of the world, 4: Amsterdam, Elsevier Science B.V., p. 123149. Pessagno, E. A., Jr., and Hull, D. M., 1996, Once upon a time in the Pacific: Chronostratigraphic misinterpretation of basal strata at ODP Site 801 (Central Pacific) and its impact on geochronology and plate tectonics models: Switzerland, 1996 Transtec Publications, GeoResearch Forum Volumes 1-2, p. 7992. Pessagno, E. A., Jr., Hull, D. M., and Hopson, C. A., 2000, Tectonostratigraphic significance of sedimentary strata occurring within and above the Coast Range ophiolite (California Coast Ranges) and the Josephine ophiolite

TEHAMA-COLUSA SERPENTINITE MLANGE

99

(Klamath Mountains), northwestern California, in Dilek, Y., Moores, E. M., Elthon, D., and Niclas, A., eds., Ophiolites and oceanic crust: New insights from field studies and the Ocean Drilling Program: Geological Society of America Special Paper 349, p. 383 394. Pessagno, E. A., Jr., Hull, D. M., and Pujana, I., 1991, Correlation of Circum-Pacific upper Tithonian Boreal and Tethyan strata in Northern and Southern hemispheres: Synthesis of radiolarian and ammonite biostratigraphic and chronostratigraphic data [abs.]: Poitiers, France, 3rd International Symposium on Jurassic Stratigraphy, Abstract Volume 97. Pessagno, E. A., Jr., and Martin, C., 2003, Tectonostratigraphic evidence for the origin of the Gulf of Mexico, in Bazrtolini, C., Buffler, R. T., and Blickwede, J., eds., The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon habitats, basin formation, and plate tectonics: American Association of Petroleum Geologists Memoir 79, p. 4674. Phipps, S. P., 1984, Ophiolitic olistostromes in the basal Great Valley sequence, Napa County, northern California Coast Ranges: Geological Society of America Special Paper 198, p. 103125. Platt, J. P., 1986, Dynamics of orogenic wedges and the uplift of high-pressure metamorphic rocks: Geologic Society of America Bulletin, v. 97, p. 10371053. Quick, J. E., 1981, The origin and significance of large, tabular dunite bodies in the Trinity peridotite, northern California: Contributions to Mineralogy and Petrology, v. 78, p. 413422. Ring, U., and Brandon, M. T., 1994, Kinematic data for the Coast Range fault and implications for the exhumation of the Franciscan subduction complex: Geology, v. 22, p. 735738. Robertson, A. H. F., 1989, Palaeoceanography and tectonic setting of the Jurassic Coast Range ophiolite, central California: Evidence from the extrusive rocks and the volcaniclastic sediment cover: Marine and Petroleum Geology, v. 6, p. 193220. Robertson, A. H. F., 1990, Sedimentary and tectonic implications of ophiolite-derived clastics overlying the Jurassic Coast Range ophiolite, northern California: American Journal of Science, v. 290, p. 109163. Robinson, P. T., Dick, H. J. B., Natland, J. H., and ODP Leg 176 Shipboard Party, 2000, Lower oceanic crust formed at an ultra slow-spreading ridge: Ocean Drilling Program Hole 735B, Southwest Indian Ridge, in Dilek, Y., Moores, E. M., Elthon, D., and Nicolas, A., eds., Ophiolites and oceanic crust: New insights from field studies and the Ocean Drilling Program: Boulder, CO, Geological Society of America Special Paper 349, p. 7586. Saleeby, J. B., 1979, Kaweah serpentinite mlange, southwest Sierra Nevada foothills, California: Geological Society of America Bulletin, v. 90, p. 2946.

Saleeby, J. B., 1981, Ocean floor accretion and volcanoplutonic arc evolution of the Mesozoic Sierra Nevada, in Ernst, W. G., ed., Geotectonic development of California (Rubey Volume 1): Englewood Cliffs, NJ, Prentice-Hall, p. 132181. Saleeby, J. B., 1984, Tectonic significance of serpentinite mobility and ophiolitic mlange: Geological Society of America Special Paper 198, p. 153168. Sedlock, R. L., and Isozaki, Y., 1990, Lithology and biostratigraphy of Franciscan-like chert and associated rocks in west-central Baja California, Mexico: Geological Society of America Bulletin, v. 102, p. 852864. Seiders, V. M., 1988, Origin of conglomerate in the Franciscan assemblage and Great Valley sequence, northern California: Geology, v. 16, p. 783787. Sharp, W. D., 1988, Pre-Cretaceous crustal evolution in the Sierra Nevada region, California, in Ernst, W. G., ed., Metamorphism and crustal evolution of the western United States (Rubey Volume 7): Englewood Cliffs, NJ, Prentice-Hall, p. 824864. Shervais, J. W., 1990, Island arc and ocean crust ophiolites: Contrasts in the petrology, geochemistry, and tectonic style of ophiolite assemblages in the California Coast Ranges, in Malpas, J., Moores, E. M., Panayiotou, A., and Xenophontos, C., eds., Ophiolites, oceanic crustal analogues: Nicosia, Cyprus, The Geological Survey Department, Ministry of Agriculture and Natural Resources, p. 507520. Shervais, J. W., 1993, Tectonic implications of oceanic basalts in the Coast Range ophiolite, California: Evidence for ridge subduction as final ophiolite-forming event [abs.]: Geological Society of America Abstracts with Programs, v. 25, no. 6, p. A445A446. Shervais, J. W., 2001, Birth, death, and resurrection: The life cycle of suprasubduction zone ophiolites: Geochemistry, Geophysics, Geosystems, v. 2, [paper number 2000GC000080; 20,925 words, 8 figures, 3 tables; published January 31, 2001]. Shervais, J. W., and Hanan, B. B., 1989, Jurassic volcanic glass from the Stonyford volcanic complex, Franciscan assemblage, northern California Coast Ranges: Geology, v. 117, p. 510514. Shervais, J. W., and Kimbrough, D. L., 1985a, Geochemical evidence for the origin of the Coast Range ophiolite: A composite island arc-oceanic crust terrane in western California: Geology, v. 13, p. 3538. Shervais, J. W., and Kimbrough, D. L., 1985b, Comment and reply on Geochemical evidence for the tectonic setting of the Coast Range ophiolite: A composite island arc-oceanic crust terrane in western California: Geology, v. 13, p. 828829. Shervais, J. W., and Kimbrough, D. L., 1987, Alkaline and transitional subalkaline metabasalts in the Franciscan Complex mlange, California: Geological Society of America Special Paper 215, p. 165182. Shervais, J. W, Kimbrough, D. L., Renne, P., Hanan, B. B., Murchey, B. L., Snow, C. A., Zoglman-Schuman,

100

HOPSON AND PESSAGNO

M. M., and Beaman, J., 2004, Multi-stage origin of the Coast Range ophiolite, California: Implications for the life cycle of supra-subduction zone ophiolites: International Geology Review, v. 46, p. 289315. Shervais, J. W., Murchey, B. L., Kimbrough, D. L., Renne, P. R., and Hanan, B. B., 2002, Radiometric and biostratigraphic age relations in the Coast Range ophiolite (CRO), northern California: Implications for Jurassic tectonic evolution of the western Cordillera [abs.]: Geological Society of America Abstracts with Programs, v. 34, no. 6, p. A-23. Shervais, J. W., Kolesar, P., and Andreasen, K., in press, a, A chemical study of serpentinizationStonyford, California: Mass balance and chemical flux: International Geology Review. Shervais, J. W., Murchey, B. L., Kimbrough, D. L., Renne, P. R., and Hanan, B., in press, b, Radioisotopic and biostratigraphic age relations in the Coast Range ophiolite: Geological Society of America Bulletin. Sinton, J. M., and Detrick, R. S., 1992, Mid-ocean ridge magma chambers: Journal of Geophysical Research, v. 97, p. 197216. Stern, R. J., and Bloomer, S. H., 1992, Subduction zone infancy: Examples from the Eocene Izu-Bonin-Mariana and Jurassic California arcs: Geological Society of America Bulletin, v. 104, p. 16211636. Suppe, J., 1973, Geology of the Leech Lake Mountain Ball Mountain region, CaliforniaA cross section of the northeastern Franciscan belt and its tectonic implications: University of California Publications in Geological Sciences, v. 107, 82 p. Suppe, J., and Armstrong, R. L., 1972, Potassium-argon dating of Franciscan metamorphic rocks: American Journal of Science, v. 272, p. 217233. Suppe, J., and Foland, K. A., 1978, The Goat Mountain schists and Pacific Ridge Complex; a redeformed but still-intact late Mesozoic Franciscan schuppen complex, in Howell, D. G., and McDougall, K., eds., Mesozoic paleogeography of the Western United States: Society of Economic Paleontologists and Mineralogists, Pacific Section, Pacific Coast Paleogeography Symposium 2, p. 431451. Thy, P., and Dilek, Y., 2000, Magmatic and tectonic controls on the evolution of oceanic magma chambers at slow-spreading ridges: Perspectives from ophiolitic and continental layered intrusions, in Dilek, Y., Moores, E. M., Elthon, D., and Nicolas, A., eds., Ophiolites and oceanic crust: New insights from field studies and the Ocean Drilling Program: Boulder, CO,

Geological Society of America Special Paper 349, p. 87104. Unruh, J. R., Loewen, B. A., and Moores, E. M., 1995, Progressive arcward contraction of a MesozoicTertiary fore-arc basin, southwestern Sacramento Valley, California: Geological Society of America Bulletin, v. 107, p. 3853. Unruh, J. R., Ramirez, V. R., Phipps, S. R., and Moores, E. M., 1991, Tectonic wedging beneath forearc basins: Ancient and modern examples from California and the Lesser Antilles: GSA Today, v. 1, no. 9, p. 187190. Wakabayashi, J., 1990, Counterclockwise P-T-t paths for amphibolites, Franciscan Complex, California: Relics from the early stages of subduction zone metamorphism: Journal of Geology, v. 98, p. 657680. Wakabayashi, J., 1999, Subduction and the rock record: Concepts developed in the Franciscan Complex, California, in Moores, E. M., Sloan, D., and Stout, D. L., eds., Classic Cordilleran concepts: A view from California: Geological Society of America Special Paper 338, p. 123133. Wakabayashi, J., and Unruh, J. R., 1995, Tectonic wedging, blueschist metamorphism, and exposure of blueschists: Are they compatible?: Geology, v. 23, p. 85 88. Wentworth, C. M., Blake, M. C., Jr., Jones, D. L., Walter, A. W., and Zoback, M. D., 1984, Tectonic wedging associated with the emplacement of the Franciscan assemblage, California Coast Ranges, in Blake, M. C., Jr., ed., Franciscan geology of northern California: Society of Economic Paleontologists and Mineralogists, Pacific Section, v. 43, p. 163173. Winchester, J. A., and Floyd, P. A., 1977, Geochemical discrimination of different magma series and their differentiation products using immobile elements: Geochemical Geology, v. 20, p. 325343. Worrall, D. M., 1981, Imbricate low-angle faulting in uppermost Franciscan rocks, south of Yolla Bolly area, northern California: Geological Society of America Bulletin, v. 92, p. 703729. Wright, J. E., and Wyld, S. J., 1994, The Rattlesnake Creek terrane, Klamath Mountains, California: An early Mesozoic volcanic arc and its basement of tectonically disrupted oceanic crust: Geological Society of America Bulletin, v. 106, p. 10331056. Zoglman, M. M., and Shervais, J. W., 1991, The Stonyford volcanic complex: Petrology and structure of a Jurassic seamount in the northern Coast Ranges of California [abs.]: Geological Society of America Abstracts with Programs, v. 23, no. 5, p. A395.

Anda mungkin juga menyukai