Sie sind auf Seite 1von 35

Late Cenozoic Tectonics of the East Ventura Basin,

Transverse Ranges, California1


Robert S. Yeats, Gary J. Huftile, and Leonard T. Stitt2

ABSTRACT
The east Ventura basin originated in the middle
Miocene as a rift system bounded on one side by the
Oak RidgeSimi Hills structural shelf and on the
other side by a granitic ridge parallel to the San
Gabriel fault. This fault began accumulating right
slip 1012 m.y. ago at a rate of 4.59 mm/yr
(depending on whether total slip is 45 or 60 km),
slowing to about 1 mm/yr in the Quaternary. North
of the Santa Clara River, rifting ended prior to deposition of the uppermost Miocenelower Pliocene
Towsley Formation. South of the Santa Clara River,
the rift axis shifted southwest toward the Oak
RidgeSimi Hills shelf as the Towsley Formation
accumulated against a normal-fault ancestor of the
Santa Susana fault. A change to contractile tectonics
occurred in the Pliocene with deposition of the Fernando Formation, when the Newhall-Potrero anticline developed as a monocline above a blind
reverse fault; the Pico anticline to the southeast and
the Temescal and Hopper RanchModelo anticlines
to the northwest may have a similar origin. Tectonic
inversion and displacement on the southwest-verging Santa Susana fault began about 0.5 Ma based on

Copyright 1994. The American Association of Petroleum Geologists. All


rights reserved.
1Manuscript received, February 10, 1992; revised manuscript received,
March 1, 1994; final acceptance, March 4, 1994.
2Department of Geosciences, Oregon State University, Corvallis, Oregon
97331.
This study was supported by contracts 14-08-0001-G1372 and 14-080001-G1798, which built on earlier work supported by contracts 14-08-000115886, 14-08-0001-16747, 14-08-0001-19138, and 14-08-0001-21279 from
the Earthquake Hazards Reduction Program of the U. S. Geological Survey.
Additional support was provided by petroleum-industry grants to the Southern
California Fault Studies program at Oregon State University. Leonard Stitt
received salary support from Arco Exploration and from Conoco, Inc. for parts
of the study. Discussions with Bill Cotton, John Crowell, Tom Dibblee, Perry
Ehlig, Ed Hall, Tom Hopps, Shaul Levi, Dick Saul, Jerry Treiman, Harold
Weber, and Jack West helped sharpen our understanding, although Harold
Weber was never able to convince us that the San Gabriel fault was not a
major strike-slip fault. John Crowell, Tom Dibblee, and Jerry Treiman provided us with unpublished geologic maps, and Crowell, in addition, provided a
detailed review of the paper. Margaret Mumford prepared the illustrations. We
especially acknowledge the detailed field and subsurface work of Alton Albin,
Mark Butler, Ibrahim emen, Kevin Lant, Lu Huafu, Kevin Lant, Jim
McDougall, Fred Nelligan, Rick Ricketts, Jill Schlaefer, Kermit Shields, and
Keith Whaley, which was the foundation on which this study was built.

1040

appearance of locally-derived clasts in the upper


Saugus Formation and its equivalents, and continues
today, along with the southwest-verging San
Cayetano fault farther west. Also active are northeast-verging backthrusts occurring in the east Ventura basin fold belt, and a segment of the San Gabriel
fault which now acts as a northeast-dipping obliqueslip reverse fault.
Northeast-trending discontinuities and structures
divide the present deformation zone into four segments. In the Hopper Canyon segment at the west
end of the area in the west Ventura basin, the San
Cayetano fault places Miocene Modelo Formation
over Pliocene-Pleistocene strata more than 5 km
thick, largely at maximum burial. To the southeast,
the Newhall-Potrero segment is characterized by
north-vergent backthrusts within the east Ventura
fold belt and by southward thrusting of the basin
sequence (tectonic inversion) over the structural
shelf on the Santa Susana fault. Farther southeast,
the Placerita segment is marked by reverse faulting
on both the Santa Susana fault and the San Gabriel
fault. Southeast of the basin in the San Fernando Valley, the Sylmar segment contains a thick PliocenePleistocene sequence overridden by basement rocks
of the San Gabriel Mountains as well as the southverging Mission HillsGranada Hills and Northridge
Hills fault zones.
INTRODUCTION
The Transverse Ranges of California cut across the
northwest grain of the Coast Ranges to the north and
the Peninsular Ranges to the south (Figure 1, inset).
In the center of the western Transverse Ranges lies a
major oil-producing province: the Ventura basin.
The basin is divided into the west Ventura basin,
which produces oil mainly from Pliocene and lower
Pleistocene strata, and the east Ventura basin, where
oil is produced mainly from upper Miocene and
lower Pliocene strata.
In the west Ventura basin, the valley of the Santa
Clara River and its extension offshore in the Santa
Barbara Channel contain one of the thickest sections
AAPG Bulletin, V. 78, No. 7 (July 1994), P. 10401074.

Figure 1Index map of the east Ventura basin and San Gabriel fault. Base map from Jennings (1975). Basement geology of San Gabriel Mountains from
Ehlig (1975). Abbreviations: AM, Alamo Mountain; CF, Canton fault; CR, Caliente Range; F, Fillmore; FM, Frazier Mountain; MC, Modelo Canyon; MF,
Morales fault; MG, Mission HillsGranada Hills fault; N, Newhall; NH, Northridge Hills fault; P, Piru; S, Saugus; SB, Sylmar basin; SGF, San Gabriel fault.

Yeats et al.
1041

1042

East Ventura Basin

of Pliocene-Pleistocene strata in the world. This section is bounded by active reverse faults: the Red
Mountain fault and San Cayetano fault on the north,
and the Oak Ridge fault on the south. The Red
Mountain, San Cayetano, and Santa Susana faults are
part of a south-verging zone of reverse faults extending from the western Santa Barbara Channel to the
San Andreas fault near San Bernardino (Figure 1). In
the west Ventura basin, this Quaternary fault system
coincides with the northern edge of the thick
Pliocene-Pleistocene trough.
In the east Ventura basin east of the town of Piru
(Figures 1, 2), the south-verging Santa Susana fault is
south of the trough, not north of it, and this fault
marks a zone in which the trough sequence is thrust
southward over its own structural shelf (Yeats,
1979), a process known as tectonic inversion. The
trough sequence and its northern structural shelf are
deformed in a fold belt. Young reverse faults cut
these folds, but except for two (the Holser and Del
Valle faults), displacements are relatively small.
The east Ventura basin is crossed by the San
Gabriel right-slip fault, a former strand of the San
Andreas system. Upper Miocene strata of the east Ventura basin cannot be correlated across the San Gabriel
fault because coeval strata have been displaced by
strike slip. In contrast, the Pliocene and Quaternary
beds are correlated without difficulty across the fault,
indicating that most strike slip preceded deposition of
these strata. Their deformation is principally by
reverse faulting and folding. The east Ventura basin,
therefore, contains evidence critical to the timing of
the shift from strike slip to contractile deformation as
the Transverse Ranges were formed.
This paper summarizes the surface and subsurface
geology of the east Ventura basin. A preliminary summary of all but the western part of the basin was published by Stitt (1986). We compiled the surface geology at 1:24,000 scale from published literature and
unpublished student theses, remapping the geology
where it was incompatible with subsurface well data.
A simplified version of the surface geology is shown
as Figure 2. For our 1:24,000 maps and subsurface
studies, see Ricketts and Whaley (1975), emen
(1977), Lant (1977), Shields (1977), Yeats et al. (1977,
1985), Nelligan (1978), Stitt (1980), Huftile (1988),
and Huftile and Yeats (in press). We relied on Mefferd
(1965) and Cordova (1966) for the subsurface geology of the Newhall-Potrero and Castaic Junction oil
fields, respectively. Surface geologic mapping of the
southeastern Ventura basin (Winterer and Durham,
1962), the northern margin of the east Ventura basin
(J. Crowell, unpublished map), the northwest part of
the Fillmore quadrangle (Dibblee, 1990), the San
Gabriel fault zone (Weber, 1982), the Newhall quadrangle (Smith, 1984; Treiman, 1986), the Mint Canyon
quadrangle (Saul and Wootton, 1983), and the Devils
Heart Peak and Cobblestone Mountain quadrangles

(T. W. Dibblee, Jr., unpublished maps), was of particular value to the study. We based our subsurface geology on data from more than 2000 wells obtained from
the California Division of Oil and Gas and from oil
operators. We did not do our own biostratigraphy but
instead relied on industry biostratigraphic zonation,
which is based on benthic foraminiferal zones of
Kleinpell (1938) and Natland (1952) supplemented
by planktonic microfossils. For an up-to-date discussion of microfossil zonation and the ages of benthic
foraminiferal zones, see Blake (1991). Seismic lines
available to us were not of high enough quality to contribute significantly to the study.
STRATIGRAPHY
The east Ventura basin formed at least in part on
crystalline rocks correlated to exposures in the San
Gabriel Mountains to the east and the Alamo Mountain region to the north. Unmetamorphosed Paleogene strata unconformably overlie the crystalline
rocks, but these were deposited in a basin framework unrelated to the east Ventura basin. The east
Ventura basin itself began with deposition of middle
Miocene strata in a trough extending southeast from
the Topatopa Mountains across the Santa Susana
Mountains and ending west of the San Fernando Valley. The Miocene and earliest Pliocene basin ends at
the San Gabriel fault, whereas younger strata are preserved on both sides of the fault.
Basement Rocks
Gneiss crops out in a narrow sliver between the
San Gabriel fault on the east and the Canton fault on
the west; gneiss is also found in the Conoco Alexander well to the southeast (Figure 3, well 3; Table 1).
The oldest strata deposited on the gneiss are Modelo
Formation of late Miocene (Mohnian) age. The gneiss
is correlated with the Mendenhall Gneiss of the western San Gabriel Mountains (Oakeshott, 1958; Ehlig,
1981) and the Mendenhall Gneiss of the Alamo
Mountain area west of the Ridge basin (Ehlig and
Crowell, 1982). Zircons from layered gneiss in the
western San Gabriel Mountains adjacent to the
Soledad basin are dated as 1715 30 Ma and 1670 20
Ma (Silver, 1966), and correlative granodioritic augen
gneiss west of the Ridge basin is dated as about 1655
Ma (L. T. Silver, in Frizzell and Powell, 1982).
West of the Canton fault, the basement consists of
the Whitaker Granodiorite of Shepherd (1960), which
is in part thrust over Paleogene strata along the
Whitaker thrust. Well data document a subsurface
ridge of granodiorite and granite close to and southwest of the San Gabriel fault overlain by Mohnian strata (Figure 3). The southeastern part of this ridge

Towsley Fm.

Tt

Modelo Fm.,
Upper Miocene
Marine
Middle
Tmm
Miocene

Fernando Fm.
Rincon Fm.

Marine Eocene
Basement

Te
gr

Tvs Vaqueros-Sespe

Tr

Tmu

Tf

Tt

5 KM

DEL

SUSA

Contact

SANTA
SUSANA

Qs

OAT
MTN.

Tmu

NEWHALL

Tc

Qs

Saugus Fm.

Qs

FAU
LT

Tmm

Tm
u

Tf

EL

RI

Tt

Te

Tmc

Qs

Tf

Qs
Qs

Tc

Tf

gr

Tf

Tmc

MINT
CANYON

11830'W

UL
T

FA

Qs

Tc

Mint Canyon Fm.

Castaic Fm.

AB

Tmc

Tc

Tvb Violin Breccia

SA

Tf

(East of San
Gabriel Fault)
Tc
Tvb

NA

LT

Reverse fault

Fault

Tt

Tf

Qs

FAULT
VALLE F
AU

HOLSER

TA

SAN

Tmu

Tt

SF

WHITAKER PEAK
VAL VERDE

Tt

Tmu

- Name of 7 1/2'
quadrangle

T.

FL

Tmm

11845'W

STRAND

(West of San
Gabriel Fault)

FAU

LT

PIRU

STRAND

Tmu

PIRU

UL

FA
COBBLESTONE
MTN.

CA

AN

Tr-Tvs

Te

PIRU

KE

gr
TA

HI

Figure 2Geologic map of the east Ventura basin, compiled from 1:24,000-scale mapping cited in the text. Location of map is on Figure 1.
Abbreviations: CF, Canton fault; DCF, Devil Canyon fault; SSF, Santa Felicia fault; WCF, Whitney Canyon fault.

Tmu

RIDGE
geology not shown

Tf

AK

Qs

IN

MA

UA

CAYETANO

Tmu

Tm 2
1

Tm

Tmm

BL

Tr-Tvs

FLT.

AG

Saugus Fm.

Te
MTN.

Tmc

Qs

SA

FILLMORE

Tr

YNEZ

NE

PI

34
22.5'N

34
30'N

DEVILS
HEART
PEAK

Tvs

FLT.

SANTA

Tr Tmm

CF DC

WCF

Tvs

Yeats et al.
1043

1044

East Ventura Basin

Figure 3Paleogeologic map of the base of Mohnian strata, approximately equivalent to the base of the upper
Miocene, east Ventura basin. There are two options for a pre-upper Miocene (pre-Mohnian) fault bounding the
basement ridge on the southwest. Option (A) curves the fault to the east, separating basement rocks on the north
from a thick Eocene and Sespe(?) sequence to the south. Option (B) continues the fault southeast between the
Argosy-Lassalle and Celeron-Towsley wells and to the Texaco 1-A Evans well, containing breccia possibly derived
from a pre-Mohnian fault scarp. Formation symbols: gn, gneiss; gr, granite; K, Upper Cretaceous marine strata; Te,
Paleocene and Eocene marine strata; Ts, Sespe Formation; Tv, Vaqueros Formation; Tr, Rincon Shale; Ttp, Topanga
Formation; Tmm, Modelo Formation with middle Miocene (Relizian and Luisian) microfossils. In footwall block of
Santa Susana fault, the paleogeologic map is at the base of the middle Miocene (Luisian) Modelo Formation. CF,
Canton fault; DCF, Devil Canyon fault; WCF, Whitney Canyon fault. Exxon D-1 NL&F and Texaco A-1 Newhall wells
contain breccia in the upper Mint Canyon Formation east of the San Gabriel fault.

Yeats et al.

1045

Table 1. Wells Penetrating Basement, East Ventura Basin


Map
No.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17

Well Name

Rock Description

Conoco Vier Kenny*


Union Alexander*
Conoco Alexander*

biotite granite and biotite granodiorite, cataclastically deformed


biotite granite, partially altered to chlorite and clays, followed by cataclasis
hornblende gneiss, sphene-bearing hornblende-biotite gneiss, epidotehornblende gneiss, cataclastic texture
Conoco 2 Rynne Fisher
biotite granodiorite partially altered to epidote and chlorite
Texaco A-38 Honor
granodiorite, slickensides and fractures, associated with contorted and
Rancho (NCT-2)
metamorphosed sediments
Texaco 20 Wayside
granite, pink, coarse grained, somewhat gneissic; associated ultramafic dike
Texaco 10 Wayside
biotite granite, coarse grained, some schistosity, slightly metamorphosed
Thomas B-3 NL&F
biotite schist
Unocal 2 NL&F*
muscovite granite and granodiorite with subsidiary hornblende; cataclastic
texture (original hole). Granite gneiss, biotite streaks (redrill)
Texaco E-1 Newhall
biotite granite
Superior 14-23 Bonelli*
biotite-muscovite schist; andalusite-sillimanite schist; granite, sheared
Conoco Georgina Swanson biotite granite, white
Mobil H&M*
chlorite schist, biotite-actinolite schist
Mobil J-2 Circle*
biotite granodiorite, some alteration to chlorite and epidote, fractured
Mobil Bermite*
biotite granodiorite, some alteration to chlorite, epidote, and calcite;
cataclastically deformed
Conoco Phillips
schist, dark green to black with thin beds of recrystallized limestone;
fractures and slickensides
Mobil Macson Mission
hornblende-plagioclase-quartz gneiss with subordinate amounts of marble

*Wells where basement rocks were examined in thin section.

includes biotite-muscovite schist, andalusite-sillimanite


schist, chlorite schist, and biotite-actinolite schist in
addition to granitic rocks (Table 1), suggesting a correlation with the Placerita Formation of Oakeshott
(1958) in the western San Gabriel Mountains south of
the San Gabriel fault. Schist with recrystallized limestone in the Conoco Phillips well (Figure 3, well 16)
and gneiss with marble in the Mobil Macson Mission
well (Figure 3, well 17) are also correlated to the
Placerita Formation. The basement rocks in the Macson Mission well are unconformably overlain by middle Miocene Topanga Formation. The basement in the
Phillips well is found beneath probable Paleogene strata, but the contact may be the Whitney Canyon fault.
Southwest of the subsurface basement ridge, the
basement beneath the east Ventura basin is not
exposed and is not reached by wells. The oldest
known rocks southwest of the Santa Susana fault are
of Late Cretaceous age; these can be correlated with
strata in the Santa Monica Mountains that unconformably overlie granitic plutons that intrude the
Jurassic Santa Monica Formation.
Sub-Middle Miocene Strata
Pre-basinal rocks consist of four roughly coeval
sequences which are shown as separate stratigraphic columns in Figure 4.

The sequence south of the Santa Ynez fault can be


traced westward through the Topatopa Mountains
and Santa Ynez Mountains (Figure 1). Upper Cretaceous deep-water strata are overlain disconformably
by an Eocene sequence 2500 m thick including the
Juncal Formation, Matilija Sandstone, Cozy Dell
Shale, and Coldwater Sandstone (Vedder et al., 1969;
Dibblee, 1982) (Figure 4). In the Sespe Creek area,
the marine Eocene sequence is overlain by nonmarine Sespe Formation of Eocene-Oligocene age, shallow-marine Vaqueros Sandstone of Oligocene age,
and Rincon Shale of early Miocene (Saucesian) age
(emen, 1977, 1989).
On the north side of the basin in the Piru Creek
area (Figure 1), upper Paleocene to upper Eocene
conglomerate, sandstone, and shale rest unconformably on granitic basement (Kriz, 1947; Nilsen
and Clarke, 1975). At the east end of the outcrop
belt, middle and upper Eocene strata (Shepherd,
1960; Squires, 1977) are in fault contact with
granitic basement along the Whitaker fault, which
Shepherd (1960) and J. C. Crowell (unpublished
mapping) regarded as an overturned, sheared
unconformity. In Canton Canyon, east of Piru Creek,
the marine Eocene strata are overlain by nonmarine
Sespe Formation (Figure 4), which contains in its
lower part conglomerate and breccia derived in part
from a distinctive source terrain in the western San
Gabriel Mountains (Crowell, 1954; Bohannon, 1975,

1046

East Ventura Basin

1976). The Sespe is overlain by the Vaqueros Sandstone, which is more conglomeratic toward the eastern end of its outcrop belt, and by Rincon Shale (Figure 4), a brown silty mudstone with a lens of pebbly
sandstone (Yeats et al., 1985). North of the Santa
Clara Valley, formations older than the Sespe are not
encountered in wells. The Sespe, Vaqueros, and Rincon formations are found in the Union 1 Moran well,
and the Rincon is found in the Gulf 1 Hathaway well
(both located on Figure 3).
The well-known sedimentary section of the Simi
Hills and Simi Valley has been traced to the footwall
of the Santa Susana fault (Yeats, 1979; Seedorf,
1983). A thickness of 2000 m of Upper Cretaceous
deep-water strata, largely turbidites, is found in the
Simi Hills, with the base not exposed. The nonmarine Paleocene Calabasas Formation of the western
Simi Valley is overlapped eastward by the PaleoceneEocene Santa Susana Formation, which is predominantly deep-water mudstone with a basal conglomerate (Yeats, 1987b). This formation is overlain
disconformably by the marine Eocene Llajas Formation, which also contains a basal conglomerate, by
the Sespe Formation of Eocene-Oligocene age, and
by the Vaqueros Sandstone. The Rincon Shale of the
northern two sections appears to grade laterally into
the Vaqueros in the Santa Susana footwall block.
North of the Santa Susana fault, the Continental 1
Phillips well in the Placerita oil field (Figure 3, well
16), penetrated approximately 1700 m of Paleocene
and Eocene marine strata in apparent fault contact
with metamorphic rocks (Stitt, 1986). Seedorf
(1983) correlated the lower part of this sequence, a
conglomerate overlain by claystone, siltstone, and
minor sandstone, with the Santa Susana Formation,
and the upper part, a conglomerate overlain by siltstone and sandstone, with the Llajas Formation.
Between the Aliso Canyon and Placerita oil fields,
upper Miocene marine strata are underlain unconformably by nonmarine beds known only from the
subsurface; these rest on marine Eocene rocks (Winterer and Durham, 1962; Nelligan, 1978) (for map
extent, see Figure 3). The sequence consists of gray,
red, green, and bluish shale, claystone, siltstone,
sandstone, and conglomerate, with local interbeds
of blue to blue-gray bentonite. The sequence is 260
m thick in the Morton & Dolley 5 Needham well
(Figure 3), where it rests on marine Eocene rock.
Elsewhere, where the marine Eocene is not reached,
the nonmarine sequence is at least 610 m thick in
the British American 1 Edwina well and possibly at
least 1100 m thick in the Basenberg 1 Hamilton well
(both located on Figure 3), where, however, part of
the section may be repeated by a fault.
Winterer and Durham (1962) questionably assigned
the sequence to the Sespe Formation based on its stratigraphic position. The Sespe of the Santa Susana footwall block is overlapped by the Modelo Formation

1013 km west of these strata (Figure 3) and is not laterally continuous with the nonmarine sequence north
of the Santa Susana fault. The Topanga Formation of
the San Fernando Valley to the east (Oakeshott, 1958;
Shields, 1977) and Oat Mountain anticline to the south
(Saul, 1975; this paper) is much coarser grained and is
interbedded with basalt. The middle Miocene strata
beneath the Pico anticline as well as those closer to the
Santa Susana fault are conformable with the overlying
Modelo Formation, whereas the unnamed nonmarine
sequence underlies the Modelo with angular unconformity. We conclude that these nonmarine strata are
most likely the same age as the Sespe Formation or the
Vasquez Formation of the Soledad basin, and they may
have been deposited in a separate, fault-bounded basin
similar to others described by Bohannon (1975).
Middle Miocene Sequence
The east-dipping strata overlying the Rincon Shale
in the hills between Sespe Creek and Piru Creek
include a lower shale member, a lower sandstone
member, a middle shale member, an upper sandstone member, and an upper shale member (Figures
5; 6A, B). Eldridge and Arnold (1907) named the
Modelo Formation for a sequence including the two
sandstone members and the upper two shale members in Modelo Canyon. They included the lowermost shale member in the Vaqueros Formation. Kew
(1924) described as Modelo Formation a thick
sequence of sandstone and shale between the Vaqueros Sandstone and the overlying Pico Formation of
his Fernando Group. Kew (1924) included the Rincon Shale in the lower shale member of his Modelo
Formation. Hudson and Craig (1929) also included
the Rincon Formation in the lower shale member,
but they preferred to call this shale member and the
overlying sandstone member and the middle shale
member the Topanga Formation, restricting the term
Modelo to the upper sandstone member and the
uppermost shale member. Bramlette (1946) and Dibblee (1989) referred to the Miocene strata above the
Rincon Shale as the Monterey Formation. We follow
emen (1977, 1989) in restricting the Modelo of
Eldridge and Arnold (1907) to those strata between
the Rincon Shale and the overlying Towsley Formation of Winterer and Durham (1962) of latest
Mioceneearliest Pliocene age, discussed below. This
results in a five-part subdivision of the Modelo Formation: Tm1, shale; Tm2, sandstone; Tm3, shale; Tm4,
sandstone; Tm5, shale (Figure 6A, B).
emen (1977) showed that the lower Modelo
shale, as restricted by Bramlette (1946), and the
lower sandstone contain benthic foraminifera of the
Relizian and Luisian stages of Kleinpell (1938),
which range in age from 17.4 to 13.9 Ma (Yeats et
al., 1989; Blake, 1991). The lower shale is in part

Yeats et al.

1047

Figure 4Stratigraphic columns of the submiddle Miocene Tertiary strata in the east
Ventura basin region.
These strata were
deposited in a Paleogene basin unrelated
to the Neogene east
Ventura basin. For
Neogene stratigraphic
columns, see Figure 5.

dark gray to brown, thin bedded to laminated, cherty to porcelaneous, and in part calcareous to
diatomaceous; these are lithologies characteristic of
the Monterey Formation to the west (Dibblee,
1989). The lower sandstone is locally interbedded
with dark brown siltstone and claystone. This twopart subdivision of middle Miocene strata is recognized throughout the Piru 712' quadrangle (emen,
1977; Huftile and Yeats, (in press), although the
lower sandstone lenses out southward 5 km northnortheast of Fillmore (emen, 1977, 1989; Dibblee,
1990) (Figure 2). To the north, in the Cobblestone
Mountain quadrangle, the lower sandstone is much
thicker in the high country between the Sespe Creek
and Piru Creek drainages, and is thinner again close
to the Pine Mountain fault (Dibblee, 1989; T. W. Dibblee, Jr., unpublished map) (Figure 2). The outcrop
pattern (Figure 2; cf. Figure 3 of Dibblee, 1989) cuts
across the axis of a depocenter south of the Pine
Mountain fault.
The five-part subdivision of the Modelo Formation

cannot be mapped in the subsurface east of Piru


Creek (Figure 6CE). Strata with Luisian and Relizian
foraminifera are 1480 m thick in the Gulf 1 Hathaway
well (Figure 6C). South of the Santa Clara River in the
hanging-wall block of the Santa Susana fault, sandstone and conglomerate overlain by an amygdaloidal
basalt flow 3060 m thick were called Topanga(?) Formation by Winterer and Durham (1962) (Figure 6E,
F), who included overlying shale containing Luisian
microfossils in the Modelo Formation. Saul (1975)
mapped Topanga Formation in the core of the Oat
Mountain anticline at Aliso Canyon oil field, placing
the contact with the overlying Modelo Formation
(Figure 5) at the top of the predominantly sandstone
unit with intercalated shale containing Luisian microfossils. The sandstone is interbedded with conglomerate and with greenish-gray siltstone, suggesting that
the Topanga there is grading to shallow marine. The
basalt in the Topanga Formation is dated by the potassium-argon method as 15.1 Ma (Turner, 1970, revised
by Weigand, 1982). West of Oat Mountain anticline,

1048

East Ventura Basin

Figure 5Neogene
stratigraphic
columns, east Ventura basin. Left column: local biostratigraphic units
correlated to time
scale. Right three
columns show representative Neogene
sections; thicknesses
are highly variable.
For Paleogene stratigraphic columns,
see Figure 4.

the middle Miocene consists of greenish-gray, micaceous, poorly sorted sandstone interbedded with
black to dark-brown siltstone. The base of the middle
Miocene is cut out by the Santa Susana fault (Figure
6DF); minimum thicknesses are 180 m in Oakridge
oil field and 762 m in Aliso Canyon oil field (both
fields located on Figure 7). Only the Mobil 1 Macson
Mission well (Figure 3, well 17) penetrates the base of
the middle Miocene; there, the sequence consists of
335 m of biotitic, locally carbonaceous gray siltstone
and sandstone resting on sheared gneiss.
Down dip from the Santa Susana fault, the middle
Miocene sequence is at least 1125 m thick in the
Celeron-Chevron 1 Towsley well, a deep test in the
Pico anticline (Figures 3, 6E). Two additional deep
tests, the Exxon 78 NLF well at Castaic Junction oil
field and the Sun A-1 RSF well at Newhall-Potrero oil
field (Figure 6D), may have reached the middle
Miocene, but faunal evidence is not available. In all
three wells, the lower part of the section is dominated by sandstone, as is the Topanga at Oat Mountain
anticline, and sub-Miocene rocks were not reached.
Farther northeast, upper Miocene strata overlap the
middle Miocene and rest directly on basement rocks.

The Santa Susana fault dies out at the surface near


the Oakridge oil field, and it is possible to correlate
electric-log markers in the Modelo across the fault
(Ricketts and Whaley, 1975; Yeats 1987a). On the
footwall side of the fault, the Modelo includes both
Luisian and Mohnian microfaunas. The Luisian Modelo consists of gray to brown, fine- to very finegrained sandstone interbedded with gray to brown
siltstone and shale, and it is overlain by light brown
to black silty shale and sandy siltstone. At Aliso
Canyon oil field, the Sesnon producing zone of the
Modelo Formation consists of very fine-grained to
fine-grained sandstone with worm impressions,
fish scales, and uncommon impressions of megafossils. In contrast to the thick middle Miocene
sequence north of the Santa Susana fault, the Modelo
south of the fault is relatively thin, comprising a
structural shelf. The Luisian Modelo rests with angular unconformity on the Vaqueros at the northwest
end of the Santa Susana fault and on Paleocene and
Cretaceous strata at the southeast end of the fault
(Figure 3). The contact between Luisian and Mohnian strata is conformable south of the fault and probably north of the fault as well.

Yeats et al.

1049

(A)
Modelo lobe and east end of west Ventura basin (Hopper Canyon segment).

(B)

East end of Oak Ridge and San Cayetano faults near Piru Creek (Hopper Canyon segment).

Figure 6Cross sections, east Ventura basin, no vertical exaggeration. Well numbers are identified in the
Appendix. Symbols: gn, gneiss; gr, granite; K, Cretaceous marine strata; Te, Paleocene and Eocene marine strata; Ts,
Sespe Formation; Tv, Vaqueros Formation; Tr, Rincon Shale; Ttp, Topanga Formation; Tm, Modelo Formation, subdivided in (A) and (6) into Tm1 (lower shale), Tm2 (lower sandstone), Tm3 (middle shale), Tm4 (upper sandstone),
and Tm5 (upper shale); Tmc, Mint Canyon Formation; Tc, Castaic Formation; Tvb, Violin Breccia; Tt, Towsley Formation; Tf, Fernando Formation; Qs, Saugus Formation; Qp, Pacoima Formation. Shading shows oil-producing strata. Bedding dips are shown by short lines at the surface and by lines extending from the well course (dipmeter).
Core dips are shown as bidirectional from the well course because dip direction is not known. Lines of section are
located on Figure 7.

Figure 6Continued.

Santa Susana fault to San Gabriel fault through Newhall-Potrero and Castaic Junction oil fields (Newhall-Potrero segment).

(D)

East Ventura basin fold and thrust belt (Newhall-Potrero segment).

(C)

1050
East Ventura Basin

Yeats et al.

1051

(E)
Santa Susana fault to San Gabriel fault through the Pico anticline (Placerita segment).

(F)
Eastern end of east Ventura basin (Placerita segment).
Figure 6Continued.

Upper Miocene Modelo Formation


We include the middle shale, upper sandstone,
and upper shale of the Modelo Canyon section in the
Mohnian (upper Miocene) part of the Modelo Formation (Figure 5). The Mohnian is 6.5 0.3 to 13.9
Ma in age in the Los Angeles basin (Blake, 1991);
Yeats et al. (1989) considered the base of the Mohni-

an to be 13.8 Ma in the Cuyama basin in the southern Coast Ranges. Dibblee (1989, 1990, 1991) correlated the upper shale to the Sisquoc Formation of
predominantly Delmontian age based on its silty or
clay shale lithology, similar to Sisquoc exposures
west of Sespe Creek and different from the LuisianMohnian siliceous shale of his Monterey Formation.
However, the upper shale contains Mohnian micro-

1052

East Ventura Basin

11845'W
SA

WHITAKER
PEAK

COBBLESTONE
MTN.

11830'W

VAL VERDE

HR

LT

TA

CH

T
FAUL

PI

NO

SA

UL

FA

PI

YE

HL

RA

DV

CJ

LT

6D
OR

LC

OF

TR
SS

5 KM

NP

TN

FI

EU

DW

FS

ET

PC

WH

EL
TW

ST

WI
RC

SANTA

OM

SU

SA

NA

SANTA
SUSANA

PL
TO

TU

FIG

TY

3422.5'N

FAU

FIG
. 6
E

CA

OAK RIDGE

MINT CANYON

WA

FA

HS

HP

NEWHALL

TC

TE

HOLSER

SAN

6C

G.

FI

OC

. 6
F

FIG. 6B

FIG. 6A

PIRU

EL
RI

3430'N

AB
G

CC

AC

OAT MTN.

fossils, not Delmontian, as does the Sisquoc Formation. In addition, in the Santa Felicia and Holser synclines west of Piru Creek (Figure 6B), the upper
shale is overlain by the Hasley conglomerate, the
basal member of the Towsley Formation, which is
itself characterized by Delmontian microfossils.
Isopachs of the Mohnian part of the Modelo Formation (Figure 8) outline a linear, northwest-deepening depocenter beneath the Pico anticline northeast of the Santa Susana fault. In the center of this
trough, the Mohnian Modelo is 2550 m thick in the
Celeron Towsley well and 1750 m thick in the Mobil
Macson Mission well farther southeast. The Mohnian
decreases in thickness southwestward to less than
700 m at the outcrop in the hanging-wall block of
the Santa Susana fault as it approaches the structural
shelf in the footwall block (Figure 6E, F). Northeast
of the Pico anticline, Mohnian thickness decreases
to zero, with abrupt decreases at coeval normal
faults (Figure 8). North of the Santa Clara River, the
Mohnian is more than 3000 m thick in the Gulf Hathaway well and adjacent outcrop. Thickness decreases eastward to 1350 m in the Oak Canyon oil field
(located on Figure 7), part of a broad structural shelf
extending east almost to Honor Rancho oil field.
Thickness decreases abruptly across the Devil
Canyon fault and gradually southeastward to

FAULT

Figure 7Oil fields


of the east Ventura
basin. Black indicates middle and
upper Miocene Modelo production; dark
shading indicates
Modelo (upper
Miocene) and
Towsley production;
light shading indicates Fernando production. Other symbols are identified
on Figure 2 and in
Table 2. Unpatterned oil fields produce from the Oak
RidgeSimi Hills
block beneath the
Santa Susana fault.
AC, Aliso Canyon;
OM, Oat Mountain;
OR, Oakridge; SS,
Santa Susana; ST,
South Tapo Canyon;
TY, Torrey Canyon.
Lines of cross sections are shown for
Figure 6.

Temescal oil field. Thicknesses decrease as the San


Gabriel fault is approached, but isopachs are generally not parallel to the fault.
Winterer and Durham (1962) pointed out that
sandstone in the Modelo is characterized by repeated graded bedding and was deposited by turbidity
currents. The sandstones are increasingly conglomeratic toward the San Gabriel fault to the northeast
and were deposited as submarine fans (Dibblee,
1989). Submarine fans characterize most of the Modelo except for the southern and eastern margins of
the basin, where the Modelo is largely characterized
by shale. The facies boundary trends west-northwest, parallel to the direction of thinning of the
Modelo toward the structural shelf in the footwall
block of the Santa Susana fault. Farther east, the
facies boundary trends north, parallel to Modelo
isopachs west of the Placerita oil field.
Towsley Formation
A sequence of light-colored sandstone, conglomerate, and brown-weathering mudstone on the north
slope of the Santa Susana Mountains (located on Figure 9), previously mapped as Modelo by Kew (1924),
was named the Towsley Formation by Winterer and

Yeats et al.

1053

Figure 8Isopachs of the upper Miocene Mohnian Modelo Formation (in meters). Dots show subsurface well control. Shaded pattern shows Mohnian Modelo exposures. Syndepositional normal faults are shown with tick marks
on downthrown side. DCF, Devil Canyon fault; HR, Honor Rancho oil field; PA, Pico anticline.

Durham (1954, 1962). They mapped the top of the


Towsley at the base of the first sandstone and conglomerate underlying the lowest concretion-bearing
olive-gray siltstone of the Fernando Formation. The
sandstone contains large, rounded brown concretions in which bedding can be recognized (Winterer
and Durham, 1962). Winterer and Durham (1962)
pointed out the difficulty of mapping the base and
top of the Towsley away from the type locality in
Towsley Canyon (located on Figure 9), and they
noted that these contacts, as defined on lithology, are
time transgressive. However, north of the Santa Clara
River, the base of the Towsley is marked by the deepwater Hasley conglomerate, which facilitates map-

ping the Towsley-Modelo contact there.


The Towsley at its type locality is characterized by
benthic foraminifera of the Delmontian Stage of
Kleinpell (1938). The Delmontian is not a true stage;
Delmontian faunas in southern California are not
time equivalent to the fauna at the type Delmontian
section in the central Coast Range (Pierce, 1972; Barron, 1976). The age of the local Delmontian of
southern California is estimated as 4.98 0.15 to 6.5
0.3 Ma (Blake, 1991). We map the base of the
Towsley at the base of the Hasley conglomerate,
although some samples from the Hasley yield Mohnian faunas. The top of the Towsley is the faunal
boundary between Delmontian and Repettian micro-

1054

East Ventura Basin

Figure 9Isopachs of the Towsley Formation (in meters). Symbols are the same as in Figure 8. TC, Towsley
Canyon; WCF, Whitney Canyon fault.

fossils, which generally corresponds to the color


change between brown mudstone below and olivegray mudstone above.
The repeated graded bedding, displaced foraminiferal faunas, broken megafossils, and abrupt facies
changes in Towsley sandstone indicate deposition by
turbidity currents (Winterer and Durham, 1962). The
sand bodies comprise large submarine fans that thicken and become more conglomeratic toward the San
Gabriel fault (Crowell, 1954). Some of the pebbly
sandstones are lenticular and channelized, and are
similar to the suprafan lobe facies of Walker (1978,
his Figure 18).
Near the Placerita oil field, the Towsley overlaps
the Modelo and rests on the Sespe(?) Formation,
Eocene strata, and basement rocks (Figures 3, 6F).
The strata change facies to shallow-marine deposits

that are difficult to distinguish from the overlying


Fernando Formation, leading Nelligan (1978) to map
them as a single unit.
Isopachs of the Towsley Formation (Figure 9)
delineate three zones of reduced thickness trending
northeast, normal to the San Gabriel fault. In detail,
isopachs within two of these zones and a zone of
thicker deposition between them trend east, suggesting that thickness changes are related to coeval
displacement on the San Gabriel fault. Farther southwest, Towsley thickness increases toward the outcrop belt in the Santa Susana Mountains (Figure 9),
indicating that the Towsley depocenter, close to the
present surface trace of the Santa Susana fault, has in
part been eroded away. However, there is no
Towsley south of the fault; the Fernando Formation
rests directly on Modelo Formation.

Yeats et al.

Miocene Strata East of San Gabriel Fault


Miocene rocks in the Soledad basin east of the San
Gabriel fault, including strata mapped as Towsley
Formation by Winterer and Durham (1962), differ
greatly from the Modelo and Towsley formations
west of the fault. Nonmarine units include the
Vasquez Formation of Oligocene(?) to early
Miocene(?) age (Muehlberger, 1958; Bohannon,
1975; Woodburne, 1975), the Tick Canyon Formation of late early to early middle Miocene age (Jahns,
1940), and the Mint Canyon Formation of middle to
late Miocene age (Winterer and Durham, 1962; Ehlig
et al., 1975; Ehlert, 1982) (Figure 5). Vertebrate fossils discussed by Winterer and Durham (1962) and
radiometrically-dated tuff beds (Terres and
Luyendyk, 1985) indicate that the Mint Canyon is
coeval with much of the Modelo Formation, but one
does not grade into the other.
The Mint Canyon Formation is overlain by and
locally interbedded with the marine Castaic Formation (Crowell, 1954), which comprises the lowest
formation of the Ridge basin (Crowell, 1975; Crowell and Link, 1982; Stitt, 1982). The Castaic contains
microfossils of the Mohnian and Delmontian stages,
and is thus the same age as parts of the Modelo and
Towsley formations. However, the microfossils
(Skolnick and Arnal, 1959; Schlaefer, 1978) and the
megafossils (Stanton, 1966, 1982) are different
across the San Gabriel fault. A possible connection
between the Castaic Formation and the east Ventura
basin is discussed by Paschall and Off (1961), Schlaefer (1978), McDougall (1982), and Stanton (1982).
The Castaic Formation grades abruptly southwestward into the Violin Breccia, which accumulated as
a talus at the base of the active San Gabriel fault, indicating that the San Gabriel fault marked the southwestern edge of the late Miocene Ridge basin. The
Violin Breccia is traced at least 1340 m southeastward into the subsurface to the Conoco 1 Alexander
well (Figure 3, well 3), but farther southeast, the
Castaic Formation abuts the San Gabriel fault without the intervening Violin Breccia.
Fernando Formation
Eldridge and Arnold (1907) described all strata
younger than the Modelo Formation and older than
terrace deposits as the Fernando Formation. Kew
(1924) raised the Fernando Formation to the Fernando Group, which consisted of a lower unit, the Pico
Formation, and an upper unit, the Saugus Formation.
Winterer and Durham (1962) did not use the term
Fernando, and their Pico Formation included strata
which Kew had earlier assigned to the Saugus. In the
eastern part of the area, Oakeshott (1950) used the
term Repetto Formation, which elsewhere refers to

1055

strata with microfauna assigned to the Repettian


Stage of Natland (1952). In this paper, we follow the
current practice of the U. S. Geological Survey (cf.
stratigraphic nomenclature discussion in Jennings
and Strand, 1969) in assigning all strata between the
Towsley and Saugus formations to the Fernando Formation. This corresponds to the Pico Formation of
Winterer and Durham (1962). In the western part of
the east Ventura basin, the Fernando Formation consists of a lower member with Repettian microfossils
and an upper member with shallower water upper
Pliocene or Pico microfossils not easily correlated
to the Venturian or Wheelerian stages of Natland
(1952) in the west Ventura basin (Figure 5). The top
of the Fernando Formation near Newhall is about 2
Ma, based on magnetic stratigraphy (Levi et al.,
1986; Levi and Yeats, 1993), but the top is younger
westward and older eastward due to a facies change
to Saugus Formation from west to east (Figure 5).
In the western part of the basin, the Fernando
consists of olive-gray and medium bluish-gray siltstone with reddish-brown concretions interbedded
in its lower part with sandstone and conglomerate
characterized by repeated graded bedding (Winterer
and Durham, 1962; Yeats et al., 1985). North of the
Holser fault, the Fernando consists of pebbly sandstone interbedded with bioturbated, sandy siltstone
deposited in relatively shallow water; the facies
boundary trends approximately east (Figure 10). The
Fernando also grades to shallow-water deposits in
the eastern part of the basin, where it is commonly
difficult to separate from the underlying Towsley
Formation, especially in the subsurface (Nelligan,
1978). This facies boundary trends north to northwest (Figure 10). The facies boundary changes trend
at the San Gabriel fault southeast of the Honor Rancho oil field in an area characterized by a westwardthickening depocenter (Figure 10).
The Fernando is the oldest formation easily correlated on both sides of the San Gabriel fault, indicating that most strike slip preceded Fernando deposition. East of the fault, the Fernando is tilted
southwestward, eroded, and overlapped by the
Saugus Formation in such a way that the isopachs of
the Fernando are in part parallel to the San Gabriel
fault (Figure 10).
Farther away from the San Gabriel fault, the thickness of the Fernando increases markedly toward the
Santa Susana Mountains and the Oak Ridge fault (Figure 10). Part of the thickness increase is due to the
lateral change southwest from the Saugus to the Fernando, documented in surface exposures and in
wells (Winterer and Durham, 1962; Yeats et al.,
1985). Farther southwest, the thickness increases
from 900 to 2500 m due to coeval development of a
southwest-verging monocline at the site of the
younger Newhall-Potrero anticline (Figures 6D, 10;
see following discussion).

1056

East Ventura Basin

Figure 10Isopachs of the Fernando Formation (in meters). Symbols are the same as in Figure 8. Hachured lines
marks facies boundary between the shallow-water and the deep-water Fernando Formation. N-P anticline, NewhallPotrero anticline; WCF, Whitney Canyon fault.

Fernando thickness is also 2500 m in the Santa


Clara Valley, where it is preserved between the San
Cayetano and Oak Ridge faults (Figure 6A, B). This
thickness is representative of that in the west Ventura basin (Yeats, 1983, 1988).
Saugus Formation
The Fernando Formation grades upward and laterally to shallow-marine to nonmarine strata which
comprise the uppermost major formation of the east
Ventura basin. Hershey (1902) referred to these strata as the Saugus division, and Kew (1924) assigned
the Saugus-Formation to the upper part of his Fernando Group. The base of the Saugus consists of
green-gray and gray pebbly sandstone and green-gray
siltstone which may contain marine to brackishwater fossils (Winterer and Durham, 1962). In many
areas, the Saugus-Fernando contact is placed at the
top of the fossiliferous sequence, but in the Santa

Clara Valley between the Oak Ridge and San


Cayetano faults, the contact is placed between finegrained Fernando strata and the fossiliferous sandstone and conglomerate of the Saugus Formation
(Figure 6A). Conglomerate predominates in more
northerly exposures west of the San Gabriel fault.
Northeast of the Holser fault, Weber (1979) subdivided the Saugus on the basis of clast composition of
conglomerate: clasts of Paleocene gray-brown sandstone vs. clasts of schist. The Saugus also occurs
south of the Santa Susana fault (Yeats, 1979, 1987a).
The uppermost deformed nonmarine strata in the
Newhall area north of the Santa Susana Mountains
(between wells 18 and 22 on Figure 6D) (Treiman,
1986) and in the Horse Flats area near Aliso Canyon
oil field (left edge of Figure 6F) south of the Santa
Susana fault (Saul, 1975) show an abrupt change
from clasts of basement rocks upward to clasts locally derived from the Modelo and Towsley formations
of the Santa Susana Mountains. Near Newhall, the
locally-derived strata have been mapped by Treiman

Yeats et al.

(1986) as the Pacoima Formation of Oakeshott


(1958) because of their fanglomeratic nature. These
strata were mapped as terrace deposits by Winterer
and Durham (1962), whose mapping indicated that
little if any of the section has been removed by erosion. South of the Santa Susana fault, similar strata
were mapped by Saul (1975) as the upper part of the
Saugus Formation. These locally derived deposits
reflect the uplift and erosion of the Santa Susana
Mountains (Saul, 1975; Levi and Yeats, 1993).
Winterer and Durham (1962) described boulders
of crystalline rocks more than 6 m in diameter resting on fine-grained Modelo Formation at an altitude
of 870 m on ridge crests on the north side of the
Santa Susana Mountains near the Aliso Canyon oil
field (located to the right of well 12 on Figure 6F).
We found crystalline boulders and cobbles overlying
Modelo rocks at altitudes of 825 to 975 m in this
same area. Winterer and Durham (1962) concluded
that these are terrace deposits, but this is unlikely.
The uppermost Saugus is dominated by clasts from
the Santa Susana Mountains (Saul, 1975), indicating
that the Santa Susana Mountains were uplifted
before the end of Saugus deposition. How, then,
could the area receive gravels of crystalline provenance unless the Santa Susana Mountains first lost
most of their topographic relief to receive the clasts,
then were uplifted again? The clasts are more likely a
lag gravel of basal Saugus, which implies that the
Towsley and Fernando formations were eroded off
the top of the Santa Susana Mountains before the
Saugus was deposited. We consider it likely that this
occurred during a pre-Saugus stage of movement on
the Santa Susana fault. The Torrey fault in the footwall block underwent reverse displacement prior to
Saugus deposition (Yeats, 1979, 1987a) (Figure 6D),
and we view the Santa Susana fault as undergoing
displacement at the same time. This pre-Saugus
reverse faulting apparently did not lead to enough
uplift to shed locally derived stones into the basal
Saugus.
The age of the Saugus is Pleistocene, based on
several magnetostratigraphic sections described by
Levi et al., (1986) and Levi and Yeats (1993). One of
the sections contains the Bishop or Friant ash (Levi
et al., 1986; Levi and Yeats, 1993). The Bishop ash is
dated as 0.759 0.003 Ma (Pringle et al., 1992), and
the Friant ash is dated as 0.62 Ma (Sarna-Wojcicki et
al., 1991). The youngest paleomagnetically dated
strata were deposited during the Brunhes Chron and
are from 0.6 to 0.5 Ma.
Post-Saugus and Post-Pacoima Deposits
Alluvium, landslide deposits, and other undeformed or locally-deformed sediments are widespread
in the area, but are not described in this paper.

1057

STRUCTURE
The structure is discussed in four parts: (1) deposition of a rifted basin in the middle and late Miocene,
(2) displacement on the San Gabriel fault, (3) faulting
and folding during deposition of the Pliocene-Pleistocene sequence, and (4) post-Saugus deformation
which, for the most part, still continues today.
Miocene Rifted Basin
A thick sequence of middle and upper Miocene
strata occupies a southeast-trending basin extending
from outcrops in the Cobblestone Mountain and Piru
quadrangles (Figure 2) into the subsurface northeast
of the Santa Susana fault (Figure 3). The great thickening of strata in outcrop does not appear to involve the
lowest shale member (Tm1) of the Relizian-Luisian
(middle Miocene) Modelo Formation, but it clearly
involves the overlying sandstone member (Tm 2),
which forms prominent ledges in the western part of
the Cobblestone Mountain quadrangle (Figure 2). All
microfossils in the thick middle Miocene Topanga
Formation southeast of the Santa Clara River are
Luisian, and a basalt flow interbedded with the Topanga at Aliso Canyon is dated within the age range of the
Luisian (Turner, 1970; Blake, 1991). Therefore, the
rifted basin became active sometime after the beginning of the Relizian at 17.4 Ma, and it clearly was
active prior to 13.9 Ma, the beginning of the Mohnian.
Mohnian strata of the Modelo Formation also show
evidence of thickening in the trough (Figure 8).
The thick Miocene sequence exposed in the Cobblestone Mountain quadrangle is not obviously bounded by faults on the south and north, although the
southeast-trending Agua Blanca fault zone (Figure 2)
may be the northern boundary of the trough. The
Agua Blanca fault and associated folded beds as young
as early Miocene Rincon Shale to the north of this fault
are covered unconformably by the Mohnian Modelo
Formation (Yeats et al., 1985), indicating strong deformation along this zone in the middle Miocene. The
Devil Canyon fault cuts Mohnian strata and is itself
overlain by Mohnian strata (Figure 2). It may form part
of the southwestern boundary of the basement ridge
and the northeast boundary of the trough.
Southeast of the Santa Clara River, deep-test wells
in the Newhall-Potrero and Castaic Junction oil fields
(Figure 6D) and at the Pico anticline (Figure 6E) penetrated a much greater thickness of Modelo Formation than that exposed to the southwest and documented in the subsurface to the northeast. The cross
section in Figure 6E shows that the boundary
between thin Modelo underlain by Sespe(?) Formation and thick Modelo beneath the Pico anticline is
so abrupt that it must be a fault. Farther southeast,
the breccia in the Modelo Formation in the Texaco

1058

East Ventura Basin

1-A Evans well (Figure 6F, well 15) is probably a


talus deposit derived from a coeval fault scarp to the
northeast. Wells northeast of the 1-A Evans well penetrate thin Mohnian strata underlain by Sespe(?) Formation, whereas wells to the southwest penetrate
thick Mohnian underlain by Topanga Formation.
The trough shallows farther southeast on the
northeast-trending lateral ramp of the Santa Susana
fault at the edge of the San Fernando Valley (Figure
8). The Mobil 1 Macson Mission well (Figure 3, well
17) penetrated a relatively thin sequence of Topanga
Formation underlain by basement; however, the
Mohnian in this well is much thicker than it is to the
northeast or southwest (Figure 8).
On the footwall side of the Santa Susana fault,
Modelo Formation of middle and late Miocene age
(Luisian and Mohnian stages) is much thinner than
coeval strata cropping out on the hanging-wall side,
and no gradation between these two sequences is
seen in the direction of the fault (Figure 6CF). This
relation leads to the suggestion that the Santa Susana
fault reactivated an old normal fault that marked the
southwest margin of the Miocene trough.
The Oak Ridge reverse fault in the west Ventura
basin may have reactivated a Miocene normal fault
(Yeats, 1987b). East of Torrey Canyon oil field (located
on Figure 7), the south strand of the Oak Ridge fault
changes eastward from a south-dipping reverse fault to
a north-dipping normal fault (Yeats, 1979, 1987a) (Figure 6B) that can be traced in the subsurface as far east
as Oakridge oil field (Figure 7). There, the Oak Ridge
fault is overridden by the Santa Susana fault, and the
change in fault dip from south to north may be related
to compression across the Santa Susana fault. The Oak
Ridge fault probably continues southeast beneath the
Santa Susana Mountains (Figure 6C), deeper than well
control, where it may have served as the zone of weakness that guided the younger Santa Susana fault.
North of the Santa Clara River, the Towsley Formation shows no evidence of the rifted trough, indicating that rifting ended there prior to 6.5 Ma. However,
south of the river, the Towsley thickens to the southwest across the axis of the older rift, and is thickest in
the outcrop section in the Santa Susana Mountains
(Figure 9). There is no Towsley in the footwall block
of the Santa Susana fault, indicating that the thick
Towsley ended at a south-side-up fault near the
younger Santa Susana fault (Yeats, 1979), about the
same place as the Modelo rift boundary. Thus, the rift
influenced Towsley thicknesses, but it was much
more restricted areally, and its axis shifted southwest.
San Gabriel Fault
The San Gabriel fault strikes southeast from near
Frazier Mountain (FM on Figure 1) and forms the
southwestern edge of the Ridge basin (Figure 1). The

fault crosses the east Ventura basin, where it marks a


contrast in stratigraphy of Miocene strata east and
west of the fault. The fault changes character in the
east Ventura basin. From Honor Rancho oil field
northwestward, the fault is straight, strikes N40W,
and dips steeply northeast with normal stratigraphic
separation (Figure 6C, D). Southeast of Honor Rancho, the fault is curved in plan and convex southward. At Saugus oil field, its strike is N55W; at
Placerita oil field, its strike is N70W. Dip is steeply to
the northeast, but with reverse separation rather than
normal (Figure 6E, F). Farther east, in the San Gabriel
Mountains, the fault bifurcates into a northern strand,
which strikes due east and is entirely within basement
rocks of the San Gabriel Mountains, and a southern
strand (Vasquez Creek fault of Powell, 1993), which
merges with the Sierra Madre reverse-fault zone marking the south-facing range front of the San Gabriel
Mountains (Figure 1). The mapped trace of the fault,
including the northern strand, extends 130 km from
the Frazier Mountain area, where the fault is overlain
unconformably by strata of the Pliocene Hungry Valley Formation, southeast to the east end of the San
Gabriel Mountains, where both strands of the fault
terminate at and are offset by northeast-trending faults
near its junction with the San Andreas fault (Matti and
Morton, 1993) (Figure 1).
Crowell (1952) first suggested large-scale right-lateral displacement on the San Gabriel fault. Total
right-lateral displacement on the fault is controversial, ranging from 42 km (Powell, 1993) and 44 km
(Matti and Morton, 1993) to 60 km (Bohannon,
1975; Ehlig and Crowell, 1982). A review of the displacement history of the San Gabriel fault in the context of an overall reconstruction of southern California fault systems is provided by Powell (1993) and
Matti and Morton (1993).
We adopt a pre-late Miocene piercing-point offset
of 60 km across the fault (Figure 11) based on (1) offset of the Precambrian Mendenhall Gneiss and
anorthositic rocks from near Frazier Mountain (FM
on Figure 1) to the western San Gabriel Mountains
(Crowell, 1962; Ehlig and Crowell, 1982), (2) offset
of Paleocene strata from the Caliente Range (CR on
Figure 1) to the Paleocene San Francisquito Formation at the northern edge of the Soledad basin (Sage,
1973) (located on Figure 1), (3) offset of the
Oligocene to lower Miocene Simmler Formation in
the Caliente Range and the Plush Ranch Formation in
the Lockwood Valley area (Carman, 1964) (Figure 1)
to the Oligocene to lower Miocene Vasquez Formation in the Soledad basin (Bohannon, 1975, 1976)
(Figure 1), (4) offset of the Blue Rock fault south of
the Caliente Range, with evidence of Miocene movement, to the San Francisquito fault in the Soledad
basin (Bohannon, 1975; Ehlert, 1982) in order to
match Charlie Canyon Megabreccia in the Soledad
basin (which overlies the Vasquez Formation with

Yeats et al.

1059

Figure 11Age vs. strikeslip displacement on the


San Gabriel fault. On the
abscissa, R = Relizian, L =
Luisian, M = Mohnian,
D = local Delmontian
(all four are benthic
foraminiferal stages of
Kleinpell, 1936), F = Fernando, and SA = Saugus
Formation. Control for
displacement are, from
right to left: VF, Vasquez
Formation; MC, conglomerate in Mint
Canyon Formation; DCF,
Devil Canyon fault; DCC,
Devil Canyon Conglomerate; CA, Castaic Formation; HA, Hasley Conglomerate; FM, Fernando
displacements in San
Gabriel Mountains; FV,
Fernando displacements
in east Ventura basin; SA,
Saugus Formation. No
age or displacement constraints in the direction
of the arrows. Two estimates of slip rate are
shown. The high slip
rates were followed
northwest of Honor
Rancho oil field by low
slip rates of 1.3 mm/yr
or less than 1 mm/yr, as
suggested by Kahle
(1986). The low slip rates
continue to the present.

angular unconformity) with a probable source in the


La Panza Range in the southern Coast Ranges (Smith,
1977; Ehlig and Joseph, 1977; Joseph et al., 1982)
(Figure 1, inset), and (5) offset of the Miocene
Caliente Formation of the Caliente RangeLockwood
Valley region from the Mint Canyon Formation of the
Soledad basin (Ehlig et al., 1975; Ehlert, 1982).
The lower displacement of Powell (1993) and
Matti and Morton (1993) is based on (1) correlation
of the San Francisquito fault with the San Andreas
fault at Frazier Mountain, requiring a displacement
of 4045 km, (2) basement offsets of 2223 km
across the north strand of the San Gabriel fault in the
central San Gabriel Mountains (Ehlig, 1981), combined with lack of evidence for large-scale displacement on the south strand (Vasquez Creek fault),
(3) displacement of the left-lateral Malibu Coast fault
and its eastern continuation north of the San Gabriel
Valley to faults in the southeastern San Gabriel

Mountains, and (4) difficulty in integrating a 60-km


displacement into an overall scheme for southern
California fault restorations.
The offset of the depositional system of the Mint
Canyon Formation determined by Ehlig et al. (1975)
and Ehlert (1982) is the same as that for all older
units, indicating that strike slip on the San Gabriel
fault largely postdates deposition of the Mint Canyon
Formation. However, Ehlert (1982) proposed that
strike slip began during deposition of the upper part
of the Mint Canyon Formation, after the deposition
of the lower Mint Canyon and Caliente conglomerates that show evidence of 60 km displacement. Stitt
(1986) showed that the Mint Canyon Formation
thickens southwestward toward the San Gabriel
fault such that the basin configuration is that of a
half graben. The Texaco A-1 Newhall and Exxon D-1
NLF wells (located on Figure 3) contain sedimentary
breccia in the upper part of the Mint Canyon Forma-

1060

East Ventura Basin

tion close to the San Gabriel fault. The breccia


includes clasts of Mendenhall Gneiss(?), granite, and
schist, which could not have come from the subModelo granite ridge on the southwest side of the
San Gabriel fault, but could have come from the
Alamo MountainFrazier Mountain region to the
northwest, which contains all rock types found in
the breccia. The breccia is an older analog to the
Violin Breccia of the Ridge basin. In addition, the
depocenter for the coarse-clastic facies of the Mint
Canyon Formation shifted northward for the turbidite facies of the Castaic Formation, suggesting
that the entry point for these coarse clastic sediments was displaced northwestward along the San
Gabriel fault (Stitt, 1982, 1986), as predicted by the
conveyor-belt model of deposition along the coeval
San Gabriel fault proposed by Crowell (1982).
Vertebrate fossils from the Mint Canyon Formation are referred to the Clarendonian and Barstovian
Vertebrate stages, the ages of which range from
about 17 to 10 Ma (Berggren, 1969). Tuff beds in the
upper part of the Mint Canyon Formation yielded zircon fission-track ages of 10.1 0.8 Ma and 11.6 1.2
Ma (J. Obradovich and T. H. McCulloh in Terres and
Luyendyk, 1985). Thus the age for initiation of strike
slip on the San Gabriel fault is in the range 1012
Ma, as proposed by Crowell (1982).
The lower Mohnian Devil Canyon Conglomerate
in the Modelo Formation thickens, contains larger
clasts, and grades to sedimentary breccia northeastward toward the San Gabriel fault (Crowell, 1952,
1954). Castaic Formation of the same age, on the
other side of the fault in the Ridge basin, could not
have been a source for the Devil Canyon Conglomerate. The conglomerate includes boulders as large as
1.5 m in diameter of anorthosite, gabbro, norite, and
gneiss similar to the rocks exposed in the western
San Gabriel Mountains (Crowell, 1952, 1962, 1982).
Right slip of at least 35 km but possibly as much as
56 km is required to place the Devil Canyon Conglomerate next to its probable source (Figure 11).
The Castaic Formation of Mohnian and Delmontian age grades southwest to the Violin Breccia,
a narrow talus deposit adjacent to the San Gabriel
fault. Right slip of about 35 km and possibly as much
as 60 km is required to match clasts of gneiss in the
Violin Breccia with their most likely source in the
Alamo MountainFrazier Mountain area west of the
fault (Crowell, 1952, 1982). Powell (1993) suggested that the 35-km length of the Violin Breccia outcrop would approximate the offset accumulated
during the time of deposition of the breccia. The
breccia in the Mint Canyon Formation close to the
San Gabriel fault would add another 11 km to the offset, using this criterion.
The Canton fault is an early formed strand of the
San Gabriel fault that juxtaposes gneiss on the east
against Whitaker Granodiorite on the west, a mini-

mum displacement of 23 km. The Canton fault cuts


the basal part of the lower Mohnian Devil Canyon
Conglomerate and is overlain by the upper part of
the conglomerate, also Mohnian (Crowell, 1954;
Yeats et al., 1985; Stitt, 1986) (Figures 2, 3, 6C).
Therefore, at least 23 km displacement on the San
Gabriel fault took place prior to the deposition of
the upper part of the conglomerate. Because the
Canton fault rejoins the San Gabriel fault north of the
Castaic Hills oil field (Figure 3), the Canton fault
could not extend southeast into the San Fernando
Valley as proposed by Powell (1993, p. 40). However, examination of Figure 3 shows that the Devil
Canyon fault, which in part forms the southwestern
edge of a basement ridge, could extend into the San
Fernando Valley and take up some right-lateral displacement measured on the San Gabriel fault farther
north. If this fault occupied the position marked
(A) on Figure 3, the Paleogene marine sequence
west of the Whitney Canyon fault would be offset
2530 km from the Paleogene of Piru Creek rather
than be correlated with the Paleogene of the Santa
Susana footwall block (Figure 4). The disadvantage
of this interpretation is that the questionable Sespe
Formation west of the Whitney Canyon fault is relatively fine grained, whereas the Sespe west of the
fault is coarse grained, with evidence of a nearby
eastern source. In summary, the extension of a
strand of the San Gabriel fault into the San Fernando
Valley is neither supported nor refuted by evidence
from the east Ventura basin.
The Hasley Conglomerate forms the base of the
Towsley Formation, and it contains both Mohnian
and Delmontian foraminifera, indicating an age of
about 6.5 Ma. The conglomerate may be in part
reworked from the underlying Devil Canyon Conglomerate, but most of it came from east of the San
Gabriel fault, an area now underlain by coeval finegrained strata of the Castaic Formation. The nearest
primary source for the Hasley Conglomerate is the
western San Gabriel Mountains, requiring that the
apex of the Hasley submarine fan is offset at least 30
km from its inferred source region (Figure 11).
Displacement of the Pliocene Fernando Formation is clearly less than that of Miocene strata. Ehlig
(1975) reported offsets of Pliocene strata in the
western San Gabriel Mountains as follows (Figure
11). (1) Conglomerate at the top of the lower
Pliocene section is offset 1029 km from its provenance. (2) A sliver of marine sandstone in the San
Gabriel fault zone is offset 1132 km from its provenance. (3) Middle to upper Pliocene conglomerate
east of Newhall is offset 619 km from its source.
The uncertainties are based on the distribution of
potential source rocks along the fault and on the
possibility of sediment transport parallel to, rather
than at right angles to, the fault. An additional uncertainty is in biostratigraphic correlation of the largely

Yeats et al.

shallow-marine Fernando rocks of the San Gabriel


Mountains with deep-water strata that have been
age-calibrated by Blake (1991).
The Fernando Formation is the oldest formation
with electric-log correlation across the San Gabriel
fault in the Castaic Hills and Honor Rancho oil fields
(Schlaefer, 1978; Stitt, 1980), indicating that most
strike slip on the fault had been completed prior to
Fernando deposition. West of the modern trace of
the San Gabriel fault, both the Castaic Hills and
Honor Rancho oil fields contain abandoned strands
of the fault that juxtaposed Towsley and Modelo formations against the Castaic Formation prior to deposition of the Fernando Formation (Figure 6D, well
29). The Fernando consists of shelf facies near the
fault, so it is not possible to separate lower Pliocene
(Repetto) from upper Pliocene faunal stages that are
recognized in deep-water Fernando rocks. The zero
isopach of Fernando Formation east of the San
Gabriel fault, marking where the Fernando is overlapped by the Saugus, is offset across the fault at least
2.4 km (Figure 10). The ease of correlation of the Fernando in electric logs across the fault makes it unlikely that the displacement is more than about 5 km.
The present trace of the San Gabriel fault is east of
the strands mentioned above in the Castaic Hills and
Honor Rancho oil fields and east of the Piru fault of
Shepherd (1960) that juxtaposes basement rocks
against nonmarine strata of the Ridge basin 8 km
northwest of Castaic. The Quaternary behavior of
the fault northwest of Honor Rancho oil field differs
greatly from its behavior to the southeast.
Northwest of Honor Rancho oil field, vertical
stratigraphic separation of the Quaternary Saugus
Formation is 200250 m with the east side down.
Surface dips in the Saugus average 2535 and do
not increase appreciably at the fault. Near Honor
Rancho oil field, distinctive clast assemblages in the
Saugus Formation have a right-lateral offset of about
500 m (Weber, 1982) (Figure 11). Kahle (1986)
reported linear ridges, trenches, hillside benches,
and ponded alluvium along the fault trace, and excavations across the fault show evidence of Holocene
displacement (Cotton and Seward, 1984; Cotton,
1985, 1986). However, Kahle (1986) concluded that
Quaternary slip rates are low, less than 1 mm/yr. At
its northwestern end, the San Gabriel fault is overlapped by the Pliocene Hungry Valley Formation,
and movement on the Frazier Mountain reverse fault
also postdates the San Gabriel fault (Crowell, 1982).
The San Gabriel fault is overridden by the postSaugus Santa Felicia reverse fault northwest of Castaic (Stitt, 1986) (Figure 2), although Kahle (1986) and
Weber (1986) disagreed with this interpretation.
In contrast, the San Gabriel fault underwent relatively large Quaternary displacement southeast of
Honor Rancho oil field where it bends eastward. Surface dips in the Saugus are steep and locally over-

1061

turned at the fault. The fault itself decreases in dip


from 80 to 85 north of Honor Rancho oil field to 60
near Saugus oil field. Vertical separation of the base
of the Saugus Formation increases from 90 m at
Honor Rancho to a maximum of 1600 m at Saugus oil
field, decreasing to about 450 m farther southeast.
Accumulation of 1600 m separation since the end of
Saugus deposition 0.60.5 m.y. ago would be at a rate
of 2.53 mm/yr., if all of it occurred in post-Saugus
time. This interpretation is preferred due to the
absence of fault-related breccia in the subsurface on
the downthrown side of the fault. This displacement
rate is faster than the slip rate farther northwest.
As described above, the San Gabriel fault dips
northeast throughout the east Ventura basin, with
normal separation from Honor Rancho oil field
northwest and reverse separation southeast of
Honor Rancho oil field. However, in both segments,
the fault dips approximately 90 to bedding in the
Saugus Formation, which dips predominantly southwest at the fault (Figure 6D, E). When the Saugus
Formation is rotated back to horizontal, the San
Gabriel fault rotates to approximately vertical, a dip
characteristic of other strike-slip faults in California.
The normal vs. reverse separation changes to east
side down from Honor Rancho oil field northwestward and to west side down farther southeast.
Pliocene Tectonics: Change from Extension to
Contraction
The Towsley Formation of latest Mioceneearly
Pliocene age (6.55 Ma) increases in thickness
southwestward toward the Santa Susana fault, yet
there is no Towsley farther southwest in the footwall block of the Santa Susana fault. The Towsley
was probably terminated southwestward by a normal fault that may have served as a zone of weakness
for the younger Santa Susana fault as far east as the
Aliso Canyon oil field. Towsley isopachs closer to
the San Gabriel fault show a pattern of local thicker
and thinner areas that reflect coeval displacement on
the San Gabriel fault. The Towsley is not thicker
beneath the syncline of Saugus near Newhall, indicating that that syncline is younger.
In contrast, Fernando isopachs show relatively little control by the San Gabriel fault, suggesting that
the fault was less active during Fernando deposition.
The Fernando thickens southwestward from about
1400 m northeast of the Newhall-Potrero anticline to
2500 m southwest of this anticline, so that the anticline does not continue upward to the surface
except locally (Winterer and Durham, 1962) (Figure
6D). The limbs of this anticline formed at different
times. The northeast limb was formed after deposition of the Saugus Formation, but the southwest
limb formed much earlier, during deposition of the

1062

East Ventura Basin

Fernando Formation (Figure 6D). Prior to postSaugus tilting, the Newhall-Potrero structure was a
southwest-facing monocline controlled by a reverse
fault that has not continued upward into strata
younger than Modelo Formation. The NewhallPotrero anticline cannot be a fault-propagation fold
(cf. Suppe and Medwedeff, 1990) because it lacks a
backlimb; the northeast limb formed after the monocline did. We have not been able to map the Repetto
and Pico separately across this structure, so we have
not determined the time of growth of this monocline other than during Fernando deposition.
Farther southeast, the Fernando is not sufficiently
preserved in the Oat Mountain syncline to demonstrate thickening across the Pico anticline analogous
to that across the Newhall-Potrero anticline. Nevertheless, we propose that the Pico anticline (Figure
6E) is another southwest-facing monocline formed
during Fernando deposition because (1) the Fernando isopach gradient north of the Newhall-Potrero
anticline curves southward toward an analogous
position with respect to the Pico anticline (Figure
10), (2) the northeast limb of the Pico anticline is
post-Saugus in age, similar to the age relations northeast of the Newhall-Potrero anticline, and (3) in the
footwall block of the Santa Susana fault in the Aliso
Canyon oil field, the Fernando is more than 2000 m
thick and increases in thickness northward toward
the fault in comparison to a thickness of 400600 m
northeast of the Pico anticline (Figure 6F, 10). This
indicates that thickening must occur between the
outcrop north of the Pico anticline and the Santa
Susana fault, which must have cut through the axis
of the Fernando trough rather than bound it on the
south, as it did the Towsley trough. In the footwall
block, the thick Fernando is terminated southward
at the Frew reverse fault (Figure 6E, F), which underwent near-vertical displacement during Fernando
deposition (Yeats, 1979, 1987a).
North of the Santa Clara Valley, we speculate that
the Temescal anticline (Figure 6B) and the Hopper
RanchModelo anticline (Figure 6A) in the Modelo
lobe of the San Cayetano fault are also folds formed as
the Fernando was deposited. The sharp isopach gradient at the Newhall-Potrero anticline continues westnorthwest to the vicinity of the Del Valle fault, where
this gradient projects west to a position immediately
south of the Hopper RanchModelo anticline (Figure
10). To the north, a gradient of 500800 m projects
west toward the Temescal anticline (Figure 10).
There are two extreme possibilities (cf. Yeats,
1983). (1) The Fernando and Saugus formations were
never deposited on the Modelo lobe, implying that
the lobe was positive throughout the time of deposition. This extreme case is unlikely because, in contrast to the Red Mountain fault in the western Ventura basin (Yeats et al., 1987), the Fernando Formation
in the footwall block shows no northward change in

grain size, and no sediments locally derived from the


north, which could be considered as evidence for a
positive Modelo lobe. (2) The Fernando and Saugus
formations were as thick atop the Modelo lobe as
they are in the Santa Clara syncline in the footwall of
the San Cayetano fault. This is also unlikely because it
would require that the Modelo still preserved would
have been overlain by up to 7 km of Towsley, Fernando, and Saugus formations. The porosity of Modelo
sandstone, which comprises the reservoir for several
small oil fields in the Modelo lobe, indicates much
less overburden than the full thickness of post-Modelo strata in the Santa Clara syncline. Furthermore, the
thickness of Fernando in the Santa Clara syncline is
2500 m, thicker than in any section measured northeast of the San Cayetano fault.
The first extreme maximizes the displacement on
the San Cayetano fault prior to the end of Saugus
deposition and minimizes post-Saugus displacement.
The second extreme requires all displacement on
the San Cayetano to be post Saugus. We favor an
intermediate explanation that calls for displacement
on a blind reverse fault generating the Temescal and
Hopper RanchModelo anticlines as monoclines
across which Fernando thicknesses increase from
900 to 2500 m. We project the thickness measured
in the east Ventura basin westward 1012 km to the
Modelo lobe, although it is likely that some westward thinning occurs, and strata younger than
Towsley may never have been deposited west of,
say, Sespe Creek.
In the footwall block of the Santa Susana fault, the
Frew fault underwent displacement during deposition
of the Fernando Formation (Yeats, 1979, 1987a). The
Torrey reverse fault cuts the Frew and is unconformably
overlain by Saugus Formation, documenting reverse
faulting prior to deposition of the Saugus (Yeats, 1987a).
This would document Pliocene reverse faulting because
the base of the Saugus south of the Santa Susana fault is
as old as it is in the Newhall magnetostratigraphic section (Levi et al., 1986; Levi and Yeats, 1993).
Contemporary Tectonics
The present-day structures began to form near the
end of Saugus deposition. The appearance of locally
derived clasts in the Pacoima Formation near Newhall
and the upper Saugus south of the Aliso Canyon oil
field document the beginning of uplift of the Santa
Susana Mountains about 0.70.6 Ma (Levi and Yeats,
1993). We view this uplift as related to upward ramping of the Santa Susana fault across the structural shelf
to the southwest (tectonic inversion; Yeats, 1987a).
The ramp dips 5055 northeast. Atop the ramp, the
leading edge of the Santa Susana fault flattens in dip
locally to near horizontal due to gravitational and
topographic effects. At the Aliso Canyon oil field, the

Yeats et al.

upper part of the fault flattened so much that the fault


moved into an unfavorable orientation with respect to
planes of high shear stress. This resulted in a younger
strand of the Santa Susana fault breaking through the
convex-upward fault surface (older strand of the Santa
Susana fault), which then became folded into an anticline (Yeats, 1987a) (Figure 6F).
An alternate possibility is raised by the January 17,
1994, Northridge earthquake (Mw = 6.7 on the
moment-magnitude scale), which occurred on a previously unknown south-dipping blind thrust beneath
the footwall block of the Santa Susana fault. Uplift of
the Santa Susana Mountains could be produced by
displacement on the south-dipping fault in addition
to displacement on the north-dipping Santa Susana
fault. A corollary to this interpretation, suggested by
T. L. Davis and J. Namson (oral communication,
1994), is that the northeast-dipping limb of the
Newhall-Potrero and Pico anticlines may represent
the forelimb of this blind thrust.
In the west Ventura basin, late Quaternary displacement on the San Cayetano fault was described by Yeats
(1983) and emen (1989), and on the Oak Ridge fault
by Yeats (1988, 1989). The post-Saugus Oak Ridge fault
continues east-northeast along the southern margin of
the Santa Clara Valley, apparently cutting stream terraces of the Santa Clara River, and dies out farther east.
The South strand of the Oak Ridge fault turns eastsoutheast beneath the Santa Susana fault; until the 1994
earthquake, it was believed that this strand was inactive
(Yeats, 1987a). The 1994 earthquake indicates that a
south-dipping fault must continue beneath the Santa
Susana Mountains, and Yeats (1994, unpublished
report) suggested that this fault is the eastward continuation of the Oak Ridge fault.
Near the town of Piru, the San Cayetano fault bifurcates into two strands (Figures 2, 6B). The northern
(Main) strand cuts an alluvial fan west of Piru Creek
and dies out. The southern (Piru) strand is close to,
but south of, the northern margin of the Santa Clara
Valley where it cuts overturned Saugus Formation
rocks (Figure 6B) and dies out farther east.
North of the Santa Susana fault and east of Piru
Creek, the sequence, including the Saugus Formation,
is deformed into a fold belt cut by several south-dipping reverse faults. The southernmost two faults (the
Del Valle and Holser faults) have relatively large separations, whereas the more northerly faults (the
Hasley, Oak Canyon, and Santa Felicia faults) have
only small displacements (Figure 6C). The northernmost fault (Santa Felicia) overrides the San Gabriel
fault, bringing strata of the east Ventura basin over
rocks of the Ridge basin. The Del Valle and Holser
faults occur close to a steep isopach gradient in the
Fernando Formation, which thickens southward
toward the upthrown sides of these faults; thus the
faults are entirely post-Fernando. Subsurface relations
in the Ramona oil field (Figure 6C) show that the

1063

Holser fault cuts obliquely across the axis of the


Ramona anticline, suggesting that the fault postdates
the anticline and is not the failed overturned limb of
it. Previous studies show the Holser fault continuing
eastward to join the San Gabriel fault, but Stitt (1986,
his Figure 21) showed that the Holser fault is not
found in a subsurface cross section southwest of and
parallel to the San Gabriel fault. Folds appear to have
much lower structural relief in the Saugus Formation
than in older beds (Yeats et al., 1985) (Figure 6C),
although there is no major angular unconformity at
the base of the Saugus.
The west Ventura basin contains an enormous
thickness of Pliocene-Pleistocene strata in the footwall block of the San Cayetano fault (Figure 6A). The
east Ventura basin contains a thinner Pliocene-Pleistocene sequence in a fold belt in the hanging wall
block of the Santa Susana fault (Figure 6BD). The
transition between the west and the east Ventura
basin is unclear. The San Cayetano and Santa Susana
faults appear to be part of a single south-verging
reverse-fault system, but they do not join at the surface. Both strands of the San Cayetano fault die out
eastward, and the Santa Susana fault dies out westward. Between these fault tips, the thick PliocenePleistocene sequence of the west Ventura basin is
tilted upward in a narrow syncline that plunges
steeply westward and is characterized by locally
overturned limbs on both flanks (Figure 6B; steep
surface dips in Figure 6C). The base of the Saugus
reaches the surface near the eastern tips of the Oak
Ridge fault and Piru strand of the San Cayetano fault.
Not only these faults but the Holser fault, Del Valle
fault, and the active north strand of the Oak Ridge
fault have their fault tips within 7 km of one another,
all in the transition zone between the west and east
Ventura basins. East of the fault tips, the synclinal
plunge levels off, and the Saugus appears farther east
where the plunge is gently eastward (Figure 2).
How is north-south shortening accommodated in
the zone where the Santa Susana fault and both
strands of the San Cayetano fault die out? The most
likely explanation is that this transition zone covers a
blind thrust of the combined San CayetanoSanta
Susana fault system. The Del Valle, Holser, and other
south-dipping faults farther north would be backthrusts of this blind thrust system. The folds in this
area may be related to blind backthrusts at depth.
Finally, that part of the San Gabriel fault southeast
of Honor Rancho oil field, discussed above, is part of
the north-south shortening, as already documented
for the southern branch of the San Gabriel fault farther east where it merges with the Sierra Madre
frontal fault marking the southern margin of the San
Gabriel Mountains. Post-Saugus dip-slip displacement on the San Gabriel fault southeast of Honor
Rancho oil field is at a higher rate than strike-slip displacement farther northwest.

1064

East Ventura Basin

Figure 12(A) East


Ventura basin at 15
Ma. Patterns represent the same formations as in Figure 1.
Sixty kilometers of
right slip removed
on the San Gabriel
fault, 10 km shortening removed on west
Ventura basin, and
clockwise rotation removed from areas
where documented by magnetic north
arrows (Hornafius, 1984; Terres and
Luyendyk, 1985; Luyendyk and Hornafius,
1987). South strand of future San Gabriel
fault strikes east-west, an unfavorable orientation for strike slip. Note diverse orientations of normal-faulted basins, suggesting
that load stress greatly exceeded all horizontal stresses. (B) East Ventura basin at about
6.5 Ma, the time of deposition of the Hasley
conglomerate, after accumulation of 36 km right slip on the San Gabriel fault. Alamo Mountain (AM) and Frazier
Mountain (FM) provide detritus for the Violin Breccia in the Ridge basin; western San Gabriel Mountains provide
detritus for the Hasley submarine fan at the base of the Towsley Formation in the east Ventura basin. Soledad basin
rotated clockwise to its present orientation, placing the San Gabriel fault in a favorable orientation to accumulate
strike slip.

Yeats et al.

DISCUSSION
Relation of Miocene Rifted Zone to Younger
Basin Depocenters
Crowell (1973, 1976) and Yerkes and Campbell
(1976) pointed out that the Ventura basin may have
had a Miocene normal-faulted precursor. Yeats
(1987b, 1989) showed that Oak Ridge was a positive
feature controlled by normal faults, inferring that the
Oak Ridge fault bounding Oak Ridge on the north
could also have had a normal-fault precursor (Figure
12A). But Miocene strata are too deep to document
this structure, and industry multichannel seismic lines
have not revealed it either. The east Ventura basin has
strong enough structural relief to shed light on
Miocene geology, revealing the trough extending longitudinally northwest-southeast through the basin.
This rift may contain an arm that extends westward
beneath the west Ventura basin around the north side
of the Oak Ridge fault, as shown in Figure 12A. This
relationship does not apply to the Los Angeles basin
on the south side of the Transverse Ranges in which
the Pliocene-Pleistocene depocenter is unrelated to
areas of thick middle and upper Miocene strata (Yeats
and Beall, 1991). However, the northwestern edge of
the San Fernando Valley is controlled by the
Chatsworth normal fault (Figure 12A), which cuts
middle Miocene rocks but predates deposition of
Mohnian strata (Shields, 1977; Yeats, 1987b) because
it controlled the position of the northwest edge of the
Mohnian Tarzana submarine fan (Sullwold, 1960).
The diverse orientation of Miocene normal-faulted
basins suggests that the load stress was much greater
than horizontal stresses, and the intermediate and
minimum principal compressive stresses were close
together in value. But what was the orientation of
these basins with respect to north? Terres and
Luyendyk (1985) showed that the Oligocene Vasquez
Formation was rotated 53 clockwise prior to deposition of at least part of the Mint Canyon Formation,
and Hornafius (1984), Luyendyk and Hornafius
(1987), and Luyendyk (1991) argued that clockwise
rotation of this age and younger characterized most
of the western and central Transverse Ranges.
Figure 12A is a speculative restoration of the Ventura basin and adjacent regions at 15 Ma, prior to the
beginning of strike slip on the San Gabriel fault. Clockwise rotations of the Soledad basin, the east Ventura
basin north of the Santa Clara River, and the Santa
Monica Mountains are restored. The Simi Hills, the
San Gabriel Mountains, and basement terranes north
of the San Francisquito fault are assumed without evidence to have been rotated as much as the Soledad
basin and Santa Monica Mountains. However, the
Cuyama basin, north of the Big Pine fault, is not rotated, following Luyendyks model. (However, Ellis et
al., 1993, found up to 20 clockwise rotation of

1065

Pliocene strata in the Cuyama basin.) Ten kilometers


of post-Miocene contraction in the west Ventura
basin is removed, and this contraction is removed in
the east Ventura basin as well. This restoration gives
the east Ventura basin rift and the granitic ridge north
of it an east-west orientation in the late Miocene.
Role of the San Gabriel Fault
The east-west orientation of the east Ventura basin
rift and the granitic ridge to the north prior to clockwise tectonic rotation was unfavorable for the accumulation of strike-slip displacement. The San Gabriel fault
is parallel to the southwest edge of the granitic ridge,
suggesting that the fault may have had a middle
Miocene normal-fault ancestor. Clockwise rotation in
the middle Miocene changed the orientation of the
granitic ridge and the rift south of it from east-west (Figure 12A) to northwest-southeast (Figure 12B). Because
of the new favorable orientation, the San Gabriel fault
began to accumulate right slip at 1012 Ma. This was at
least 23 m.y. after the initiation of the middle Miocene
rift in the east Ventura basin. However, deposition in
the rift continued as major right slip on the San Gabriel
fault accumulated (Figure 13), suggesting an extensional component of strike slip on the fault. An extensional
component would have been present adjacent to the
Ridge basin for the same reason. This indicates that the
San Gabriel fault may have propagated along one of the
rift boundary faults at the time the boundary fault
became favorably oriented for this to occur.
Right slip began rather abruptly 1012 m.y. ago
and continued through the Pliocene at rates of 69
mm/yr (Figure 11). The arrows for DC and CA show
that the displacement could have occurred at any
time after the time shown; the arrow for DCF and HA
show that the displacement could have been larger
than that shown up to the maximum of 60 km. The
69 mm/yr rate (4.55.5 mm/yr if the total displacement is 45 rather than 60 km) takes into account the
initiation of strike slip prior to the end of Mint
Canyon deposition, the relatively large displacements
of Fernando Formation in the western San Gabriel
Mountains, and the small displacement of Fernando
Formation in the east Ventura basin prior to deposition of the Saugus Formation, discussed below.
Uncertainty in timing and amount of displacements shown on Figure 11 leads to uncertainty in
the age of the slowdown of strike slip on the San
Gabriel fault. A slowdown by the Pliocene is suggested by the overlap of the San Gabriel fault at its northwestern end by the Pliocene Hungry Valley Formation (Crowell, 1982). However, the Fernando
displacements in the western San Gabriel Mountains
(Ehlig, 1975) (FM on Figure 11) suggest that the
slowdown took place in the Pleistocene. The ease of
correlation of the Fernando Formation across the

1066

East Ventura Basin

Figure 13Summary diagram of tectonic evolution of


east Ventura basin.
Width of the structural activity bars
are proportional to
the displacement
rate. Inset boxes
are diagrammatic
maps and cross sections showing five
stages of development. On the maps,
solid lines are
faults with high
slip rates, dashed
lines show faults
with low slip rates,
dotted lines show
inactive faults.

San Gabriel fault at Honor Rancho oil field (FV on


Figure 11) favors the earlier slowdown. If the Fernando Formation in the western San Gabriel Mountains is older than assumed by Ehlig (1975), then the
1.3 mm/y rate could be extended back to 4 Ma, the
age of the opening of the Gulf of California. Uncertainty in the Fernando age is likely; Nelligan (1978)
was unable to separate the Fernando and Towsley in
the easternmost Ventura basin. However, this would
increase the pre-Pliocene slip rate on the San Gabriel
fault to at least 9 mm/y (Figure 11) or 4.56 mm/yr
based on a total displacement of 45 km.
Low strike-slip rates have characterized this part
of the San Gabriel fault in the east Ventura basin for
at least the last 0.7 m.y. (SA on Figure 11), an expression of the near abandonment of the San Gabriel
fault as a major plate-boundary displacement zone.
The reason for this is probably the bend in the San
Gabriel fault to east-west in the San Gabriel Mountains; the fault is no longer in a favorable orientation
to accumulate strike slip. The N40W strike of the
fault adjacent to the Ridge basin is favorable to accumulate right slip, but fault displacement is blocked
by the bend to the east.

However, the bend to the east orients the San


Gabriel fault parallel to the Santa Susana fault, a zone
of major late Quaternary displacement, and perpendicular to the direction of tectonic shortening according to Zoback et al. (1987). Post-Saugus reverse slip
(or, more likely, reverse-oblique slip) appears to be as
high as 3 mm/yr, a much higher rate than coeval
strike slip farther northwest. This assumes that the
1500-m separation of the base of the Saugus Formation at Saugus oil field took place since the end of
Saugus deposition 0.5 m.y. ago. This eastward
increase in late Quaternary displacement implies a
rotational component of slip on the San Gabriel fault.
The parallelism of the reverse faults and associated folds to the Mojave segment of the San Andreas
fault is analogous to the parallelism of the San
Andreas fault to folds and thrusts in the central Coast
Ranges farther north (Zoback et al., 1987).
Segmentation of the Quaternary Deformed Belt
The south-verging reverse-fault system of the Transverse Ranges is interrupted by lateral ramps marked

Yeats et al.

1067

Figure 14Quaternary segmentation of the east Ventura basin. Northeast-trending structures (segment boundaries) mark the termination of faults and folds and the change in late Quaternary history of the San Gabriel fault.

by very steep fault dips and local left-lateral displacements (Figure 14). The Santa Susana fault is marked
by two lateral ramps, one at Gillibrand Canyon and
one at the northwest margin of the Sylmar basin in the
San Fernando Valley (Yeats, 1987a) (located on Figure
14). The latter structure contains the main shock of
the 1971 San Fernando earthquake. Aftershocks of the
1971 earthquake on this zone yielded northeast-trending left-lateral fault-plane solutions (Whitcomb et al.,
1973), consistent with the lateral-ramp geology. However, aftershocks of the 1994 Northridge earthquake
do not appear to be affected by this boundary (E.
Hauksson, oral communication, 1994). Whitcomb et
al. (1973) further suggested that the northeast-trending aftershock zone was part of a longer lineament
that could be correlated from the Pacific coast to the
Mojave Desert. Yeats (1987a) suggested that the Gillibrand Canyon lateral ramp (located on Figure 14)
could be traced northeast through the east Ventura
basin to the San Gabriel fault.
The lateral ramp marking the northwest margin of

the San Fernando Valley is here named the Chatsworth


ramp after the Chatsworth normal fault that forms the
northwest margin of the valley farther southwest (Figure 12A). As discussed above, the Chatsworth fault
formed in the middle Miocene and influenced the
position of the west edge of the upper Miocene (Mohnian) Tarzana submarine fan. Farther north, the Whitney Canyon fault had two stages of movement (Figure
6F). Pre-Towsley separation was normal and down to
the west, bringing a thick Paleogene sequence in fault
contact with basement rocks of the San Gabriel Mountains to the east. Post-Towsley separation, relatively
minor compared to the pre-Towsley displacement,
was reverse and down to the east.
The Gillibrand Canyon structure was described in
detail by Yeats (1987a). It is mainly expressed by the
lateral ramp in the Santa Susana fault, but the preSaugus Torrey fault (Figure 6CE) also bends eastward
across it (Yeats, 1987a). In the east Ventura basin, the
Oat Mountain syncline and Pico anticline terminate
westward against it. The San Gabriel fault is south-

1068

East Ventura Basin

ward-convex with late Quaternary reverse displacement southeast of the structure and straight with normal separation and slow late Quaternary strike-slip
displacement northwest of it. The Del Valle and
Holser faults and Newhall-Potrero anticline terminate
eastward against it. A zone of reduced Towsley thickness follows the Gillibrand Canyon structure to the
San Gabriel fault. A northeast-trending zone of high
seismicity follows the zone across the east Ventura
basin and San Gabriel fault (Pechmann, 1987).
The Piru segment boundary occurs at the concentration of fault tips at the transition between the west and
east Ventura basin (Figure 14). As noted above, the Oak
Ridge fault and both strands of the San Cayetano fault
terminate eastward and the Holser and Del Valle faults
terminate westward against it, as do other backthrusts
farther north. The Santa Felicia fault follows the boundary for a considerable distance southwest before dying
out. Saugus and Fernando thicknesses in the Santa
Clara Valley are much greater west of the boundary.
The Saugus is at maximum burial west of the boundary,
but is largely removed by erosion east of it. Aftershocks
of the 1994 Northridge earthquake terminated westward close to this segment boundary.
These three boundaries separate four segments of
the late Quaternary deformed zone (Figure 14). On
the west is the Hopper Canyon segment, named for
a prominent gorge in the Modelo lobe of the San
Cayetano fault (Figure 6A). Within this segment, the
Pliocene-Pleistocene sequence of the Santa Clara Valley is largely at maximum burial, cut by the San
Cayetano fault on the north and the Oak Ridge fault
on the south. Late Quaternary displacements north
of the San Cayetano fault and south of the Oak Ridge
fault are relatively small. The western boundary of
this segment, not described in this paper, may be
the western edge of the Modelo lobe at Sespe Creek.
The Newhall-Potrero segment is bounded by the
Piru and Gillibrand Canyon structures. This segment
is characterized by south-dipping backthrusts, with
the largest being the Del Valle and Holser faults, by a
fold belt, and by thrusting of the east Ventura basin
sequence southward on the Santa Susana fault (Figure 6C, D). The Placerita segment is bounded by the
Gillibrand Canyon and Chatsworth structures. This
segment is also characterized by southward thrusting on the Santa Susana fault, but the fold belt and
backthrusts farther north are absent (Figure 6E, F).
The San Gabriel fault is characterized by late Quaternary reverse slip in this segment.
The Sylmar segment is east of the Chatsworth
structure. As described by Shields (1977), it is characterized by a great thickness of Pliocene-Pleistocene
strata largely at maximum burial overridden on the
north by the Santa Susana fault zone. This fault has the
structural role of the San Cayetano fault in the west
Ventura basin, except that the hanging wall is made
up of crystalline rocks of the San Gabriel Mountains.

This fault did not move in either the 1971 or the 1994
earthquake; the 1971 San Fernando fault forms the
southern edge of the Sylmar Pliocene-Pleistocene
trough, not the northern edge. The 1971 San Fernando fault forms part of the Granada HillsMission Hills
fault zone (located on Figure 1), which has tectonic
geomorphic expression and is traced westward to the
Chatsworth structure where it disappears. Farther
south, the Northridge Hills fault zone also appears to
be active, but it, too, can be traced only as far west as
the Chatsworth structure. The south-dipping fault
marking the southern edge of the Sylmar basin may be
a surface expression of the otherwise buried southdipping fault that ruptured in the 1994 Northridge
earthquake (Yeats, 1994, unpublished report).
Oil and Gas Production in the East Ventura
Basin
The distribution of hydrocarbons in the east Ventura basin (shown on Figure 7) is sufficiently different from that in the west Ventura basin that it should
be considered as a separate petroleum subbasin. Oil
production is concentrated in turbidite sandstones
of the Towsley and Modelo formations, whereas in
the west Ventura basin between the Oak Ridge and
San CayetanoRed Mountain fault zones, it is concentrated in Fernando turbidite sandstones. Fernando production is found in shelf-facies strata in the
northeast and east margins of the basin, including
two oil fields (Wayside Canyon, Tapia Canyon) east
of the San Gabriel fault (Figure 7). Middle Miocene
strata are productive in the northern margin of the
basin, which has undergone very deep erosion.
According to Magoon and Taylor (1975), oil fields in
the east Ventura basin have an ultimate recovery of
377.1 million BOE (barrels of oil plus gas equivalent).
These are listed in Table 2. Of this total, 250 million BOE
occur in just four oil fields: Newhall-Potrero, Castaic
Junction, Del Valle, and Placerita. All occur in areas
where Pliocene or younger strata are at the surface. The
distribution of major oil reservoirs is consistent with
migration into structures that formed during deposition
of the Fernando and Saugus formations, structures that
were later deformed during the late Quaternary. The
most likely source rocks are fine-grained organic clastic
rocks of Monterey Shale lithology of middle and late
Miocene age (Modelo Formation). As oil migrated north
and west from the thick, deeply buried Miocene trough,
it must have crossed stratigraphic boundaries so that it
occurs in Pliocene shelf-edge strata in the Placerita,
Elsmere, Tapia Canyon, and Wayside Canyon oil fields.
The occurrence of oil in the Tapia Canyon and Wayside
Canyon oil fields is evidence that most of the strike slip
on the San Gabriel fault had already occurred prior to oil
migration, and that the San Gabriel fault was not a barrier to oil migration.

Yeats et al.

CONCLUSIONS
The east Ventura basin originated as a rift basin in
the middle Miocene extending from the outcrop
north of the Modelo lobe of the San Cayetano fault
southeast to the subsurface north of the Santa Susana
Mountains. This rift contains a great thickness of middle and upper Miocene strata dominated by sandstone. The rift is bounded on the northeast by a basement ridge parallel and close to the San Gabriel fault
and on the southwest by the Oak RidgeSimi Hills
structural shelf. The area continued to be strongly
negative during deposition of the Mohnian Modelo
Formation (Figure 13). South of the Santa Clara River,
the Towsley Formation may have been rift related,
and the depocenter appears to have shifted southwest toward a normal-fault precursor to the Santa
Susana fault, so that the thickest Towsley is found in
its outcrop belt in the Santa Susana Mountains.
The San Gabriel fault may have started as a normal
fault related to rifting and striking east-west prior to
clockwise rotation of the Soledad basin and adjacent
areas (Figure 12A). Rotation changed the strike of the
fault to northwest-southeast, a favorable orientation to
begin to accumulate 60 km of right-lateral displacement (Figures 12B, 13). The rift continued as a negative
feature as strike slip accumulated on the San Gabriel
fault in the late Miocene, suggesting that the strike slip
had an extensional component. Strike slip juxtaposed
the east Ventura basin against the Ridge basin and
occurred at a rate of 4.59 mm/y, the uncertainty being
the age of initiation and termination of major strike slip
and the amount of total displacement (Figure 11).
The Fernando Formation rests on the contrasting
Miocene sequences of the east Ventura and Ridge
basins, and it is itself displaced right laterally much
less than older strata. Southwest of the San Gabriel
fault, the Newhall-Potrero anticline developed as a
monocline above a blind reverse fault during Fernando deposition, and the Pico anticline to the southeast and the Temescal and Hopper RanchModelo
anticlines to the northwest may have formed in the
same way at the same time. In addition, there was
probably displacement on the San Cayetano fault
during Fernando deposition.
Late Quaternary structures include the south-verging San Cayetano and Santa Susana reverse faults, the
north-verging backthrusts of the east Ventura basin
fold belt, a blind thrust connecting the San Cayetano
and Santa Susana faults at depth, and the San Gabriel
fault. North of the Santa Clara River, where the San
Gabriel fault is straight with a N40W strike, right
slip on the fault slowed to 1.3 mm/yr or less by 0.7
Ma and possibly as early as 4 Ma. The more easterly
striking segment of the fault south of the Santa Clara
River was reactivated as a reverse fault with a slip
rate of several millimeters/year. The late Quaternary
deformed belt is divided by northeast-trending trans-

1069

Table 2. Oil Fields of the East Ventura Basin

Oil Field

Map
Symbol Reservoir*

Canton Creek
Castaic Hills
Castaic Junction
Del Valle
De Witt Canyon
Elsmere
Eureka Canyon
Hasley Canyon
Holser Canyon
Honor Rancho
Hopper Canyon
Lyon Canyon
Newhall-Potrero
Oak Canyon
Pico Canyon
Piru
Placerita
Ramona
Rice Canyon
Saugus
Tapia Canyon
Tapo North
Tapo Ridge
Temescal
Townsite
Towsley Canyon
Tunnel
Wayside Canyon
Whitney Canyon
Wiley Canyon

CC
CH
CJ
DV
DW
EL
EU
HS
HL
HR
HP
LC
NP
OC
PC
PI
PL
RA
RC
SA
TC
TN
TR
TE
TO
TW
TU
WA
WH
WI

MM
MU
MU, TO
MU, TO, FE
MU
FE
MU, TO
TO
MU, TO
MU, TO
MM, MU
MU
MU, TO
MM, MU
MU
MU
FE
TO
MU
MU
FE
MU
MU
MM, MU
MU
MU
MU, TO
FE
TO
MU

Recoverable
Reserves
Trap** (106 BOE)
OT
<0.1
ST
12.2
AN, FA
61.6
AN
42.6
AN
<0.1
AN, FA
3.2
ST, OT
0.5
FA
0.1
AN, FA, OT 0.7
ST
37.5
AN
3.3
ST, FA
0.3
AN
101.5
AN
16.8
AN
4.0
AN, ST
0.5
FA
46.0
FA, ST
28.8
AN
0.2
FA
0.5
ST
1.1
FA, OT
1.5
AN
<0.1
AN
7.2
FA
0.2
AN
<0.1
FA
3.0
ST
3.0
FA
0.4
AN, ST
0.5

Total recoverable reserves

377.1

*MM = middle Miocene, MU = Mohnian, TO = Towsley, FE = Fernando.


**ST = stratigraphic, FA = fault, AN = anticline, OT = other.
BOE = barrels of oil and gas equivalent.

verse structures into four segments: (1) the Hopper


Canyon segment in which most displacement is on
the San Cayetano and Oak Ridge faults, and the intervening Pliocene-Pleistocene sequence is largely at
maximum burial, (2) the Newhall-Potrero segment
in which deformation is by a southward-vergent
blind thrust along with the west third of the Santa
Susana fault, northeast of which are backthrusts and
the east Ventura basin fold belt, (3) the Placerita segment dominated by the southward-vergent Santa
Susana fault and a reactivated segment of the San
Gabriel fault, and (4) the Sylmar segment in which a
very thick Pliocene-Pleistocene sequence is overridden by basement rocks of the San Gabriel Mountains
on the Santa Susana fault, and the smaller displacement, southward-verging Granada HillsMission
Hills and Northridge Hills active reverse fault zones
occur farther south.

1070

East Ventura Basin

APPENDIX: WELLS IN FIGURE 6


Figure 6A
1
2
3
4
5
6
7
8
9
10
11

Texaco 165 Shiells


Signal 1 Calumet
Kencal 1 Calumet
Unocal 1 William Shiells
Arco 1 Laura Lawton
Texaco 1 Lawton
Sovereign 1 Leavins Goodenough
Continental 1 Elkins
McCulloch 1A Hopper Canyon
McCulloch 1 Hopper Canyon
Unocal 1 Moran

Figure 6B
1
2
3
4
5
6
7
8
9
10

Unocal 90 Torrey
Unocal 107 Torrey
Chevron 1 Stevens Eureka
Phillips 1 Sloan Ranch
Texaco B-1 Sloan
Shell 1 Sloan
McCulloch 1 Burger
Mobil 1 Camulos Ranch
McCulloch 1-16 IDS
Getty 14 Temescal

Figure 6C
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32

Chevron 3 Tapo Core Hole


Havenstrite 1 Tapo
Unocal 1 South Tapo
Unocal 5 South Tapo
Unocal 7 South Tapo
Unocal 109 Broad Oaks
Unocal 28 Simi
Unocal 103 Broad Oaks
Unocal 11 Simi
McCulloch 1 NL&F
Conoco 2 NL&F
Conoco 1 NL&F
Unocal C-1 NL&F
Rheem 1 Larinan
Superior 13 Black
Superior 15 Black
Superior 9 Black
Superior 12 Black
Superior 5 Black
Superior 3 Black
Superior 4 Black
Arco 7-15 Black Ranch
Arco Hamilton Bros. 12-84 Black Ranch
Arco 1 Lechler
Gulf 1 WSW
Gulf 4 Lechler
Amax 2 Ayala O. H.
R. Edgar 1 Jesus
Superior 51-28 Romero
Aminoil 77-21 Romero Loma Verde
Conoco 1 Vier Kenny
Unocal 1 Alexander

Figure 6D
1
2
3
4
5
6
7
8
9
10
11

Unocal 29 Simi
Unocal 23 Simi
Unocal 1 Chivo Canyon
Unocal 1 Blue Sage
Sun A-2 NL&F
Sun 103 RSF
Unocal 1 RSF
Sun 9 RSF
Sun 90 RSF
Sun A-1 RSF
Sun 8 RSF

12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35

Sun 44 RSF
Sun 97 RSF
Sun 141 RSF
Sun 148 RSF
Exxon 78 NL&F
Exxon 18 NL&F
Exxon 8 NL&F
Exxon 27 NL&F
Exxon 33 NL&F
Exxon 63 NL&F
Exxon 76 NL&F
Exxon 69 NL&F
Texaco F-1 Newhall
Texaco D-3 Newhall
Texaco 25 Wayside
Texaco 20 Honor Rancho A (NCT-1)
Texaco 6 Honor Rancho A (NCT-2)
Texaco 8 Honor Rancho A (NCT-2)
Texaco 42 Honor Rancho A (NCT-2)
Texaco 12 Honor Rancho A (NCT-2)
Texaco 28 Wayside Canyon Unit
Texaco 25 Honor Rancho A (NCT-2)
Texaco 39 Wayside Canyon Unit
Texaco 29 Wayside Canyon Unit

Figure 6E
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20

Placerita 2
Placerita 1
Unocal 1-1 Del Aliso
Unocal 1-2 Del Aliso
Sun 1 Limbocker
International (Mayock) 1 Foster
Montara 1 Patrick Towsley Canyon
Celeron 1 Chevron Towsley
Sun 2 Lassalle
Sun 1 Lassalle
Von Glahn 1 Lassalle
Argosy 1 Scott Lassalle
Rheem 1 Miller Ranch
Rheem 1 Happy Valley Unit
Eagle 1 Eagle
Mobil J-1 Circle
Mobil J-2 Circle
Unocal 1 Bermite
Termo 1 TB
Terminal Ind. 1 Chiggion

Figure 6F
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29

P. S. 6 Limekiln
S. C. G. 5 Sesnon Fee
Getty 40 Porter
Getty 42 Porter Sesnon
Getty 38 Porter
Getty 56 Porter
Getty 60 Porter
Getty 71 Porter
Getty 25 Porter
Getty 20 Porter Sesnon
Getty 66 Porter
Hamilton Sherman 1 Orcutt
Chevron 2-1 Orcutt Trustee
Mobil 1 Mendota
Texaco 1-A Evans
Terminal 1 Hamilton
H. S. Russell 1 Needham
Morton & Dolley B-1 Needham
Morton & Dolley 5 Needham
Terminal 1 Phillips
Albert 102 Albert
Conoco 1 Phillips
Albert 23 Albert
Albert 19 Albert
Range & L. A. Ault 1 Dorothy
W. Y. Lee 1 Heil
San Gabriel 1 San Gabriel
Lee 1 Lee
Conoco 1 Wallace USL

Yeats et al.

REFERENCES CITED
Barron, J. A., 1976, Marine diatom and silicoflagellate biostratigraphy of the type Delmontian Stage and the type Bolivina obliqua zone, California: Journal of Research of the U. S. Geological Survey, v. 4, no. 3, p. 339351.
Berggren, W. A., 1969, Cenozoic chronostratigraphy, planktonic
foraminiferal zonation and the radiometric time scale: Nature,
v. 224, no. 5214, p. 10721075.
Blake, G. H., 1991, Review of the Neogene biostratigraphy and
stratigraphy of the Los Angeles basin and implications for
basin evolution, in K. T. Biddle, ed., Active margin basins:
AAPG Memoir 52, p. 135184.
Bohannon, R. G., 1975, Mid-Tertiary conglomerates and their bearing on Transverse Range tectonics, southern California: California Division of Mines and Geology Special Report 118, p. 7582.
Bohannon, R. G., 1976, Mid-Tertiary nonmarine rocks along the
San Andreas fault in southern California: Ph.D. dissertation,
University of California, Santa Barbara, California, 327 p.
Bramlette, M. N., 1946, Monterey Formation of California and origin of its siliceous rocks: U. S. Geological Survey Professional
Paper 212, 57 p.
California Division of Oil and Gas, 1974, California oil and gas
fields, v. II, south, central coastal and offshore California: California Division of Oil and Gas Report TR 12.
Carman, M. F., 1964, Geology of the Lockwood Valley area, Kern
and Ventura counties, California: California Division of Mines
and Geology Special Report 81, 62 p.
emen, I., 1977, Geology of the SespePiru Creek area, Ventura
County, California: M.S. thesis, Ohio University, Athens, Ohio,
69 p.
emen, I, 1989, Near-surface expression of the eastern part of the
San Cayetano fault: a potentially active thrust fault in the California Transverse Ranges: Journal of Geophysical Research,
v. 94, p. 96659677.
Cordova, S., 1966, Castaic Junction oil field: California Division of
Oil and Gas, Summary of Operations, California Oil Fields,
v. 52, no. 2, part 2, p. 5565.
Cotton, W. R., 1985, Holocene behavior of the San Gabriel fault,
Saugus/Castaic area, Los Angeles County, California: Final
Technical Report to U. S. Geological Survey, Contract No.
1408000121950, 26 p.
Cotton, W. R., 1986, Holocene paleoseismology of the San
Gabriel fault Saugus/Castaic area, Los Angeles County, California, in P. L. Ehlig, compiler, Neotectonics and faulting in
southern California: Guidebook and volume prepared for the
82nd annual meeting of the Cordilleran Section of the Geological Society of America, p. 3341.
Cotton, W. R., and A. E. Seward, 1984, Engineering geologic
investigation of the San Gabriel fault: Report to the Newhall
Land and Farming Company, Valencia, California, April 11,
1984, v. I, 34 p.; v. II, 13 plates.
Crowell, J. C., 1952, Probable large lateral displacement on the
San Gabriel fault, southern California: AAPG Bulletin, v. 36,
p. 20262035.
Crowell, J. C., 1954, Strike-slip displacement on the San Gabriel
fault, southern California: California Division of Mines Bulletin
170, p. 4852.
Crowell, J. C., 1962, Displacement along the San Andreas fault, California: Geological Society of America Special Paper 71, 61 p.
Crowell, J. C., 1973, Problems concerning the San Andreas fault
system in southern California, in R. L. Kovach, and A. Nur,
eds., Proceedings of the conference on tectonic problems of
the San Andreas fault system: Stanford University Publications
Geological Sciences, v. 13, p. 125135.
Crowell, J. C., 1975, The San Gabriel fault and Ridge basin, southern California: California Division of Mines and Geology Special Report 118, p. 208219.
Crowell, J. C., 1976, Implication of crustal stretching and shortening of coastal Ventura basin, California, in D. G. Howell, ed.,

1071

Aspects of the geologic history of the California continental


borderland: Pacific Section AAPG Miscellaneous Publication
24, p. 365382.
Crowell, J. C., 1982, The tectonics of Ridge basin, southern California, in J. C. Crowell, and M. H. Link, eds., Geologic history
of Ridge basin, southern California: Pacific Section SEPM,
p. 2542.
Crowell, J. C., and M. H. Link, 1982, eds., Geologic history of
Ridge basin, Southern California: Pacific Section SEPM, 304 p.
Dibblee, T. W., Jr., 1982, Geology of the Santa YnezTopatopa
Mountains, southern California, in D. L. Fife, and J. A. Minch,
eds., Geology and mineral wealth of the California Transverse
Ranges: South Coast Geological Society, p. 4156.
Dibblee, T. W., Jr., 1989, Mid-Tertiary conglomerates and sandstones on the margins of the Ventura and Los Angeles basins
and their tectonic significance, in I. P. Colburn, P. L. Abbott,
and J. Minch, eds., Conglomerates in basin analysis: a symposium dedicated to A. O. Woodford: Pacific Section SEPM,
v. 62, p. 207226.
Dibblee, T. W., Jr., 1990, Geologic map of the Fillmore quadrangle Ventura County, California: Santa Barbara, California, Dibblee Geological Foundation, scale 1:24,000.
Dibblee, T. W., Jr., 1991, Geologic map of the Piru quadrangle
Ventura County, California: Santa Barbara, California, Dibblee
Geological Foundation, scale 1:24,000.
Ehlert, K. W., 1982, Basin analysis of the Miocene Mint Canyon
Formation, southern California, in R. V. Ingersoll, and M. O.
Woodburne, eds., Cenozoic nonmarine deposits of California
and Arizona: Pacific Section SEPM, p. 5164.
Ehlig, P. L., 1975, Geologic framework of the San Gabriel Mountains:
California Division of Mines and Geology Bulletin 196, p. 718.
Ehlig, P. L., 1981, Origin and tectonic history of the basement terrane of the San Gabriel Mountains, central Transverse Ranges,
in W. G. Ernst, ed., The geotectonic development of California: Englewood Cliffs, New Jersey, Prentice-Hall, p. 253283.
Ehlig, P. L., and J. C. Crowell, 1982, Mendenhall Gneiss and
anorthosite-related rocks bordering Ridge basin, California: in
J. C. Crowell, and M. H. Link, eds., Geologic history of Ridge
basin, southern California: Pacific Section SEPM, p. 199202.
Ehlig, P. L., and S. E. Joseph, 1977, Polka dot granite and correlation of La Panza quartz monzonite with Cretaceous batholithic
rocks north of Salton trough, in D. G. Howell, J. G. Vedder,
and K. McDougall, eds., Cretaceous geology of the California
Coast Ranges west of the San Andreas fault: Pacific Coast paleogeography field guide 2: Pacific Section SEPM, p. 9196.
Ehlig, P. L., K. W. Ehlert, and B. M. Crowe, 1975, Offset of the
upper Miocene Caliente and Mint Canyon formations along
the San Gabriel and San Andreas faults: California Division of
Mines and Geology Special Report 118, p. 8392.
Eldridge, G. H., and R. Arnold, 1907, The Santa Clara Valley,
Puente Hills, and Los Angeles oil districts, southern California:
U. S. Geological Survey Bulletin 309, 266 p.
Ellis, B. J., S. Levi, and R. S. Yeats, 1993, Magnetic stratigraphy of
the Morales Formation: late Neogene clockwise rotation and
compression in the Cuyama basin, California Coast Ranges:
Tectonics, v. 12, p. 11701179.
Frizzell, V. A., Jr., and R. E. Powell, 1982, Crystalline rocks near
Frazier and Alamo mountains, western Transverse Ranges, California: a comparative study (abs.): Geological Society of America Abstracts with Programs, v. 14, p. 164.
Hershey, O. H., 1902, Some Tertiary formations of southern California: American Geologist, v. 29, p. 349372.
Hornafius, J. S., 1984, Paleomagnetism of the Monterey Formation
in the western Transverse Ranges, California: Ph.D. thesis, University California, Santa Barbara, California, 442 p.
Hornafius, J. S., 1985, Neogene tectonic rotation of the Santa
Ynez Range, western Transverse Ranges, California, suggested
by paleomagnetic investigation of the Monterey Formation:
Journal of Geophysical Research, v. 90, p. 1250312522.
Hudson, F. S., and E. K. Craig, 1929, Geologic age of the Modelo
Formation, California: AAPG Bulletin, v. 5, p. 509518.

1072

East Ventura Basin

Huftile, G. J., 1988, Structural geology of the Upper Ojai Valley


and Chaffee Canyon areas, Ventura County, California: M.S.
thesis, Oregon State University, Corvallis, Oregon, 103 p.
Huftile G. J., 1991, Thin-skinned tectonics of the Upper Ojai Valley and Sulphur Mountain area, Ventura basin, California:
AAPG Bulletin, v. 75, p. 13531373.
Huftile, G. J., and R. S. Yeats, in press, Cenozoic structure of the
Piru 712-minute quadrangle, California: U. S. Geological Survey
OpenFile Report, 33 sheets, scale 1:24,000.
Jahns, R. H., 1940, Stratigraphy of easternmost Ventura basin, California: Carnegie Institution of Washington Publication 514,
p. 145194.
Jennings, C. W., compiler, 1975, Geologic map of California: California Division of Mines and Geology, scale 1:750,000.
Jennings, C. W., and R. G. Strand, 1969, Geologic map of California 1:250,000: California Division of Mines and Geology, Los
Angeles sheet.
Joseph, S. E., T. E. Davis, and P. L. Ehlig, 1982, Strontium isotopic correlation of the La Panza Range granitic rocks with similar rocks in
the central and eastern Transverse Ranges, in D. L. Fife, and J. A.
Minch, eds., Geology and mineral wealth of the California Transverse Ranges: South Coast Geological Society, p. 310320.
Kahle, J. E., 1986, The San Gabriel fault near Castaic and Saugus,
Los Angeles County: California Division of Mines and Geology
Fault Evaluation Report FER 178, 8 p.
Kew, W. S. W., 1924, Geology and oil resources of a part of Los
Angeles and Ventura counties, California: U. S. Geological Survey Bulletin 753, 202 p.
Kleinpell, R. M., 1938, Miocene stratigraphy of California: AAPG,
450 p.
Kriz, S. J., 1947, Stratigraphy and structure of the Whitaker
PeakReasoner Canyon area, Ventura and Los Angeles Counties, California: Ph.D. dissertation, Princeton University,
Princeton, New Jersey, 68 p.
Lant, K. J., 1977, Structure of the Aliso Canyon area, eastern Ventura basin, California: M.S. thesis, Ohio University, Athens,
Ohio, 79 p.
Levi, S., and R. S. Yeats, 1993, Paleomagnetic constraints on the
initiation of uplift on the Santa Susana fault, western Transverse Ranges, California: Tectonics v. 12, p. 688702.
Levi, S., D. L. Schultz, R. S. Yeats, L. T. Stitt, and A. M. SarnaWojcicki, 1986, Magnetostratigraphy and paleomagnetism of the
Saugus Formation near Castaic, Los Angeles County, California, in P. L. Ehlig, compiler, Guidebook and volume, neotectonics and faulting in southern California: Geological Society
of America, p. 103108.
Luyendyk, B. P., 1991, A model for Neogene crustal rotations,
transtension, and transpression in southern California: Geological Society of America Bulletin, v. 103, p. 15281536.
Luyendyk, B. P., and Hornafius, J. S., 1987, Neogene crustal rotations, fault slip, and basin development in southern California,
in R. V. Ingersoll and W. G. Ernst, eds., Cenozoic basin development of coastal California: Englewood Cliffs, New Jersey,
Prentice-Hall, p. 259283.
Magoon, L. B., and J. C. Taylor, 1975, The four petroleum zones of
the Ventura basin, Santa Barbara, Ventura and Los Angeles
counties, California, U. S. A.: U. S. Geological Survey Branch of
Oil and Gas Resources, 87 p.
Matti, J., and Morton, D. M., 1993, Paleogeographic evolution of the
San Andreas fault in southern California: a reconstruction based
on a new crossfault correlation, in R. E. Powell, R. J. Weldon II,
and J. C. Matti, eds., The San Andreas fault system: displacement,
palinspastic reconstruction, and geologic evolution: Boulder, Colorado, Geological Society of America Memoir 178, p. 107159.
McDougall, K., 1982, Microfossil assemblages from the Castaic
Formation, Ridge basin, southern California, in J. C. Crowell,
and M. H. Link, eds., Geologic history of Ridge basin, southern
California: Pacific Section SEPM, p. 219228.
Mefferd, M. G., 1965, Newhall-Potrero oil field: California Division
of Oil and Gas Summary of Operations, California Oil Fields,
v. 51, no. 2, p. 4151.

Muehlberger, W. R., 1958, Geology of northern Soledad basin,


Los Angeles County, California: AAPG Bulletin, v. 42,
p. 18121844.
Natland, M. L., 1952, Pleistocene and Pliocene stratigraphy of
southern California: Ph.D. thesis, University of California at
Los Angeles, Los Angeles, California, 165 p.
Nelligan, F. M., 1978, Geology of the Newhall area of the eastern
Ventura and western Soledad basins, Los Angeles County, California: M.S. thesis, Ohio University, Athens, Ohio, 117 p.
Nilsen, T. H., and S. H. Clarke, Jr., 1975, Sedimentation and tectonics in the early Tertiary continental borderland of central California: U. S. Geological Survey Professional Paper 925, 64 p.
Oakeshott, G. B., 1950, Geology of Placerita oil field, Los Angeles
County, California: California Journal of Mines and Geology,
v. 46, p. 4379.
Oakeshott, G. B., 1958, Geology and mineral deposits of San Fernando quadrangle, Los Angeles County, California: California
Division of Mines Bulletin 172, 147 p.
Obradovich, J. D., and C. W. Naeser, 1981, Geochronology bearing on the age of the Monterey Formation and siliceous rocks
in California, in R. E. Garrison and R. G. Douglas, eds., The
Monterey Formation and related siliceous rocks of California:
Pacific Section SEPM, p. 8796.
Paschall, R. H., and T. Off, 1961, Dip-slip versus strike-slip movement on the San Gabriel fault, southern California: AAPG Bulletin, v. 45, p. 19411956.
Pechmann, J. C., 1987, Tectonic implications of small earthquakes in the central Transverse Ranges: U. S. Geological Survey Professional Paper 1339, p. 97111.
Pierce, R. L., 1972, Revaluation of the late Miocene biostratigraphy of California; summary evidence: Proceedings of the Pacific Coast Miocene Biostratigraphic Symposium, Pacific Section
of SEPM, p. 334340.
Powell, R. E., 1993, Balanced palinspastic reconstruction of prelate Cenozoic paleogeology, southern California: geologic and
kinematic constraints on evolution of the San Andreas fault
system, in R. E. Powell, R. J. Weldon II, and J. C. Matti, eds.,
The San Andreas fault system: displacement, palinspastic
reconstruction, and geologic evolution: Geological Society of
America Memoir 178, p. 1106.
Ricketts, E. W., and K. R. Whaley, 1975, Structure and stratigraphy of the Oak Ridge faultSanta Susana fault intersection,
Ventura basin, California: M.S. thesis, Ohio University, Athens,
Ohio, 81 p.
Sage, O. G., Jr., 1973, Paleocene geography of California: Ph.D. dissertation, Univ. of California, Santa Barbara, California, 250 p.
Sarna-Wojcicki, A. M., K. B. Lajoie, C. E. Meyer, D. P. Adam, and
H. J. Rieck, 1991, Tephrochronologic correlation of upper
Neogene sediments along the Pacific margin, conterminous
United States, in R. B. Morrison, ed., Quaternary nonglacial
geology: Conterminous U. S.: Geological Society of America,
The Geology of North America, v. K2, p. 117140.
Saul, R. B., 1975, Geology of the southeast slope of the Santa
Susana Mountains and geologic effects of the San Fernando
earthquake: California Division of Mines and Geology Bulletin
196, p. 187194.
Saul, R. B., and T. M. Wootton, 1983, Geology of the south half of
the Mint Canyon Quadrangle, Los Angeles County, California:
California Division of Mines and Geology OpenFile Report
8324LA, 139 p.
Schlaefer, J. T., 1978, The subsurface geology of the Honor Rancho area of the east Ventura and western Soledad basins, California: M.S. thesis, Ohio University, Athens, Ohio, 86 p.
Seedorf, D. C., 1983, Upper Cretaceous through Eocene subsurface stratigraphy, Simi Valley and adjacent regions, California,
in R. L. Squires and M. V. Filewicz, eds., Cenozoic geology of
the Simi Valley area, southern California: Pacific Section SEPM
Field Trip Volume and Guidebook, p. 109128.
Shepherd, G. L., 1960, Geology of the Whitaker PeakCanton
Canyon area, southern California: M.A. thesis, University of
California at Los Angeles, Los Angeles, California, 64 p.

Yeats et al.

Shields, K. E., 1977, Structure of the northwestern margin of the


San Fernando Valley, Los Angeles County, California: M.S. thesis, Ohio University, Athens, Ohio, 82 p.
Silver, L. T., 1966, Preliminary history of the crystalline complex
of the central Transverse Ranges, Los Angeles County, California (abs.): Geological Society of America Abstracts with Programs, v. 8, p. 201.
Smith, D. P., 1977, San JuanSt. Francis faulthypothesized major
middle Tertiary right-lateral fault in central and southern California: California Division of Mines and Geology Special
Report 129, p. 4150.
Smith, D. P., 1984, Geology of the northeast quarter of the
Newhall quadrangle, Los Angeles County, California: California Division of Mines and Geology Open File Report 8449 LA,
42 p.
Squires, R. L., 1977, Middle Eocene molluscan assemblage and
stratigraphy, lower Piru Creek, Transverse Ranges, California:
California Division of Mines and Geology Special Report 129,
p. 8186.
Stanton, R. J., Jr., 1966, Megafauna of the upper Miocene Castaic
Formation, Los Angeles County, California: Journal of Paleontology, v. 40, p. 2140.
Stanton, R. J., R., 1982, Molluscan paleoecology and depositional
environment of the Miocene Castaic Formation, Ridge basin,
southern California, in J. C. Crowell and M. H. Link, eds., Geologic history of Ridge basin, southern California: Pacific Section SEPM, p. 211218.
Stitt, L. T., 1980, Geology of the Ventura and Soledad basins in the
vicinity of Castaic, Los Angeles County, California: M.S. thesis,
Oregon State University, Corvallis, Oregon, 124 p.
Stitt, L. T., 1982, Structure and late Miocene paleogeography of
southern Ridge basin, southern California, in J. C. Crowell and
M. H. Link, eds., Geologic history of Ridge basin, southern California: Pacific Section SEPM, p. 4352.
Stitt, L. T., 1986, Structural history of the San Gabriel fault and
other Neogene structures of the central Transverse Ranges,
California, in P. L. Ehlig, compiler, Guidebook and volume,
neotectonics and faulting in southern California: Geological
Society of America, p. 43102.
Stitt, L. T., and R. S. Yeats, 1982, A structural sketch of the Castaic
area, northeastern Ventura basin and northern Soledad basin,
California, in D. L. Fife and J. A. Minch, eds., Geology and mineral wealth of the California Transverse Ranges: South Coast
Geological Society, p. 390394.
Sullwold, H. H., Jr., 1960, Tarzana fan, deep submarine fan of late
Miocene age, Los Angeles County, California: AAPG Bulletin,
v. 44, p. 433457.
Suppe, J., and D. A., Medwedeff, 1990, Geometry and kinematics
of faultpropagation folding: Eclogae Geologicae Helvetiae,
v. 83, p. 409454.
Terres, R. R., and B. P., Luyendyk, 1985, Neogene tectonic rotation of the San Gabriel region, California, suggested by paleomagnetic vectors: Journal of Geophysical Research, v. 90,
p. 1246712484.
Treiman, J. H., 1986, Geological hazards in the west half of the
Newhall quadrangle, Los Angeles County, California: California Division of Mines and Geology Open-File Report 86-6 LA.
Turner, D. L., 1970, Potassiumargon dating of Pacific Coast
Miocene foraminiferal stages: Geological Society of America
Special Paper 124, p. 91129.
Vedder, J. G., H. C. Wagner, and J. E. Schoellhamer, 1969, Geologic framework of the Santa Barbara Channel Region: U. S. Geological Survey Professional Paper 679-A, p. 111.
Walker, R. G., 1978, Deepwater sandstone facies and ancient
submarine fans: models for exploration for stratigraphic traps:
AAPG Bulletin, v. 62, p. 932966.
Weber, F. H., Jr., 1979, Geologic and geomorphic investigation of
the San Gabriel fault zone, Los Angeles and Ventura Counties,
California: Final Technical Report to U.S. Geological Survey,
Contract No. 14-08-0001-16600, Modification 1, 78 p.
Weber, F. H., Jr., 1982, Geology and geomorphology along the

1073

San Gabriel fault zone, Los Angeles and Ventura Counties, California: California Division of Mines and Geology Open File
Report 82-2 LA.
Weber, F. H., Jr., 1986, Geological relationships along the San
Gabriel fault between Castaic and the San Andreas fault, Kern,
Los Angeles, and Ventura counties, California, in P. L. Ehlig,
compiler, Neotectonics and faulting in southern California:
Geological Society of America Cordilleran Section Annual
Meeting Guidebook and Volume, Los Angeles, California State
University Department of Geology, p. 109122.
Weigand, P. W., 1982, Cenozoic volcanism of the western Transverse Ranges, in D. L. Fife and J. A. Minch, eds., Geology and
mineral wealth of the California Transverse Ranges: South
Coast Geological Society Annual Symposium and Guidebook
10, p. 170188.
Whitcomb, J. H., C. R. Allen, J. D. Garmany, and J. A. Hileman,
1973, San Fernando earthquake series, 1971: focal mechanisms and tectonics: Reviews of Geophysics and Space
Physics, v. 11, p. 693730.
Winterer, E. L., and D. L. Durham, 1954, Geology of a part of the
eastern Ventura basin, Los Angeles, County: California Division of Mines Bulletin 170, map sheet 5.
Winterer, E. L., and Durham, D. L., 1962, Geology of the southeastern Ventura basin, Los Angeles County, California: U. S.
Geological Survey Professional Paper 334-H, p. 275366.
Woodburne, M. O., 1975, Cenozoic stratigraphy of the Transverse
Ranges and adjacent areas, southern California: Geological
Society America Special Paper 162, 91 p.
Yeats, R. S., 1979, Stratigraphy and paleogeography of the Santa
Susana fault zone, Transverse Ranges, California, in J. M.
Armentrout, M. R. Cole, and H. Ter Best, eds., Cenozoic paleogeography of the western United States: Pacific Coast Paleogeography Symposium 3, p. 191204.
Yeats, R. S., 1983, Large-scale Quaternary detachments in Ventura
basin, southern California: Journal of Geophysical Research,
v. 88, p. 569583.
Yeats, R. S., 1987a, Late Cenozoic structure of the Santa Susana
fault zone: U. S. Geological Survey Professional Paper 1339,
p. 137160.
Yeats, R. S., 1987b, Changing tectonic styles in Cenozoic basins of
southern California, in R. V. Ingersoll and W. G. Ernst, eds.,
Cenozoic basin development of coastal California: Rubey Volume VI, p. 284298.
Yeats, R. S., 1988, Late Quaternary slip rate on the Oak Ridge
fault, Transverse Ranges, California: implications for seismic
risk: Journal of Geophysical Research, v. 93, p. 1213712149.
Yeats, R. S., 1989, Oak Ridge fault, Ventura basin, California: U. S.
Geological Survey Open-File Report OFR 89-343, 30 p.
Yeats, R. S., and J. M. Beall, 1991, Stratigraphic controls of oil
fields in the Los Angeles basin: a guide to migration history, in
K. T. Biddle, ed., Active margin basins: AAPG Memoir 52,
p. 221235.
Yeats, R. S., M. L. Butler, and J. C. Schlueter, 1977, Geology of the
central Santa Susana fault area, Ventura and Los Angeles counties, California: Part III of Final Technical Report to U. S. Geological Survey, Contract 14-08-0001-15886, 29 p.
Yeats, R. S., J. W. McDougall, and L. T. Stitt, 1985, Cenozoic structure
of the Val Verde 712minute quadrangle and south half of the
Whitaker Peak 712 minute quadrangle, California: U. S. Geological
Survey Open-File Report 85-587, 32 p., map scale 1:24,000.
Yeats, R. S., W. H. K. Lee, and R. F. Yerkes, 1987, Geology and
seismicity of the eastern Red Mountain fault, Ventura County:
U. S. Geological Survey Professional Paper 1339, p. 161167.
Yeats, R. S., J. A. Calhoun, B. B. Nevins, H. F. Schwing, and H. M.
Spitz, 1989, The Russell fault: an early strike-slip fault of the
California Coast Ranges: AAPG Bulletin, v. 73, p. 10891102.
Zoback, M. D., M. L. Zoback, V. S. Mount, J. Suppe, J. P. Eaton,
J. H. Healy, D. H. Oppenheimer, P. A. Reasenberg, L. M. Jones,
C. B. Raleigh, I. G. Wong, O. Scotti, and C. M. Wentworth,
1987, New evidence on the state of stress of the San Andreas
fault system: Science, v. 238, p. 11051111.

1074

East Ventura Basin

ABOUT THE AUTHORS


Robert S. Yeats

Gary J. Huftile

Robert S. Yeats is professor of geology and oceanography at Oregon


State University at Corvallis. He
received his B.A. degree in geography
from the University of Florida in 1952
and his Ph.D. in geology from the University of Washington in 1958. He
worked for Shell Oil Company in
southern California as an exploitation
engineer, production geologist, and
exploration geologist until 1967,
when he joined the faculty at Ohio University in Athens. In
1977, he moved to Oregon State University. His research
interests are in tectonics and earthquake potential of active
margins, focusing on southern California and offshore Oregon and Washington. He is chairman of a task group in paleoseismology in the International Lithosphere Program, and
he is writing a book, Geology of Earthquakes, to be published by Oxford University Press.

Gary J. Huftile is a postdoctoral


research associate at Oregon State
University. He specializes in the calculation of deformation rates across
active geologic structures. He graduated from the University of California at
Davis in 1982 with a B.S. degree in
geology. He worked as an exploration
geophysicist for Gulf Oil Exploration
and Production Company for three
years. He received his M.S. degree
(1988) and Ph.D. (1992) in geology from Oregon State University.
Leonard T. Stitt
Leonard T. Stitt is a geologist with
Rancho Energy Consultants. His main
emphasis is exploration geology,
especially regional basin analysis. He
has a B.S. degree (1977) from the University of Cincinnati and a M.S.
degree (1980) from Oregon State University. His research emphasis has
been the structural history of southern California. Recent work with Rancho Energy includes regional evaluations of the subsurface geology and exploration potential of
the Ventura basin and of the Los Padres National Forest.

Das könnte Ihnen auch gefallen