SlideShare a Scribd company logo
Student Name: Bud Benneman
Geology 105 Spring 2020
Paper Outline
The Newport–Inglewood fault zone (NIFZ) of southern
California.
I. The Newport–Inglewood fault zone (NIFZ) was first
identified as a significant threat to southern California residents
in 1933 when it generated the Magnitude 6.3 Long Beach
Earthquake, killing 115 people.
A. The Newport Inglewood fault is located in southern Los
Angeles County in the city of Inglewood and transverses south
to Newport Beach in Orange County where it becomes an off
shore fault.
B. The NIFZ then connects to the Rose Canyon Fault none,
becomes a landward fault in San Diego County.
C. This is a stress reliever Strike-Slip Fault Zone associated
with the San Andréas Fault Zone.
D. Ground water basins in Los Angeles County may be used as
predictors of fault movement due to sudden changes in ground
water level.
E. In southern California, tectonic deformation between
the Pacific and North American plates is accommodated
primarily by a zone of strike-slip faults,
II. This fault is important to the exploration of Oil.
A. The NIFZ has been studied extensively in the Los Angeles
basin by petroleum geologists.
B. The NIFZ overlies a major tectonic boundary separating
eastern continental basement rocks of granitic and associated
metamorphic rocks of Santa Catalina Island schist.
C. The NIFZ follows along a former Mesozoic subduction zone.
D. This fault is associated with several oil basins including
Newport Beach, Huntington Beach, Seal Beach, Long Beach,
and Signal Hill.
III. Earthquake Potential for Los Angeles, Orange and San
Diego Counties.
A. The epicenter for the 1933 Long Beach Quake was located
in Newport Beach just south of the Santa Ana River discharge
into the Pacific Ocean.
B. Orange County in 1933 consisted of small farm-towns, cattle
ranching and agricultural fields.
C. The closest city of significant development was Lang Beach,
which was devastated by the 1933 6,3 quake.
D. Today cities of Newport Beach, Huntington Beach, Costa
Mesa, are large population centers, which have replaced
agriculture as the primary economic sector.
E. The rate of ground water basin contraction is important in
determining possible earthquake releases in the Los Angeles
Basin.
F. Many Orange and Los Angeles county buildings are located
along high risk development zones during earthquakes.
IV. Earthquake Monitoring and Prediction
a. The location of the 2000 cluster is southwest of an area of
active faulting from 1982 to 1990 along this zone and reported
strike-slip offshore fault between Newport Beach and the San
Joaquin Hills was traced from seismic activity.
b. Analyzed microseismicity from a cluster of epicenters
between 2 and 2.5 km along the fault zone suggest potential
earthquakes of higher magnitude are possible.
c. A possible 7.5 magnitude earthquake could result from
rupture of the entire fault. The 6.5-km depth of the Newport
Beach seismicity cluster does not provide information on the
geometry or interaction between the strike-fault zone and
Oceanside thrust faults.
d. Los Angeles groundwater basin contraction could cause faults
such as the Whittier
E. Oil companies and water basin studies using seismic
studies and human
induced micro-seismic studies have identified several fault
structures such as
blind thrust faults in the Los Angeles basin. Blind thrust
along with strike slip
fault structures suggest potential earthquakes of higher
magnitude are
possible for the Los Angeles and Orange County basins.
IV. Conclusion
a. The Newport Inglewood Fault is a major stress reliever for
the San Andreas Fault Zone.
b. This fault zone is tied in between the North American Plate
and Pacific Plates Strike Slip Zones and follows a Mesozoic
Subduction zone.
c. The location of the NIFZ has potential for a large magnitude
earthquake due to it being sandwiched between the eastern San
Andreas and Pacific Margin plate Boundary.
Effects of episodic fluid flow on hydrocarbon migration in
the Newport-Inglewood Fault Zone, Southern California
B. JUNG1, G. GARVEN 2 AND J. R. BOLES3
1Department of Earth Sciences, Uppsala University, Uppsala,
Sweden; 2Department of Earth and Ocean Sciences, Tufts
University, Medford, MA, USA; 3Department of Earth Science,
University of California, Santa Barbara, CA, USA
ABSTRACT
Fault permeability may vary through time due to tectonic
deformations, transients in pore pressure and effective
stress, and mineralization associated with water-rock reactions.
Time-varying permeability will affect subsurface
fluid migration rates and patterns of petroleum accumulation in
densely faulted sedimentary basins such as those
associated with the borderland basins of Southern California.
This study explores the petroleum fluid dynamics of
this migration. As a multiphase flow and petroleum migration
case study on the role of faults, computational
models for both episodic and continuous hydrocarbon migration
are constructed to investigate large-scale fluid
flow and petroleum accumulation along a northern section of
the Newport-Inglewood fault zone in the Los
Angeles basin, Southern California. The numerical code solves
the governing equations for oil, water, and heat
transport in heterogeneous and anisotropic geologic cross
sections but neglects flow in the third dimension for
practical applications. Our numerical results suggest that fault
permeability and fluid pressure fluctuations are cru-
cial factors for distributing hydrocarbon accumulations
associated with fault zones, and they also play important
roles in controlling the geologic timing for reservoir filling.
Episodic flow appears to enhance hydrocarbon accu-
mulation more strongly by enabling stepwise build-up in oil
saturation in adjacent sedimentary formations due to
temporally high pore pressure and high permeability caused by
periodic fault rupture. Under assumptions that
fault permeability fluctuate within the range of 1–1000
millidarcys (10�15–10�12 m2) and fault pressures fluctuate
within 10–80% of overpressure ratio, the estimated oil volume
in the Inglewood oil field (approximately 450 mil-
lion barrels oil equivalent) can be accumulated in about 24 000
years, assuming a seismically induced fluid flow
event occurs every 2000 years. This episodic petroleum
migration model could be more geologically important
than a continuous-flow model, when considering the observed
patterns of hydrocarbons and seismically active
tectonic setting of the Los Angeles basin.
Key words: episodic fluid flow, fluid flow in faults, multiphase
flow in siliciclastic sedimentary basins, petroleum
migration
Received 21 May 2013; accepted 16 October 2013
Corresponding author: Byeongju Jung, Department of Earth
Sciences, Uppsala University, Gl227 Geocentrum,
Villav€agen 16B, 753 36 Uppsala, Sweden.
Email: [email protected] Tel: +46 018 471 2264. Fax: +1 617
627 3584.
Geofluids (2014) 14, 234–250
INTRODUCTION
Large-scale faults in sedimentary basins have become
increasingly studied due to their important role in convey-
ing and compartmentalizing hydrocarbons (Aydin 2000;
Boles et al. 2004; Karlsen & Skeie 2006; Kroeger et al.
2009; Zhang et al. 2009; Gong et al. 2011). The hydro-
mechanical properties of faults in active continental mar-
gins are strongly affected by tectonic deformation, so
considering the fluid dynamics of faults will likely improve
our understanding of petroleum migration (Reynolds &
Lister 1987; Blanpied et al. 1992; Sibson 1994; Appold &
Garven 2000; Yamaguchi et al. 2011). For example, there
is abundant geological evidence, at both macroscopic and
microscopic scales, that faults focus fluid flow over long
periods of time but later are sealed by mechanical compac-
tion and chemical reactions causing mineral precipitation
(Eichhubl & Boles 2000; Caine et al. 2010; Faulkner et al.
2010). The hydrologic activity of fault zones may also
depend highly on earthquakes, which in turn may induce
periodic fluctuations in pore fluid pressure and fault per-
meability (Evans 1992; Sibson 1994). Furthermore, large-
© 2013 John Wiley & Sons Ltd
Geofluids (2014) 14, 234–250 doi: 10.1111/gfl.12070
scale fault zones may affect regional hydrocarbon migration
by regulating the spatial distribution of overpressure in the
subsurface (Sperrevik et al. 2002; Fisher et al. 2003; Sork-
habi & Tsuji 2005). Laboratory experiments on fractured
rock show that active shear faults are more permeable than
the adjacent country rock by two to three orders in magni-
tude. These faults then become less permeable when deac-
tivated (Aydin 2000).
The mechanism of recurring fluid pressure build-up, hy-
drofracturing, fluid surge, and fault sealing can also be a
potential means for hydrocarbon migration (Bradley 1975;
Walder & Nur 1984; Mandl & Harkness 1987). For exam-
ple, field observations of brecciated rocks and hydrother-
mal veins from the Stillwater fault zone in Nevada indicate
that petroleum migration was not completed as one single
flow event, but rather accumulation too place over many
episodes of oil flow during the deformation history (Caine
et al. 2010). Geochemical evidence from hydrocarbon con-
densates found in the South China basin also support the
notion that petroleum migration occurs simultaneously
with episodes of hydrothermal fluid flow (Guo et al.
2011).
We further hypothesize that the hydrodynamic effects of
multiphase flow are more effective for long-distance trans-
port, during periods of strong overpressuring associated
with episodes of seismically controlled fluid flow. Episodic
flow associated with large faults may also be more effective
than long-term continuous or steady flow of hydrocarbons,
as might be envisioned for a slowly subsiding sedimentary
basin. To test this hypothesis, we conduct 2D finite ele-
ment simulations for multiphase flow in geologically com-
plex cross sections through a basin. We introduce two
migration scenarios of continuous flow and of episodic
flow, which likely account for the current distribution of
hydrocarbon pools such as the Inglewood oil field. The
continuous models assume constant fault permeability and
fluid pressure throughout the simulation time, while those
conditions are time varying in the episodic models. We first
compared the fluid pressure, subsurface temperature, and
petroleum saturations from these models and then per-
formed sensitivity studies on the fault permeability and the
frequency of episodic flow pulses to understand how these
permeability transients might affect overall hydrocarbon
migration and accumulation patterns in a faulted sedimen-
tary basin.
GEOLOGIC SETTING
The Los Angeles (LA) basin is one of the most prolific
hydrocarbon-producing areas on Earth, and it hosts histor-
ically giant oil fields. From the geological survey, it was
recognized that the upper Miocene formations contain
organic-rich sediments and play an important role as the
primary hydrocarbon source rock. Most of the hydrocar-
bons were thermally matured in the central part of the
basin (the central syncline area) and have migrated to the
edges, which are laterally confined by regional-scale fault
zones (Biddle 1991; Jeffrey et al. 1991). The Newport-
Inglewood fault zone is one of the major regional fault sys-
tems that structurally border the southwestern side of the
LA basin, and numerous hydrocarbon reservoirs are closely
associated with the fault structure (Fig. 1A). In the Ingle-
(A) (B)
Fig. 1. Faults and oil fields in the Los Angeles basin (after
Wright 1991): (A) LA basin, (B) Inglewood oil field.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
Effects of episodic fluid flow 235
wood oil field located on the northern part of the fault
zone, the main production area exists at the intersection
between the Newport-Inglewood and Sentous faults and is
elongated along the trends of both faults (Fig. 1B). Geo-
chemical indicators and biomarkers also suggest the hydro-
carbons discovered in this area have mostly migrated
approximately 10–15 km from the deep basin center with
minor mixing of indigenous petroleum (Kaplan et al.
2000). Tectonic deformation and subsequent seismic activ-
ity make this region attractive for studying the relationship
between hydrocarbon migration and active fault structures
in young sedimentary basins.
The Newport-Inglewood fault zone consists of a series
of en echelon strike slip faults that were reactivated during
the late Pliocene transpressional deformation (Pasadena
Orogeny) and transformed into more complicated anticline
structures containing normal and reverse faults (Fig. 2).
Hydrocarbon reservoirs exist in multiple sedimentary for-
mations but are more concentrated in Pliocene strata con-
taining a high proportion of sandstone. Many productive
petroleum reservoirs align with the trend of the Newport-
Inglewood fault zone, which extends and merges with the
Central Basin d�ecollement at about 10 km depth (Shaw &
Suppe 1996).
The permeability of these sandstones ranges from 10’s
to 1000’s millidarcys (10�14–10�12 m2) for sandy Plio-
cene formations (Hesson & Olilang 1990). The thickness
of sediments in the central syncline is approximately
10 km and becomes gradually thinner toward the north-
ern and southwestern edges of the basin (Blake 1991;
Fig. 3). Figure 4A shows an outcrop of channel-fill sand-
stones (Sespe Formation) in the adjacent Santa Barbara
area. The faults in these rocks are filled with carbonate
mineral precipitates and lithified hydrocarbons, which is
strong evidence that the fault zone provided active chan-
nels for hydrocarbon fluids, but later these were sealed
by subsequent reactions involving diagenetic-hydrother-
mal mineralization as fluids cooled or the pressure rap-
idly dropped (Fig. 4B). Five separate hydrogeologic units
were considered here: middle and pper Miocene, lower
and upper Pliocene, and Pleistocene formations (Fig. 3).
The middle Miocene unit (Topanga Formation) consists
of medium to coarse sandstones with intercalated shale
layers. The upper Miocene sediments (Puente Formation)
are mostly siltstone and silty sandstone with interbedded
pelagic mudstone and shale layers (nodular shales in
some areas) that contain high organic carbon contents
(10–16%). This formation is often considered to be the
primary source rock for petroleum generation (Jeffrey
et al. 1991). The lower Pliocene unit (Repetto forma-
tion) serves as a major reservoir for hydrocarbon accu-
mulation. The Repetto consists of fine to coarse
sandstones with interbedded siltstone and shale layers
that have relatively high permeability of 10–100 md
(10�14–10�13 m2) (Higgins & Chapman 1984). The
geology of the upper Pliocene unit (Pico formation) is
very similar to the Repetto formation, consisting of inter-
bedded sandstone and siltstone, but with slightly lower
permeability. The Pleistocene unit (San Pedro formation)
consists of relatively uniform and highly permeable sand
layers interbedded with minor gravel, silt, and shale lay-
ers (Olson 1978).
We propose that north–south transpressional tectonic
stresses pressurized the basin continuously from the late
Pliocene to the present. A coupled mechanism of tec-
tonic loading, pore fluid pressure build-up, fault instabil-
ity, and fluid flow may have induced episodic fluid flow
events (Sibson 1994). Elevated effective stress and pore
pressure by tectonic loading compact the sedimentary
rocks during the interseismic periods, making the fault
zone mechanically unstable. When ruptured, high perme-
ability and low pore pressure temporarily create focused
fluid flow in the fault zone. The fault stays open for a
relatively short period of time due to hydromechanical
compaction as the pore pressure decreases, and then
sealed further by hydrothermal mineral precipitation
(Fig. 5). The continuous tectonic stress and rebuilding of
pore pressure cause this cycle to repeat after the fault
zone becomes sealed. Field observations of petroleum
and carbonate mineral deposits in other siliciclastic faults
suggest an episodic nature of injected fluids (Eichhubl &
Boles 2000; Garden et al. 2001), and these episodic flow
patterns may affect the geohydrologic controls on hydro-
carbon accumulation. We modeled this dynamic aspect of
fault zone to understand its effects on petroleum migra-
tion and entrapment.
Fig. 2. Cross section of the Newport-Inglewood fault zone along
the tran-
sect X–X′ (courtesy of Plains Exploration and Production
Company, 2008).
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
236 B. JUNG et al.
MULTIPHASE FLOW MODEL
To model petroleum migration in the basin, we need to
predict fluid saturation and pressure for both the wetting
phase (formation water) and nonwetting phase (oil). The
governing equations for multiphase fluid flow can be
derived from the fluid mass conservation equations, as
written by Bear (1972):
(A) (B)
Fig. 4. Outcrop pictures of organic-rich sedimentary rocks in
California: (A) channel-fill sandstones in nonmarine Sespe
Formation (Oligocene) from Old San
Marcos road, Santa Barbara County, CA, (B) tar-filled fault
breccia in Monterey Formation at Arroyo Burro beach, Santa
Barbara County, CA.
Fig. 3. Lithostratigraphy of the Los Angeles
basin, modified from Blake (1991).
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
Effects of episodic fluid flow 237
/
@ðqwSwÞ
@t
¼ �O � ðqwvwÞ þ qwmw ð1Þ
/
@ðqnSnÞ
@t
¼ / @ðqnð1 � SwÞÞ
@t
¼ �O � ðqnvnÞ þ qnmn ð2Þ
where / is the effective porosity of a formation, and S is
the saturation of the phase: the subscripts n and w denote
the nonwetting (liquid petroleum) and wetting (water)
phases, respectively. Additionally, q is mass density of the
fluid, m is a source/sink term, and v is the Darcy velocity
(specific discharge) of each fluid phase, expressed as fol-
lows:
vw ¼ �kwkðOqw � qwgÞ ð3Þ
vn ¼ �knkðOqn � qngÞ ¼ �knðOqw þ Oqc � qgÞ ð4Þ
where k is the intrinsic permeability tensor, p is fluid pres-
sure, g is a gravitational vector (g = (0, 0, �g)), and pc is
capillary pressure (pc = pn – pw). The parameter, k, is the
mobility coefficient and defined by the ratio of relative per-
meability (kr) and dynamic viscosity (l) as:
kw ¼ krw=lw; kn ¼ krn=ln ð5Þ
After a few steps of algebraic manipulation, the pressure
and saturation equations can be decoupled. If slightly com-
pressible fluids are assumed, the final form of average pres-
sure and saturation equations can be written as follows
(Geiger et al. 2004):
/ct
@ �P
@t
¼ O � kfktO�P � 0:5ðkw � knÞOpc�
ðkwqw þ knqnÞggþmt
ð6Þ
/
@Sn
@t
¼ O � ½fnvt � �kkfOqc þ ðpw � pnÞgg� � mt ¼ 0 ð7Þ
where ct is bulk compressibility of a medium, and mt is a
source/sink term. vt is the sum of water and petroleum
velocites (vt = vw + vn). The average pressure �P is an arith-
metic mean of the water and petroleum pressure, and f is a
fractional flow coefficient that is also defined for simplicity:
fw ¼
kw
kt
; fn ¼
kn
kt
; and �k ¼ kwkn
kt
ð8Þ
The first, second, and third terms in the right-hand side
of the pressure equation (Eq. 6) represent advection, capil-
larity, and buoyancy flow terms, respectively.
Conventional multiphase flow equations were decoupled
in terms of average pressure and petroleum saturation.
These equations were solved using the implicit-pressure
explicit-saturation (so called ‘IMPES’) technique (Helmig
1997; Huber & Helmig 2000; Class et al. 2002; Reichen-
berger et al. 2006), a technique that produces solutions
faster than those requiring time-consuming nonlinear itera-
tions.
Solution
s to the pressure and saturation equations
were computed using a hybrid numerical method called
FEFVM suggested by Geiger et al. (2004, 2006). This
method applies a finite element method (FEM) for com-
puting average pressure and then a finite volume method
(FVM) for computing fluid saturation. A fully upwind for-
mulation and total velocity diminishing (TVD) method
(Harten 1997) were also used for solutions to avoid both
numerical dispersion and spurious oscillation at the satura-
tion front.
Capillary pressure and relative permeability models devel-
oped by van Genuchten (1980) were used for describing
two-phase fluid–solid interaction in the porous media.
The van Genuchten model has been widely used in a
multiphase flow modeling and well known for providing
stable numerical solutions when applying continuous capil-
lary pressure functions for a whole saturation interval.
Stress or temperature dependent parameters were not
included in the numerical formulation.
Petroleum density and viscosity were computed using
empirical equations suggested by Glasø (1980) and Eng-
land et al. (1987). Liquid petroleum density generally
decreases with increasing pressure because the solubility of
gaseous component increases. The dynamic viscosity of oil
Fig. 5. Coupled hydromechanic and hydrother-
mal processes for episodic fluid flow (as known
as fault-valve mechanism), adopted from Sibson
(1994). The parameter and variables: pf is fluid
pressure; s is shear stress along the fault; C is
the cohesive strength of the fault; ls is the sta-
tic coefficient of rock friction, and rn is the nor-
mal stress on the fault.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
238 B. JUNG et al.
varies between 5 9 10�4 and 5 9 10�2 Pa�s, and generally
decreases exponentially with increasing temperature (Eng-
land et al. 1987). Fluid properties of the formation water
were obtained from a set of state equations proposed by
Phillips et al. (1981, 1983) and Watson et al. (1980),
which consider the effects of temperature, pressure, and
salinity/NaCl concentration on water density.
MODEL CONFIGURATION
The models, in this research, are based on the cross section
in Fig. 2, which illustrates the present geology of the New-
port-Inglewood fault zone and associated oil fields (e.g.,
Inglewood oil field) along the transect X–X′ in Fig. 1B.
This profile, which consists of anticline folds with heavily
faulted rocks, was chosen because it represents the typical
geologic settings of oil reservoirs along the NIFZ. The
numerical grid was discretized into 7655 triangular ele-
ments using a Delaunay triangulation method for detailed
rendering of geological structures (de Berg et al. 2008;
Fig. 6). The Newport-Inglewood fault zone and surround-
ing areas were divided into smaller elements to increase the
resolution of the numerical solution.
Because this model is only two-dimensional, a character-
istic fault length or effective flow field length was intro-
duced for petroleum volume calculation. The total length
of Inglewood Oil Field is about 5 km, so we assume that
20% of the total fault length (Approximately 1 km) in pro-
file is available for petroleum migration. The width of fault
zone in the numerical grid is approximately 50 m. Model
parameters used in the simulations are listed in Table 1.
Permeability and porosity of each hydrogeologic unit were
obtained from several publications (Yerkes 1972; Olson
1978; Higgins & Chapman 1984; Hesson & Olilang
1990; Nishikawa et al. 2009) and chosen within the ranges
considered to be representative for these local formations.
Fault permeability was not available from any local field
measurements, so it was systematically varied or dynami-
cally changed as part of model parameter sensitivity analy-
sis. Anisotropic permeability ratio values of up to 100:1
(kx/kz), typical for a regional-scale flow, were chosen for
most formations except the Pleistocene sediments, which
are known to be the most permeable and yet not fully con-
solidated. A thermal dispersivity a approximately 100 m
was assumed for all formations, reflecting the longitudinal
solute dispersivity value for a regional groundwater flow
system (de Marsily 1986; Gelhar et al. 1992). Matrix ther-
mal conductivity values of 3.0 W m�1°C were assigned to
most sandstone-dominant units and values of 2.5–
2.8 W m�1°C were assigned to the units having high con-
tent of siltstone and interbedded shale (Blackwell & Steele
1989). A specific heat capacity value of 750 J kg�1°C, typi-
cal of sandstone and shale, was assigned for all hydrogeo-
logic units and faults (Sabins 1997). Formation of water
salinity in the LA basin usually ranges from 20 000 to
34 000 ppm TDS (Hesson & Olilang 1990; California
Department of Conservation 1992) for most oil fields. The
groundwater salinity was set to 25 000 ppm TDS for the
entire model profile.
Capillary pressure in porous sandstone is generally
<0.1 bar (approximately 0.01 MPa) but may increase to
tens of bars in source rocks with clay grain size (Ingebrit-
sen et al. 2006). Capillarity model parameters were chosen
within the range of typical rock types obtained from other
publications (Levorsen 1967; Wendebourg 1994; Bloom-
field et al. 2001). Typically, the capillary pressure of more
permeable formations exhibit lower values, but sharp
increases occur near the irreducible water saturation point
(Swr). The sum of water and petroleum relative permeabil-
ity is usually less than one when both phases are mobile.
Initially, hydrostatic conditions and conductive thermal
profile were assumed throughout the basin. The levels of
overpressure were chosen considering that the pore pres-
sure in the Southern California faults may approach
lithostatic pressure due to mechanical compaction (Gra-
tier et al. 2002, 2003). The values of overpressure ratio,
fault permeability, and the period of episodic flow pulses
are presented in Table 2. The over pressure ratio (k*) in
a sedimentary basin can be defined as follows (Wang
et al. 2006)
Fig. 6. Model numerical grid, boundary conditions and
hydrostratigraphy
of the Newport-Inglewood fault zone based on the cross section
along the
transect X–X’ (Fig. 1b). The upper margin is the prescribed
pressure head
of 200 m and the isothermal boundary condition of 4°C. The left
and right
margins of the grid were assigned to be hydrostatic for pressure
and ther-
mally insulated (no flow) for heat. The bottom margin were
assigned as no
flow and a constant temperature of 160°C. Overpressure is
applied to the
right end of the fault boundary (marked as a white arrow) as a
prescribed
pressure condition, and the petroleum saturation at this
boundary is con-
stant (Sn = 0.6).
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
Effects of episodic fluid flow 239
k� ¼ P � PF
PL � PF
� 100 ð9Þ
where P is pore fluid pressure in the formation, PF is
hydrostatic pressure, and PL is lithostatic pressure. Porosity
of the fault also changes dynamically in the range of 0.1–
0.3, accompanying with the permeability variation.
In middle Miocene to early Pleistocene, the basin is still
under a shallow marine environment (Wright 1991), so
the prescribed pressure head of 200 m and the isothermal
boundary condition of 4°C were assigned to the top
boundary of the grid. The temperature value was chosen
within the ranges of estimated Miocene shallow seawater
temperature (0–6°C) (Zachos et al. 2001). Boundary con-
ditions along the bottom margin of the grid were
assigned as no fluid flow (impermeable) and a constant
temperature of 160°C, based on geothermal gradients
reported in this area of 35–40°C km�1 (Higgins & Chap-
man 1984; Jeffrey et al. 1991). The left and right mar-
gins of the grid were assigned to be hydrostatic for
pressure and thermally insulated for heat. Petroleum was
injected through the fault boundary at the right boundary
(white arrow in Fig. 6), and this condition is physically
possible only when we assume that most of petroleum
was generated in the deep basin center and migrated
through the fault zone.
CONTINUOUS FLOW MODEL
First, we considered a continuous hydrocarbon migration
scenario through the Newport-Inglewood fault zone,
assuming a slightly overpressured sedimentary basin envi-
ronment with constant fault zone permeability. The level
of overpressure at the fault boundary (marked as a white
arrow in Fig. 6) was held constant during the simulation,
and the fault zone was considered as a channel for releas-
ing overpressure of the basin center as well as a conduit for
petroleum migration. For Model 1, the vertical fault per-
meability kz was set to 10 md (10
�14 m2), which is within
the range of kx for the Pliocene unit and higher than that
of Upper Miocene unit. The permeability value was
inferred from the fact that a fault zone may be more per-
meable than its host rock, especially when the fault was
ruptured by shear stresses and brecciated to include sur-
rounding rocks (Aydin 2000). The change of fault perme-
ability during deformation depends on fault rock clay
content and burial history. Laboratory experimental data
show permeability increasing over 2–4 orders of magnitude
(Fisher & Knipe 2001).
Fluid pressure gradients in a compacting sedimentary
basin usually vary between hydrostatic (approximately
10.1 MPa km�1) and lithostatic pressure (approximately
23.3 MPa km�1), and the values used in the simulations
Table 1 Hydrogeologic model parameters.
Parameter Topanga Puente Repetto Pico San Pedro NIFZ
kx, Horizontal permeability (md) 3.2 0.13 18 50 50 1–1000*
kx/kz, Anisotropy 100 100 100 100 100 0.1
/, Porosity 0.16 0.17 0.23 0.33 0.30 0.20
ct, Bulk compressibility (Pa
�1) 1 9 10�10 3 9 10�9 1 9 10�10 1 9 10�10 1 9 10�10 3 9
10�9
e, Thermal dispersivity (m) 100 100 100 100 100 100
k, Bulk thermal conductivity (W m�1°C) 3.0 2.5 2.7 2.8 2.8 3.0
c, Specific heat of matrix (J kg�1 °C) 750 750 750 750 750 750
a, van Genuchten coefficient (m�1) 3.0 9 10�4 4.0 9 10�4 2.5
9 10�3 2.0 9 10�3 2.0 9 10�3 1.0 9 10�2
n, van Genuchten coefficient 5.0 4.0 6.0 4.5 4.3 8.0
Swr, Irreducible wetting phase saturation (%) 10 18 12 12 12 12
Snr, Irreducible nonwetting phase saturation (%) 6 10 7 7 7 3
*Fault zone permeabilities are varied depending on model
settings described in Table 2. Permeability and porosity values
were selected within the ranges
from a technical report written by the California Department of
Conservation (1991) and unpublished data from Plains
Exploration and Production Company.
1 md = 10�15 m2. Capillarity model parameters are obtained
from publications (see the text).
Table 2 Petroleum migration scenarios for the Newport-
Inglewood fault zone.
Model Scenario Overpressure ratio (%) Fault permeability
(md)* Period of fluid pulse (years) Petroleum saturation
1 Continuous 10 10 0.6
2 Continuous 10 5 0.6
3 Continuous 10 20 0.6
4 Continuous 10 50 0.6
5 Episodic 10–80 1–1000 3000 0.6
6 Episodic 10–80 1–1000 2000 0.6
7 Episodic 10–80 1–1000 1000 0.6
8 Episodic 10–80 1–1000 500 0.6
*1 md = 10�15 m2.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
240 B. JUNG et al.
were selected based on the current LA basin fluid pressure
gradient of 9.9–12.6 MPa km�1 (Berry 1973; McCulloh
1979). These values can be converted to �1.5 to 19.0% in
overpressure ratio. These relatively low-pressure gradients,
despite the high subsidence and sedimentation rates, seem
to originate from the high content of coarse-grained sedi-
ments that can dissipate pressure very effectively in the LA
basin (Hayba & Bethke 1995).
We assumed the whole domain was initially hydrostatic,
and then, the lower right end of the fault on the boundary
of the mesh (marked by the white arrow in Fig. 6) was set
to 10% overpressure. Petroleum saturation at the inlet of
the fault zone was maintained as a constant value of 0.6,
which is taken from the average value of Miocene strata in
the Inglewood oil field (California Department of Conser-
vation 1992).
Figure 7 illustrates the pore pressure change in terms of
hydraulic head (h) of the continuous flow model, assuming
a constant fault permeability of 10 md (10�14 m2) and
overpressure ratio of 10% during the whole simulation time
(Model 1). The overpressured areas are sustained in the
lower part of the fault zone and the Miocene formations,
but the Pliocene and Pleistocene formations remain largely
hydrostatic (colored in white). The high pore fluid pressure
change is shown in the lower part of the fault. High fluid
pressure in the fault zone is confined in the Upper Mio-
cene formation due to its low permeability. The pressure
pattern reaches the equilibrium state approximately after
40 000 years. Subsurface temperature under the continu-
ous flow model displays dominantly conductive transport
patterns with nearly horizontal isotherms, except for a
slight elevation along the lower part of the fault zone
(Fig. 8A). Elevated temperatures are the result of upward
fluid flow into the fault zone, and contrast in thermal con-
ductivity. The petroleum velocity distribution of the
domain (Fig. 8B) shows that high velocity (approximately
0.2 m year�1) area appears at the lower part of the fault
zone, and it gradually decreases moving up along the fault.
The petroleum velocity at the upper part of the fault is
<0.03 m year�1.
Petroleum migrates mainly through the fault zone, and
the upper Miocene formation provides an effective seal for
the migration (Fig. 9A). Highly saturated areas (warmer
color) are found in the boundaries between the fault zone
and the upper Miocene formation because petroleum dis-
charge away from the fault is retarded by the permeability
contrast. Petroleum continues to rise due to buoyancy and
saturates the tributary fault structures such as the Sentous
fault before smearing into the Pliocene formation near the
fault zone, and eventually vents on the sea floor (Fig. 9C,
D).
EFFECT OF FAULT PERMEABILITY
Effects of fault zone permeability on petroleum distribu-
tion were investigated by comparing four continuous
flow models (Models 1–4). The parameters assumed are
tabulated in Table 2. For each model scenario, the per-
meability and pore pressure in the fault zone are assumed
to be constant. We also compared total accumulation
time required to reach the current estimate of total oil
reserve in the Inglewood oil field: approximately 450
million barrels of oil equivalent (MMBOE = 106 barrels)
(California Department of Conservation 1992). The total
petroleum volume in a unit of barrel of oil equivalent
(BOE), computed to provide a base of comparison, was
obtained by multiplying petroleum saturation, porosity,
and characteristic length (1 km) for considering the vol-
ume as a three-dimensional domain. The total petroleum
volume within the domain can be calculated using petro-
leum saturation, porosity, finite elemental area, and char-
acteristic length:
PV ¼
XN
i¼1
Sn/Ai
!
� Lc ð10Þ
where PV is the total petroleum volume at a specified time
step, N the number of finite elements, Sn the petroleum
saturation, / is porosity, Ai is the area of the i-th finite ele-
ment, and Lc the characteristic length.
The temporal variation of total petroleum volume of
four continuous models are depicted for showing the rela-
tionship of fault permeability and flow simulation time
required for the total volume of 450 million barrels
Fig. 7. Hydraulic head change (Dh = h – h0, h0 is hydrostatic
pressure) of
the Newport-Inglewood fault zone applied with the continuous
petroleum
migration model having constant fault zone permeability of 10
md
(10�14 m2) (Model 1) at 100 000 years after starting
simulation. Overpres-
sure appears along the fault zone. The domain was initially
hydrostatic and
then the 10% overpressure was applied at the lower end of the
fault
boundary. Numbers are in meter.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
Effects of episodic fluid flow 241
(Fig. 10). Not surprisingly, the results indicate that higher
fault permeability results in faster accumulation. For exam-
ple, the reservoir takes about 62 000 years to fill when the
fault permeability is 10 md (10�14 m2), but this time per-
iod reduces to about 34 000 years when the permeability
increases to 50 md. It is also worth noting that the filling
time increases with decreasing fault permeability almost in
linear proportion. The spatial patterns of petroleum satura-
tion are also affected significantly by fault permeability.
Figure 11 shows petroleum saturations for Models 1–4
when the total volume is equal to 450 million barrels. In
the high fault permeability model (Model 4, Fig. 11D),
petroleum migrates more vigorously laterally to invade
adjacent formations through fingering and channeling
(A) (B)
Fig. 8. (A) Subsurface temperature (°C) and (B)
petroleum velocity of the fault zone (Model 1)
at 100 000 years after starting simulation. The
top margin of the grid was set as the isothermal
boundary condition of 4°C, and the bottom
margin were assigned as a constant tempera-
ture of 160°C, based on geothermal gradients
reported in this area of 35–40°C km�1. The left
and right sides were assumed to be thermally
insulated for heat.
(A) (B)
(C) (D)
Fig. 9. Evolution of petroleum saturation for
Model 1 at (A) 40 000 years, (B) 60 000 years,
(C) 80 000 years, and (D) 100 000 years of
simulation time showing the advancement of
hydrocarbons through the fault zone. Petroleum
invades into the upper Miocene units because
of the relatively high fluid pressure in the lower
part of the fault. The outline of this magnified
area is marked with dashed lines in Fig. 6.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
242 B. JUNG et al.
effects and accumulates significantly enhanced pools in the
Pliocene sediments compared with that of low permeabil-
ity cases. In the low permeability model (Fig. 11A, Model
2), however, accumulation tends to be more abundant in
the upper Miocene sediments and less in the Pliocene
unit.
Fig. 10. Effect of fault permeability on flow
duration needed to accumulate the current
known volume of petroleum in the Inglewood
oil field (1 MMBOE = 106 barrels). Flow dura-
tion assumes that 20% of the total fault length
in transverse is available for petroleum migra-
tion and therefore characteristic or effective
flow length of about 1 km.
(A) (B)
(C) (D)
Fig. 11. Comparison of continuous flow models
showing the effects of fault permeability on
petroleum migration and distribution, indicated
by saturation. Numbers in parenthesis indicate
times needed for accumulating total oil reverse
in Inglewood field (see Fig. 10).
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
Effects of episodic fluid flow 243
EPISODIC FLOW MODEL
Models 5 through 8 were designed to investigate the
effects of episodic fluid flow on petroleum migration,
based on a fault-valve mechanism (Wright 1991), so a
repeating pattern of changing fault permeability and pore
fluid pressure was assigned (Fig. 12). A sharp increase in
permeability from k = 1 to 1000 md (10�15 to 10�12 m2)
denotes fault rupturing caused by tectonic stresses or due
to hydrofracturing by fluid pressure accumulation (Fisher
& Knipe 2001). Elevated fault permeability and accompa-
nied focused fluid flow are, however, considered to be
short-lived (<100 years) events (Sibson 1994), so the fault
will be closed hydromechanically as the pore pressure
decreases and will be assumed to sealed by mineral precipi-
tation (Dieterich & Kilgore 1996; Scholz 2002; Johnson
& Jia 2005). Recent field measurements of time-varying
permeability in faults after large earthquakes also suggest
that transient fluid pulses play a major role for water circu-
lation in deep fault zones (Xue et al. 2013).
To simulate these sealing effects, the fault zone perme-
ability was decreased from 1000 to 10 md during the first
100 years after rupturing, and then, it continued to be
reduced to 1 md until the next pulse occurs 2000 years
later. We assumed that the hydromechanical compaction
causing the first stage of permeability drop is abrupt and
much faster than the second stage of drop dominated by
hydrothermal precipitation.
We also assume that the fault ruptured at overpressure
ratio of 80%, and pore fluid pressure dropped to 10% over-
pressure level for the episodic flow model. The fluid pressure
gradually builds back up to 80% during the sealing period.
Qualitative behavior of these fluctuating pore pressure pat-
terns was first suggested by Sibson (see review by Roberts &
Nunn 1995), and the pressure value at the fault rupture was
chosen from work by Gratier et al. (2003) and Appold et al.
(2007). Petrologic evidence of multiple pulses of petroleum
flow in fault zones is reported in publications that deal with
Neogene basins in southern California (Boles & Ramseyer
1987; Eichhubl & Boles 2000; Perez & Boles 2007), but
there are a lack of articles providing a quantitative estimation
of possible flow duration. In this study, the period of epi-
sodic flow pulse was chosen within the range of 500–
2000 years based on Myers’s research on the Puente Hill
blind fault in the LA basin (Myers et al. 2003), where he
estimated the recurrence period of earthquakes in this fault
zone to be approximately 1700–3200 years, assuming a
Richter magnitude of M = 6.0–7.5. We assume that earth-
quakes in that range could impose significant hydromechan-
ical impacts on the adjacent aquifers.
Fig. 12. Fault permeability and pore fluid pres-
sure settings for the episodic flow model (Model
6). Fault permeability changes applied to the
whole fault zone, and pore pressure changes
applied to the end of the fault on the boundary
elements (marked by the white arrow in Fig. 6)
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
244 B. JUNG et al.
When the fault zone is sealed, pore fluid pressure builds
up in the lower part of the fault so it cannot reach the upper
formations due to the low fault permeability (Fig. 13A).
After the fault is open, previously accumulated high pressure
is released, and a temporary focus of fluid flow is created in
the fault zone (Fig. 13B). The released overpressure per-
vades the Miocene formations and the center of the anticline
structure. This strong and episodic pressure anomaly may
act as a strong transient driving force for both groundwater
and petroleum. Furthermore, a subsurface heat anomaly is
also detected along the fault zone due to the increased flow
rate and heat advection (Fig. 14). Temperature elevation in
the fault zone can facilitate the petroleum migration by
decreasing viscosity of the fluid. However, the heat anomaly
disappears when the fault is sealed, and isotherms become
horizontal and conductive.
The pattern of petroleum migration in the early stages is
similar to that of continuous migration, which mostly fol-
lows the fault surface (Fig. 15A), but some distinctive and
different patterns appear in later stages. In the episodic
model, petroleum migration seems to be more dispersed
than in the continuous model and accumulates in broader
areas in the middle of the shallow Pliocene sediments.
Petroleum migrates from the fault structures themselves
and extends into the adjacent sedimentary formations.
Effective spreading of petroleum to the adjacent rocks is
due to the high permeability, Darcy flow rates, and fluid
pressure concentrated in this area during the episodic flow
events. Especially, the high fluid pressure facilitates petro-
leum to overcome entry pressure of reservoir rocks sur-
rounding the fault zone.
EFFECT OF EPISODIC FLOW FREQUENCY
To further investigate the effect of frequency of episodic
fluid flow on petroleum accumulation, the temporal varia-
tion of total petroleum volumes for Models 5 to 8 can be
compared graphically (Fig. 16). The total volume of petro-
(A) (B)
Fig. 13. Hydraulic head change (Δh = h – h0)
for the episodic flow model (Model 6): (A) just
before the fault opening due to fault rupture,
and (B) 50 years after the fault opening. h0 is
the hydrostatic pressure. Numbers are in meter.
(A) (B)
Fig. 14. Subsurface temperature for the epi-
sodic flow model (Model 6): (A) before the fault
opening, and (B) 50 years after the fault open.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
Effects of episodic fluid flow 245
leum increases with flow frequency. If we set the known
450 million barrels as the comparison base, the flow model
with a shorter period of episodic flow pulses reaches the
base faster than the model with a longer period. For exam-
ple, it takes approximately 5700 years if the period of epi-
sodic pulse is 500 years (Model 8) but takes more than
24 100 years if the period is 3000 years (Model 5). Petro-
leum distributions of Models 5 to 8 at the comparison
base were compared in Fig. 17. The overall patterns of
petroleum distributions from four models look similar.
(A) (B)
(C) (D)
Fig. 15. Evolution of petroleum saturation for
the episodic flow model (Model 6) at four time
steps.
Fig. 16. Comparison of total petroleum volume
changes of episodic flow models, showing the
effect of fault permeability on recharging (fill-
ing) times needed to accumulate the observed
volume of total oil reserve in the Inglewood oil
field. Filling times (durations) assume that only
20% of the fault zone is available for petroleum
migration along the length of the oil field.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
246 B. JUNG et al.
They all caused significant accumulations in the Pliocene
unit and channeled flow in the middle Miocene unit, but
the shorter period scenario (Model 8, Fig. 17D) enables
more pervasive accumulations in the center of the Pliocene
unit, in the Sentous fault, and in the upper part of the
Newport-Inglewood fault zone. It also caused relatively
smaller intrusion into the Middle Miocene unit, which
probably means that the episodic migration having a
shorter period is more active in transferring petroleum
upward because the fault zone is open more frequently
and assumed to be sealed for less time.
The episodic flow model having the fault k range of 1–
1000 md (10�15 to 10�12 m2) and a pulse interval of
3000 years (Model 5) can be compared with a continuous
model (Model 4) having k values of 50 md to demonstrate
differences between continuous and episodic flow models.
These two cases were selected because their simulation
times required for fill the reservoir were most similar
although the pattern of the driving force is different. The
episodic model (Model 5) takes 24 100 years to accumu-
late the petroleum at the level of comparison point
(approximately 450 million barrels). This value is faster
than the elapsed time of the continuous model, which is
34 000 years, having fault permeability of 50 md. When
we consider the relatively short sum of open fault duration
(<1400 years) in Model 5, it could be said that the epi-
sodic flow can drive migration of petroleum more effi-
ciently compared to the continuous models with constant
fault permeability higher than 50 md.
For the given amount of total petroleum volume, epi-
sodic fault pressure-driven migration generates broader and
more highly saturated petroleum accumulations in the Pli-
ocene sediments and also showed more dispersed patterns
in the Middle Miocene unit. On the other hand, the con-
tinuous migration made relatively low-saturated accumula-
tions simply following the fault structures. Petroleum in
the continuous model also tends to stay in the lower part
of the fault zone and rise slowly by the forces of buoyancy.
CONCLUSIONS
To understand the dynamics of petroleum migration in
active margin basins, both continuous and episodic fluid
migration models were constructed for the Newport-Ingle-
wood fault zone in the Los Angeles basin. The effects of
time-varying fault permeability on hydrocarbon accumula-
tion rate and pattern are also quantitatively investigated
using multiphase numerical modeling.
In continuous flow models, the petroleum migration
rate in the fault zone ranges 0.03–0.2 m year�1, and the
(A) (B)
(C) (D)
Fig. 17. Comparison of episodic flow models
showing the effects of flow frequency on petro-
leum migration and distribution. Numbers in
parenthesis indicate times needed for accumu-
lating total oil reverse in Inglewood field (see
Fig. 16).
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
Effects of episodic fluid flow 247
duration of time needed to form the Inglewood oil field
(approximately 450 million barrels) is approximately
34 000–62 000 years depending on the fault permeability
assigned varying 50–5 md. High fault permeability also
allows more petroleum accumulation and channeling in the
Pliocene formation because larger overpressures can propa-
gate further and reach at the upper part of the fault.
In episodic models, the total oil volume in the reservoir
can be accumulated in about 24 000 years under assump-
tions that fault permeability fluctuates between 1–1000 md
and seismically induced flow pulse occurs every 2000 years.
Episodic migration forms larger areas of oil accumulation
with highly petroleum saturation accumulations in a
broader region surrounding the fault zone. Considering
the observed pattern of hydrocarbon accumulation in this
region, the episodic model could be more geologically
meaningful than continuous migration model.
The frequency of episodic flow affects petroleum migra-
tion and distribution patterns in the fault zone. Faster and
broader accumulations appear in the Pliocene sediments if
the pulse period becomes shorter. The frequent episodic
fluid pulse also forms accumulations with higher petroleum
saturation and flow channels in adjacent formations.
ACKNOWLEDGEMENTS
This research was supported by a grant from the U.S.
Department of Energy – Basic Energy Sciences (Grant
number: DE-FG02-96ER14620), The GDL Foundation
also provided a Ph.D. scholarship to the first author, for
which he is extremely grateful.
REFERENCES
Appold MS, Garven G (2000) Reactive flow models of ore
formation in the Southeast Missouri district. Economic
Geology,
95, 1605–26.
Appold MS, Garven G, Boles JR, Eichhubl P (2007) Numerical
modeling of the origin of calcite mineralization in the Refugio-
Carneros fault, Santa Barbara Basin, California. Geofluids, 7,
79–
95.
Aydin A (2000) Fractures, faults, and hydrocarbon entrapment,
migration and flow. Marine and Petroleum Geology, 17, 797–
814.
Bear J (1972) Dynamics of Fluids in Porous Media. Dover
Publications Inc., New York.
de Berg M, Cheong O, van Kreveld M, Overmars M (2008)
Computational Geometry, 3rd edn. Springer-Verlag, Berlin, 386
p.
Berry FAF (1973) High fluid potentials in California Coast
Ranges and their tectonic significance. American Association of
Petroleum Geologists Bulletin, 57, 1219–45.
Biddle KT (1991) The Los Angeles basin: an overview. In:
Active
Margin Basins (ed. Biddle KT), pp. 1–24. AAPG Memoir 52,
Tulsa, OK.
Blackwell DD, Steele JL (1989) Thermal conductivity of
sedimentary rocks: measurement and significance. In: Thermal
History of Sedimentary Basins: Methods and Case Histories
(eds
Naeser ND, McCulloh TH), pp. 13–36. Springer-Verlag, New
York.
Blake GH (1991) Review of the Neogene biostratigraphy and
stratigraphy of the Los Angeles Basin and implications for basin
evolution. In: The Los Angeles Basin: An Overview (ed. Biddle
KT), pp. 135–84. AAPG Memoir 52, Tulsa, OK.
Blanpied ML, Lockner DA, Byerlee JD (1992) An earthquake
mechanism based on rapid sealing of faults. Nature, 358,
574–6.
Bloomfield JP, Gooddy DC, Bright MI, Williams PJ (2001)
Pore-
throat size distributions in Permo-Triassic sandstones from the
United Kingdom and some implications for contaminant
hydrogeology. Hydrogeology Journal, 9, 219–30.
Boles JR, Ramseyer K (1987) Diagenetic carbonate in Miocene
sandstone reservoir, San Joaquin basin, California. American
Association of Petroleum Geologists Bulletin, 71, 1475–87.
Boles JR, Eichhubl P, Garven G, Chen J (2004) Evolution of a
hydrocarbon migration pathway along basin-bounding faults:
evidence from fault cement. American Association of Petroleum
Geologists Bulletin, 88, 947–70.
Bradley JS (1975) Abnormal formation pressure. American
Association of Petroleum Geologists Bulletin, 59, 957–73.
Caine JS, Bruhn RL, Forster CB (2010) Internal structure, fault
rocks, and inferences regarding deformation, fluid flow, and
mineralization in the seismogenic Stillwater normal fault, Dixie
Valley, Nevada. Journal of Structural Geology, 32, 1576–89.
California Department of Conservation (1992) California Oil
and
Gas Fields: Volume II – Southern, Central Coastal, and
Offshore
California Oil and Gas Fields. California Department of
Conservation, Division of Oil, Gas, and Geothermal Resources,
Sacramento, CA.
Class H, Helmig R, Bastian P (2002) Numerical simulation of
non-isothermal multiphase multicomponent processes in porous
media. 1. An efficient solution technique. Advances in Water
Resources, 25, 533–50.
Dieterich JH, Kilgore BD (1996) Imaging surface contacts:
power
law contact distributions and contact stresses in quartz, calcite,
glass and acrylic plastic. Tectonophysics, 256, 219–39.
Eichhubl P, Boles JR (2000) Focused fluid flow along faults in
the
Monterey Formation, coastal California. Geological Society of
America Bulletin, 112, 1667–79.
England WA, Mackenzie AS, Mann DM, Quigley TM (1987)
The
movement and entrapment of petroleum fluids in the
subsurface. Journal of the Geological Society, London, 144,
327–
47.
Evans B (1992) Greasing the fault. Nature, 358, 544–5.
Faulkner DR, Jackson CAL, Lunn RJ, Schlische RW, Shipton
ZK,
Wibberley CAJ, Withjack MO (2010) A review of recent
developments concerning the structure, mechanics and fluid
flow properties of fault zones. Journal of Structural Geology,
32,
1557–75.
Fisher QJ, Knipe RJ (2001) The permeability of faults within
siliciclastic petroleum reservoirs of the North Sea and
Norwegian Continental Shelf. Marine and Petroleum Geology,
18, 1063–81.
Fisher QJ, Casey M, Harris SD, Knipe RJ (2003) Fluid-flow
properties of faults in sandstone: the importance of temperature
history. Geology, 31, 965–8.
Garden IR, Guscott SC, Burley SD, Foxford KA, Walsh JJ,
Marshall J (2001) An exhumed palaeo-hydrocarbon migration
fairway in a faulted carrier system, Entrada Sandstone of SE
Utah, USA. Geofluids, 1, 195–213.
Geiger S, Burri A, Roberts S, Matthai SK, Zoppou C (2004)
Combining finite element and finite volume methods for
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
248 B. JUNG et al.
efficient multiphase flow simulations in highly heterogeneous
and structurally complex geologic media. Geofluids, 4, 284–99.
Geiger S, Driesner T, Heinrich CA, Matthai SK (2006)
Multiphase thermohaline convection in the earth’s crust: I. A
new finite element – finite volume solution technique combined
with a new equation of state for NaCl-H2O. Transport in
Porous Media, 63, 399–434.
Gelhar LW, Welty C, Rehfeldt KR (1992) A critical review of
data
on field-scale dispersion in aquifers. Water Resources Research,
28, 1955–74.
van Genuchten MT (1980) A closed-form equation for
predicting
the hydraulic conductivity of unsaturated soils. Soil Science
Society of America Journal, 44, 892–8.
Glasø Ø (1980) Generalized pressure-volume-temperature
correlations. Journal of Petroleum Technology, 32, 785–95.
Gong ZS, Huang LF, Chen PH (2011) Neotectonic controls on
petroleum accumulations, offshore China. Journal of Petroleum
Geology, 34, 5–28.
Gratier JP, Favreau P, Renard F (2002) Fluid pressure evolution
during the earthquake cycle controlled by fluid flow and
pressure solution crack sealing. Earth, Planets and Space, 54,
1139–46.
Gratier JP, Favreau P, Renard F (2003) Modeling fluid transfer
along California faults when integrating pressure solution crack
sealing and compaction processes. Journal of Geophysical
Research, 108, 1–25.
Guo X, He S, Liu K, Cao F, Shi H, Zhu J (2011) Condensates in
the PY30-1 structure, Panyu Uplift, Pearl River Mouth Basin,
South China Sea: evidence for hydrothermal activity associated
with petroleum migration and accumulation. Journal of
Petroleum Geology, 34, 217–32.
Harten A (1997) High resolution schemes for hyperbolic
conservation laws. Journal of Computational Physics, 135, 260–
78.
Hayba DO, Bethke CM (1995) Timing and velocity of petroleum
migration in the Los Angeles Basin. Journal of Geology, 103,
33–49.
Helmig R (1997) Multiphase Flow and Transport Processes in
the
Subsurface: A Contribution to the Modeling of Hydrosystems.
Springer, Berlin. 367 pp.
Hesson HB, Olilang HR (1990) Seal Beach Oil Field: Alamitos
and Marine Areas TR39. Division of Oil and Gas, California
Department of Conservation, Sacramento, CA.
Higgins CT, Chapman RH (1984) Geothermal energy at Long
Beach Naval Shipyard and Naval Station and at Seal Beach
Naval Weapons Station, California. DMG Open-File Report
84-32, California Department of Conservation, Division of
Mines and Geology, Sacramento, CA.
Huber R, Helmig R (2000) Node-centered finite volume
discretizations for the numerical simulation of multiphase flow
in heterogeneous porous media. Computational Geosciences, 4,
141–64.
Ingebritsen SE, Neuzil CE, Sanford WE (2006) Groundwater in
Geologic Processes, 2nd edn. Cambridge University Press,
Cambridge. 536 pp.
Jeffrey AWA, Alimi HM, Jenden PD (1991) Geochemistry of
Los
Angeles basin and gas system. In: Active Margin Basins (ed.
Biddle KT), pp. 197–220. AAPG Memoir 52, Tulsa, OK.
Johnson PA, Jia X (2005) Nonlinear dynamics, granular media
and dynamic earthquake triggering. Nature, 437, 871–4.
Kaplan IR, Alimi MH, Hein C, Jeffrey A, Lafferty MR,
Mankiewicz MP (2000) The geochemistry of hydrocarbons and
potential source rocks from the Los Angeles and Ventura basins,
data synthesis and text, v.I. In: Collection of Papers Written in
the Mid-to-Late 1980’s and in 1997 by Staff Members of Global
Geochemistry Corporation about the Oil, Gas, and Source Rock
Investigations Carried Out in the San Joaquin, Santa Maria,
Santa Barbara, Ventura, and Los Angeles Basins, California
(ed.Kaplan IR), pp. 1–238. Pacific Section – AAPG,
Bakersfield,
CA.
Karlsen DA, Skeie JE (2006) Petroleum migration, faults and
overpressure, Part I: calibrating basin modelling using
petroleum in traps – a review. Journal of Petroleum Geology,
29,
227–55.
Kroeger KF, di Primio R, Horsfield B (2009) Hydrocarbon flow
modeling in complex structures (Mackenzie Basin, Canada).
American Association of Petroleum Geologists Bulletin, 93,
1209–34.
Levorsen AI (1967) Geology of Petroleum, 2nd edn. W. H.
Freeman, San Francisco, CA.
Mandl G, Harkness RM (1987) Hydrocarbon migration by
hydraulic fracturing. In: Deformation of Sediments and
Sedimentary Rocks 29 (eds Jones ME, Preston RMF), pp. 39–
53. Geological Society of London Special Publication, London.
de Marsily G (1986) Quantitative Hydrogeology: Groundwater
Hydrology for Engineers. Academic Press, Orlando, FL, 440 p.
McCulloh TH (1979) Implication for petroleum appraisal. In:
Geologic Studies of the Point Conception Deep Stratigraphic
Test
Well OCS-CAL 78-164 No. 1 Outer Continental Shelf Southern
California, United States (ed. Cook HE), pp. 26–42. USGS
Open File Report 79-1218, U.S. Geological Survey, Menlo
Park, CA.
Myers DJ, Nabelek JL, Yeats RS (2003) Dislocation modeling
of
blind thrusts in the eastern Los Angeles basin, California.
Journal of Geophysical Research-Solid Earth, 108, ESE 14, 1–
19.
Nishikawa T, Siade AJ, Reichard EG, Ponti DJ, Canales AG,
Johnson TA (2009) Stratigraphic controls on seawater intrusion
and implications for groundwater management, Dominguez
Gap area of Los Angeles, California, USA. Hydrogeology
Journal,
17, 1699–725.
Olson L (1978) Shallow Aquifers and Surface Casing
Requirements for Wilmington and Belmont Offshore Oil Fields
TR22. California Department of Conservation, Division of
Mines and Geology, Sacramento, CA.
Perez RJ, Boles JR (2007) Mineralization, fluid flow, and
sealing
properties associated with an active thrust fault: San Joaquin
basin, California. American Association of Petroleum
Geologists
Bulletin, 88, 1295–314.
Phillips SL, Ingbene A, Fair JA, Ozbek H, Tavana M (1981) A
technical data book for geothermal energy utilization. DE81-
029868.
Phillips SL, Ozbek H, Silvester LF (1983) Density of sodium
chloride solutions at high temperatures and pressures. DE84-
004883.
Reichenberger V, Jakobs H, Bastian P, Helmig R (2006) A
mixed-dimensional finite volume method for two-phase flow in
fractured porous media. Advances in Water Resources, 29,
1020–36.
Reynolds SJ, Lister GS (1987) Structural aspects of fluid-rock
interactions in detachment zones. Geology, 15, 362–6.
Roberts SJ, Nunn JA (1995) Episodic fluid expulsion from
geopressured sediments. Marine and Petroleum Geology, 12,
195–204.
Sabins EF (1997) Remote Sensing: Principles and
Interpretation.
Wiley & Sons, New York. 494 pp.
Scholz CH (2002) The Mechanics of Earthquakes and Faulting.
2nd edn. Cambridge University Press, Cambridge. pp. 471.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
Effects of episodic fluid flow 249
Shaw JH, Suppe J (1996) Earthquake hazards of active blind-
thrust faults under the central Los Angeles basin, California.
Journal of Geophysical Research, 101, 8623–42.
Sibson RH (1994) Crustal stress, faulting and fluid flow. In:
Geofluids: Origin, Migration, and Evolution of Fluids in
Sedimentary Basins (ed. Parnell J) 78, pp. 69–84. Geological
Society Special Publication, London.
Sorkhabi R, Tsuji Y (2005) The place of faults in petroleum
traps.
In: Faults, Fluid Flow, and Petroleum Traps (eds Sorkahabi R,
Tsuji Y), AAPG Memoir 52, Tulsa, OK.
Sperrevik S, Gillespie PA, Fisher QJ, Halvorsen T, Knipe RJ
(2002)
Empirical estimation of fault rock properties. In: Hydrocarbon
Seal Quantificaiton (eds Koestler AG, Hunsdale R) 11, pp. 109–
25. Norwegian Petroleum Society Special Publication,
Stavanger,
Norway.
Walder J, Nur A (1984) Porosity reduction and crystal pore
pressure development. Journal of Geophysical Research, 89,
11539–48.
Wang K, He J, Hu Y (2006) A note on pore fluid pressure ratios
in the Coulomb wedge theory. Geophysical Research Letters,
33,
L19310. doi:10.1029/2006GL027233.
Watson JT, Basu RS, Sangers JV (1980) An improved
representative equation for the dynamic viscosity of water
substance. Journal of Physical and Chemical Reference Data, 9,
1255–90.
Wendebourg J (1994) Simulating hydrocarbon migration and
stratigraphic traps, PhD thesis. Standford University.
Wright TL (1991) Structural geology and tectonic evolution of
the Los Angeles Basin, California. In: Active Margin Basins
(ed.
Biddle KT), pp. 35–134. AAPG Memoir 52, Tulsa, OK.
Xue L, Li HB, Brodsky EE, Xu ZQ, Kano Y, Wang H, Mori
JJ, Si JL, Pei JL, Zhang W, Yang G, Sun ZM, Huang Y
(2013) Continuous permeability measurements record healing
inside the Wenchuan earthquake fault zone. Science, 304,
1555–9.
Yamaguchi A, Cox SF, Kimura G, Okamoto S (2011) Dynamic
changes in fluid redox state associated with episodic fault
rupture along a megasplay fault in a subduction zone. Earth
and Planetary Science Letters, 302, 369–77.
Yerkes RF (1972) Geology and oil resources of the western
Puente Hills area, southern California. USGS Professional
Paper,
420-C, 63.
Zachos JC, Pagani M, Lisa S, Thomas E, Billups K (2001)
Trends, rhythms, and aberrations in global climate 65 Ma to
present. Science, 292, 686–93.
Zhang Y, Gable CW, Zyvoloski GA, Walter LM (2009)
Hydrogeochemistry and gas compositions of the Uinta Basin: a
regional-scale overview. American Association of Petroleum
Geologists Bulletin, 93, 1087–118.
© 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250
250 B. JUNG et al.
Volume 14, Number 2, May 2014
ISSN 1468-8115
Geofluids
This journal is available online at Wiley Online Library.
Visit onlinelibrary.wiley.com to search the articles and register
for table of contents and e-mail alerts.
Geofluids is abstracted/indexed in Chemical Abstracts
CONTENTS
127 Review of ‘Too Hot To Touch: The Problem of High-Level
Nuclear Waste’ by
William M. Alley and Rosemarie Alley
E.M. Kwicklis
128 Diffusion and kinetic control of weathering layer
development
D. Reeves and D.H. Rothman
143 Carbon dioxide controlled earthquake distribution pattern in
the NW Bohemian
swarm earthquake region, western Eger Rift, Czech Republic –
gas migration in the
crystalline basement
F.H. Weinlich
160 Fractal analysis of veins in Permian carbonate rocks in the
Lingtanchang anticline,
western China
B. Deng, S. Liu, L. Jansa, S. Yong and Z. Zhang
174 Fluid effect on hydraulic fracture propagation behavior: a
comparison between water
and supercritical CO2-like fluid
X. Zhou and T.J. Burbey
189 Cementation and the hydromechanical behavior of
siliciclastic aquifers and reservoirs
D.F. Boutt, K.E. Plourde, J. Cook and L.B. Goodwin
200 Impacts of Pleistocene glacial loading on abnormal pore-
water pressure in the
eastern Michigan Basin
O. Khader and K. Novakowski
221 Numerical simulation of mylonitization and structural
controls on fluid flow and
mineralization of the Hetai gold deposit, west Guangdong,
China
J. Zhu, Z. Li, G. Lin, Q. Zeng, Y. Zhou, J. Yi, G. Gong and G.
Chen
234 Effects of episodic fluid flow on hydrocarbon migration in
the Newport-Inglewood
Fault Zone, Southern California
B. Jung, G. Garven and J.R. Boles
gfl_14_2_Issue toc_OC 4/11/2014 12:51 PM Page 1
Copyright of Geofluids is the property of Wiley-Blackwell and
its content may not be copied
or emailed to multiple sites or posted to a listserv without the
copyright holder's express
written permission. However, users may print, download, or
email articles for individual use.
747
Bulletin of the Seismological Society of America, Vol. 94, No.
2, pp. 747–752, April 2004
Activity of the Offshore Newport–Inglewood Rose Canyon Fault
Zone,
Coastal Southern California, from Relocated Microseismicity
by Lisa B. Grant and Peter M. Shearer
Abstract An offshore zone of faulting approximately 10 km
from the southern
California coast connects the seismically active strike-slip
Newport–Inglewood fault
zone in the Los Angeles metropolitan region with the active
Rose Canyon fault zone
in the San Diego area. Relatively little seismicity has been
recorded along the off-
shore Newport–Inglewood Rose Canyon fault zone, although it
has long been sus-
pected of being seismogenic. Active low-angle thrust faults and
Quaternary folds
have been imaged by seismic reflection profiling along the
offshore fault zone, raising
the question of whether a through-going, active strike-slip fault
zone exists. We
applied a waveform cross-correlation algorithm to identify
clusters of microseis-
micity consisting of similar events. Analysis of two clusters
along the offshore fault
zone shows that they are associated with nearly vertical, north-
northwest-striking
faults, consistent with an offshore extension of the Newport–
Inglewood and Rose
Canyon strike-slip fault zones. P-wave polarities from a 1981
event cluster are con-
sistent with a right-lateral strike-slip focal mechanism solution.
Introduction
The Newport–Inglewood fault zone (NIFZ) was first
identified as a significant threat to southern California resi-
dents in 1933 when it generated the M 6.3 Long Beach earth-
quake, killing 115 people and providing motivation for pas-
sage of the first seismic safety legislation in the United States
(Barrows, 1974; Hauksson and Gross, 1991; Yeats, 2001).
The Rose Canyon fault (RCF) zone in the San Diego area
has ruptured several times during the Holocene (Lindvall
and Rockwell, 1995) and poses significant seismic hazard to
San Diego area residents (Anderson et al., 1989). An off-
shore zone of faulting approximately 10 km from the south-
ern California coast connects the strike-slip NIFZ in the Los
Angeles metropolitan region with the strike-slip RCF zone
in the San Diego area. The activity and seismic potential of
the intervening offshore fault zone, herein referred to as the
offshore Newport–Inglewood Rose Canyon (ONI-RC) fault
zone, has been the subject of debate for decades (e.g., Hill,
1971; Barrows, 1974; Fischer and Mills, 1991; Rivero et al.,
2000; Grant and Rockwell, 2002). Recent attention has fo-
cused on blind thrusts that may intersect the ONI-RC fault
zone and accommodate some of the regional deformation
(Grant et al., 1999; Rivero et al., 2000). Interaction with the
thrust system could limit the magnitude of earthquakes on
strike-slip faults in the ONI-RC fault zone, if they are active.
Sparse offshore microseismicity has been difficult to locate
accurately (Astiz and Shearer, 2000) but has been interpreted
to be associated with an active ONI-RC fault zone (Fischer
and Mills, 1991). We identified, relocated, and analyzed two
clusters of microearthquakes within the northern and central
ONI-RC fault zone to examine the fault structure, minimum
depth of seismic activity, and source fault mechanism. The
results suggest that the ONI-RC fault zone is a steeply dip-
ping, active strike-slip fault to seismogenic depths.
Tectonic and Geologic Setting
In southern California, tectonic deformation between
the Pacific and North American plates is accommodated pri-
marily by a zone of strike-slip faults (Walls et al., 1998)
with maximum deformation along the San Andreas system,
decreasing westward into the offshore inner Continental
Borderland region (Fig. 1). The inner Continental Border-
land has complex structure including low-angle detachment
faults, thrust faults, and high-angle strike-slip faults resulting
from late Cenozoic rifting prior to the current transpressional
tectonic regime (Crouch and Suppe, 1993; Bohannon and
Geist, 1998). The geometry and slip rate of faults in the inner
Continental Borderland are poorly constrained relative to
onshore faults (Astiz and Shearer, 2000), yet they may pose
significant seismic risk because they are close to populated
areas, and several offshore faults appear to displace seafloor
sediments (Legg, 1991). Grant and Rockwell (2002) pro-
posed that an active �300-km-long Coastal Fault zone
(Fig. 2) extends between the Los Angeles basin (USA) and
coastal Baja California (Mexico), including the NIFZ and
RCF zone in southern California, the Agua Blanca fault in
748 Short Notes
Figure 1. Regional fault map (modified from Grant and
Rockwell [2002]; after
Walls et al. [1998]) summarizing regional tectonic deformation
according to slip rate
on major faults, including the San Andreas fault (SAF), San
Jacinto fault zone (SJFZ),
Elsinore fault zone (EFZ), Whittier fault (WF), Palos Verdes
fault (PVF), NIFZ near Los
Angeles (LA), RCF zone (RCFZ) near San Diego (SD), Agua
Blanca fault zone (ABFZ),
San Miguel fault zone (SMFZ), Imperial fault (IF), Cerro Prieto
fault (CPF), and Laguna
Salada fault (LSF). Offshore faults of the inner Continental
Borderland include the
Coronado Bank fault zone (CBF), San Diego Trough fault
(SDTF), San Clemente fault
zone (SCFZ), Santa Cruz Island fault (SCIF), and Santa Rosa
Island fault (SRIF). The
San Gabriel fault (SGF), San Cayetano fault (SCF), Oak Ridge
fault (ORF), and Santa
Ynez fault (SYF) are located in the Transverse Ranges.
Baja California, and contiguous offshore fault zones. The
structural style and slip rates of the NIFZ and RCF zone sug-
gest that the intervening ONI-RC fault zone is an active,
strike-slip fault zone with complex structure and a slip rate
of 0.5–2.0 mm/yr.
The NIFZ has been studied extensively in the Los An-
geles basin by petroleum geologists. Hill (1971) proposed
that the NIFZ overlies a major tectonic boundary separating
eastern continental basement facies of granitic and associ-
ated metamorphic rocks from oceanic Catalina schist facies
of the inner Continental Borderland province to the west.
The current strike-slip faulting regime apparently reactivated
a structural weakness along the Mesozoic subduction zone
(Hill, 1971) and initiated right-lateral motion in the mid-to-
late Pliocene (Wright, 1991; Freeman et al., 1992). The NIFZ
is a structurally complex series of discontinuous strike-slip
faults with associated folds and shorter normal and reverse
faults (Yeats, 1973), described by Wilcox et al. (1973) and
Harding (1973) in a set of classic papers on wrench tecton-
ics. Dextral displacement of late Miocene and younger sed-
iments reveals a long-term slip rate of 0.5 mm/yr (Freeman
et al., 1992). Multiple surface ruptures since the early Ho-
locene indicate that the minimum slip rate is at least 0.3–0.6
mm/yr and may be substantially higher (Grant et al., 1997).
A Holocene slip rate of �1 mm/yr has been assumed for
seismic hazard assessment (SCECWG, 1995). Seismically
active strands of the NIFZ have been mapped along the west-
ern margin of the Los Angeles basin between the cities of
Beverly Hills and Newport Beach (Bryant, 1988; Grant et
al., 1997) where the fault steps over and continues offshore
parallel to the coast (Morton and Miller, 1981).
The subsurface structure and tectonic history of the RCF
zone have not been studied as thoroughly as the NIFZ be-
cause it lacks petroleum reserves. The Holocene sense of
movement and slip rate of the RCF have been investigated
for seismic hazard studies. Lindvall and Rockwell (1995)
Short Notes 749
Figure 2. Map shows approximate location of major strands in
the Coastal fault
zone and dates of most recent rupture. The Coastal fault zone
includes the northern
NIFZ near Beverly Hills (BH) (dashed), southern NIFZ (bold)
near Newport Beach (NB),
the ONI-RC fault zone (gray) offshore of the San Joaquin Hills
(SJH) and city of Ocean-
side (O), the RCF zone (bold line) between La Jolla (LJ) and
San Diego, the Coronado
Bank fault zone (CBF), and the Agua Blanca fault zone (ABFZ,
bold). Modified from
Grant and Rockwell (2002); after Grant et al. (1997), Lindvall
and Rockwell (1995),
and Rockwell and Murbach (1999).
and Rockwell and Murbach (1999) reported that it is pri-
marily a strike-slip fault zone with a slip rate of 1–2 mm/yr,
although individual strands display varying amounts of dip-
slip motion in combination with strike slip.
Location and Structural Models of the ONI-RC
Fault Zone
The location of the ONI-RC fault zone has been mapped
by seismic reflection profiling in shallow water along the
inner continental margin between Newport Beach and La
Jolla, north of San Diego. Fischer and Mills (1991) sum-
marized the results of eight investigations between 1972 and
1988, including unpublished mapping for the nuclear gen-
erating station at San Onofre. Interpretations of Fischer and
Mills (1991) and prior studies disagree about the location
and number of traces, but all reveal a fairly continuous zone
of faulting within a few kilometers of the coastline. Fischer
and Mills (1991) presented several interpretive cross sec-
tions with a steeply dipping fault and flower structure sug-
gestive of strike-slip motion. They described an inner and
outer “thrust fault-fold complex” consisting of multiple
thrust faults and thrust-generated folds associated with a
positive flower structure, and they described an offshore ex-
tension of the San Joaquin Hills as being dextrally offset by
the ONI-RC fault zone. Grant and others (1999, 2000, 2002)
reported late Quaternary and Holocene uplift of the San Joa-
quin Hills inboard of the ONI-RC fault zone and suggested
that seismic activity of the San Joaquin Hills blind thrust and
ONI-RC fault zone are linked.
Recent offshore investigations have focused on the
structure and potential activity of a low-angle fault, the
Oceanside thrust, and its relationship to the ONI-RC fault
zone (Bohannon and Geist, 1998; Rivero et al., 2000). Riv-
ero et al. (2000) proposed that the San Joaquin Hills blind
thrust is a backthrust soling into the larger Oceanside thrust.
In this model, Quaternary uplift of the San Joaquin Hills and
coastline south to San Diego is associated with movement
of the Oceanside thrust. They presented four potential con-
figurations for thrust interaction with strike-slip faults in the
ONI-RC fault zone, each leading to a different maximum
magnitude estimate, based on the activity and depth of the
strike-slip system. Our analysis of microseismicity helps
constrain the most viable model for interaction between the
Oceanside thrust and ONI-RC fault zone.
Seismicity and Event Location Method
No significant historic earthquakes have occurred on the
ONI-RC fault zone. Scattered seismicity has been recorded
along the ONI-RC fault zone (Fig. 3), although events are
difficult to locate accurately due to poor station coverage
(Fischer and Mills, 1991; Astiz and Shearer, 2000). With the
exception of the energetic 1986 ML 5.3 Oceanside earth-
quake sequence, rates of microseismicity have been low
since digital waveform data became available in 1981 (Astiz
and Shearer, 2000). Fischer and Mills (1991) reported that
the rate of seismicity along the ONI-RC fault zone is ap-
proximately 10 times lower than along the onshore NIFZ.
Rivero et al. (2000) interpreted the 1986 Oceanside earth-
quake sequence as generated by the Thirtymile Bank thrust
fault, a reactivated low-angle detachment fault.
Figure 3 shows microseismicity of the inner Continental
Borderland between 1981 and February 2003. The earth-
quake locations were computed from the archived P and S
picks using the source-specific station term method of
Richards-Dinger and Shearer (2000). This method improves
the relative location of events within compact clusters by
solving for custom station terms for subsets of the events,
thus removing the biasing effects of unmodeled three-
750 Short Notes
-118.0 -117.5
33.2
33.4
33.6
33.8
Oceanside
Newport Beach
A
B
Elsinore
San Clemente
Newport-Inglewood
P
alos Verdes
0 10 km
Figure 3. Earthquakes (black dots) near the
coastal cities (circles) of Newport Beach, San Cle-
mente, and Oceanside, California, recorded from
1981 to February 2003. Black boxes indicate clusters
of similar events near the ONI-RC fault zone. A 1981
cluster near Oceanside (box A) is shown in Figure 4,
and a 2000 cluster near Newport Beach (box B) is
shown in Figure 5. The approximate locations of the
Elsinore, Newport–Inglewood, and Palos Verdes
faults are also shown.
-117.52 -117.51
33.26
33.27
A
A'
0.0 0.4 0.8
12.0
12.5
13.0
13.5
Distance (km)
D
e
p
th
(
km
)
A A'
0.5 km0
Figure 4. The 1981 Oceanside cluster (from box
A on Fig. 3) is plotted at higher resolution in
map view and cross section, after relocation using
differential times computed using waveform cross-
correlation.
dimensional velocity and the varying station coverage
among the events. In general, this results in sharper and
better-defined seismicity alignments. The method does not,
however, necessarily yield improvements in the absolute lo-
cations of the clusters. We obtained waveforms for these
events and performed waveform cross-correlation between
each event and 100 neighboring events (Shearer, 1997; Astiz
et al., 2000). We then identified clusters of similar events
and relocated the microearthquakes within similar event
clusters using the method of Shearer (1998, 2002). On Fig-
ure 3, each cluster is so compact that these appear mostly as
single dots. Two of these clusters, in boxes labeled A and B
in Figure 3, are associated with the ONI-RC fault zone and
are the focus of this article.
Results
The Oceanside Cluster
The most interesting cluster is from a 1981 swarm of
19 M �3.0 earthquakes approximately 10 km northwest of
Oceanside. This cluster should not be confused with the
more energetic 1986 ML 5.3 Oceanside earthquake sequence
located much further offshore. As shown in Figure 4, the
events align along a north-northwest trend about 0.5 km
long. In cross section, the events define a nearly vertical
plane between 12.5 and 13.0 km depth. The strike, dip, and
location of a plane fit by these events are consistent with
active strike-slip faulting on the ONI-RC fault zone.
Fischer and Mills (1991) reported that the 1981 cluster
represents “a direct correlation of seismicity” (p. 30) with
the ONI-RC fault zone because they located the earthquakes
within 0.5 km of where they mapped the Oceanside segment
of the fault zone. Our relocation results support this inter-
pretation by showing that the events within the cluster form
a linear trend that is approximately parallel to the fault zone.
Fischer and Mills (1991) also calculated a strike-slip single
event first motion focal mechanism solution for an earth-
quake in the 1981 cluster. We are doubtful that unique focal
mechanisms can be computed for the 1981 events, given the
sparse distribution of seismic stations. However, we did ex-
amine composite waveform polarities from the 1981 cluster
using the method of Shearer et al. (2003) and found that the
polarities are consistent with a right-lateral strike-slip focal
mechanism solution, aligned with the trend of the ONI-RC
Short Notes 751
-117.870 -117.860
33.545
33.550
33.555
B
B'
0.0 0.5 1.0 1.5
6
7
Distance (km)
D
e
p
th
(
km
)
B B'
0.5 km0
Figure 5. The 2000 Newport Beach cluster (from
box B on Fig. 3) is plotted at higher resolution in
map view and cross section, after relocation using
differential times computed using waveform cross-
correlation.
fault zone. We cannot, however, entirely eliminate other
possible focal mechanisms.
The Newport Beach Cluster
A cluster of seven similar microearthquakes occurred
offshore of Newport Beach in 2000 at a depth of approxi-
mately 6.5–7.0 km. The cluster is near the offshore NIFZ as
mapped by Morton and Miller (1981) and compiled by
Fischer and Mills (1991). In map view and cross section
(Fig. 5), five of the seven events are aligned in a pattern
consistent with a shallow (7 km), north-northwest-striking,
vertical or steeply dipping active fault. Observed polarities
for this cluster are too sparse to provide any meaningful
constraint on the focal mechanisms.
The location of the 2000 cluster is southwest of an area
of active faulting reported by Fischer and Mills (1991) from
mapping offshore of Newport Beach and the San Joaquin
Hills. Fischer and Mills (1991) analyzed microseismicity
from 1982 to 1990 along this zone and reported strike-slip
and reverse-fault first motion solutions from a cluster of epi-
centers between 2 and 2.5 km on either side of the mapped
fault zone.
Discussion
The mean locations of each cluster are determined using
a standard 1D velocity model for all of southern California
and are probably not very accurate, particularly in depth. We
estimate uncertainties on the absolute cluster locations as
roughly �2 km in horizontal position and �5 km in depth.
However, the relative location accuracy among events
within each cluster is much better constrained, so the trends
defined by the seismicity alignments should be reliable. The
estimated relative location accuracy is generally less than
50 m for the events in the Oceanside cluster and less than
150 m for the aligned events in the Newport Beach cluster.
The two outlying events in the Newport Beach cluster have
standard errors of about 500 m; thus it is entirely possible
that their true locations lie within the linear trend of the other
events.
The location, alignment, and apparent dip of the Ocean-
side and Newport Beach clusters of microearthquakes are
consistent with the presence of active, steeply dipping faults
in the ONI-RC fault zone. Waveform polarities are also con-
sistent with strike-slip motion, although other mechanisms
cannot be ruled out.
For hazard estimation of offshore faults, it is not as im-
portant to precisely locate active traces as it is for onshore
faults in populated areas. A more important consideration is
the potential for a through-going rupture and an estimate of
the maximum magnitude earthquake. Our relocated micro-
seismicity is too sparse to reveal whether or not there is a
through-going strike-slip fault zone. The structure of the
ONI-RC fault zone may be similar to that of the onshore
NIFZ, which contains multiple strike-slip traces. Others have
mapped a fairly continuous structurally complex zone of
faulting �110 km long, subparallel to and within 10 km of
the coast (Fischer and Mills, 1991). The maximum magni-
tude of an ONI-RC fault rupture can be estimated from the
length of the fault zone. Assuming a 110-km surface rupture
length of a strike-slip fault zone yields an estimated M 7.4
earthquake (Wells and Coppersmith, 1994).
If strike-slip faults do not terminate the Oceanside
thrust, Rivero et al. (2000) estimate an Mw 7.5 maximum
magnitude earthquake could result from rupture of the entire
thrust fault. The �6.5-km depth of the Newport Beach seis-
micity cluster does not provide information on the geometry
or interaction between the strike-slip ONI-RC fault zone and
the Oceanside thrust. However, the location and �13 km
depth of the Oceanside cluster suggests that the Oceanside
thrust is terminated by active strike-slip faults. According to
Rivero et al. (2000), this geometry would lead to an Mw 7.3
maximum magnitude earthquake on the Oceanside thrust.
The maximum magnitude estimate is at the upper range of
magnitude estimated for an earthquake that uplifted the San
Joaquin Hills circa A.D. 1635–1769 (Grant et al., 2002).
752 Short Notes
Acknowledgments
This research was supported by the Southern California
Earthquake
Center. SCEC is funded by NSF Cooperative Agreement EAR-
0106924
and USGS Cooperative Agreement 02HQAG0008. The SCEC
Contribution
Number for this article is 746.
References
Anderson, J. G., T. K. Rockwell, and D. C. Agnew (1989). Past
and possible
future earthquakes of significance to the San Diego region,
Earth-
quake Spectra 5, 299–335.
Astiz, L., and P. M. Shearer (2000). Earthquake locations in the
inner Con-
tinental Borderland, offshore southern California, Bull. Seism.
Soc.
Am. 90, 425–449.
Astiz, L., P. M. Shearer, and D. C. Agnew (2000). Precise
relocations and
stress-change calculations for the Upland earthquake sequence
in
southern California, J. Geophys. Res. 105, 2937–2953.
Barrows, A. G. (1974). A review of the geology and earthquake
history of
the Newport–Inglewood structural zone, Calif. Div. Mines Geol.
Spe-
cial Report 114, 115 pp.
Bohannon, R. G., and E. Geist (1998). Upper crustal structure
and Neogene
tectonic development of the California continental borderland,
Geol.
Soc. Am. Bull. 110, no. 6, 779–800.
Bryant, W. A. (1988). Recently active traces of the Newport–
Inglewood
fault zone, Los Angeles and Orange Counties, California,
California
Division of Mines and Geology, DMG OFR 88-14.
Crouch, J. K., and J. Suppe (1993). Late Cenozoic tectonic
evolution of the
Los Angeles basin and inner California borderland: a model for
core
complex-like crustal extension, Geol. Soc. Am. Bull. 105,
1415–1434.
Fischer, P. J., and Mills, G. I. (1991). The offshore Newport–
Inglewood–
Rose Canyon fault zone, California: structure, segmentation and
tec-
tonics, in Environmental Perils San Diego Region, P. L. Abbott
and
W. J. Elliot (Editors), San Diego Association of Geologists, San
Di-
ego, Geological Society of America Annual Meeting, 17–36.
Freeman, S. T., E. G. Heath, P. D. Guptil, and J. T. Waggoner
(1992).
Seismic hazard assessment, Newport–Inglewood fault zone, in
En-
gineering Geology Practice in Southern California, B. W. Pipkin
and
R. J. Proctor (Editors), Star, Belmont, California, 211–231.
Grant, L. B., and T. K. Rockwell (2002). A northward-
propagating earth-
quake sequence in coastal southern California? Seism. Res.
Lett. 73,
no. 4, 461–469.
Grant, L. B., J. T. Waggoner, C. von Stein, and T. K. Rockwell
(1997).
Paleoseismicity of the north branch of the Newport–Inglewood
fault
zone in Huntington Beach, California, from cone penetrometer
test
data, Bull. Seism. Soc. Am. 87, 277–293.
Grant, L. B., K. J. Mueller, E. M. Gath, H. Cheng, R. L.
Edwards, R. Munro,
and G. L. Kennedy (1999). Late Quaternary uplift and
earthquake
potential of the San Joaquin Hills, southern Los Angeles basin,
Cali-
fornia, Geology 27, 1031–1034.
Grant, L. B., K. J. Mueller, E. M. Gath, and R. Munro (2000).
Late Qua-
ternary uplift and earthquake potential of the San Joaquin Hills,
south-
ern Los Angeles basin, California—Reply, Geology 28, 384.
Grant, L. B., L. J. Ballenger, and E. E. Runnerstrom (2002).
Coastal uplift
of the San Joaquin Hills, southern Los Angeles Basin,
California, by
a large earthquake since 1635 A.D., Bull. Seism. Soc. Am. 92,
no. 2,
590–599.
Harding, T. P. (1973). Newport–Inglewood Trend, California:
an example
of wrenching style of deformation, Am. Assoc. Petrol. Geol.
Bull. 57,
no. 1, 97–116.
Hauksson, E., and S. Gross (1991). Source parameters of the
1933 Long
Beach earthquake, Bull. Seism. Soc. Am. 81, 81–98.
Hill, M. L. (1971). Newport–Inglewood zone and Mesozoic
subduction,
California, Geol. Soc. Am. Bull. 82, 2957–2962.
Legg, M. R. (1991). Sea Beam evidence of recent tectonic
activity in the
California Continental Borderland, in The Gulf and Peninsular
Prov-
ince of the Californias, P. Dauphin and G. Ness (Editors),
American
Association of Petroleum Geologists Memoir 47, 179–196.
Lindvall, S. C., and T. K. Rockwell (1995). Holocene activity of
the Rose
Canyon fault zone in San Diego, California, J. Geophys. Res.
100,
no. B12, 24,121–24,132.
Morton, P. K., and R. V. Miller (1981). Geologic map of Orange
County
California, showing mines and mineral deposits, Calif. Div.
Mines
Geol. Bull. 204, plate 1, 1:48,000 scale.
Richards-Dinger, K. B., and P. M. Shearer (2000). Earthquake
locations in
southern California obtained using source-specific station
terms,
J. Geophys Res. 105, 10,939–10,960.
Rivero, C., J. H. Shaw, and K. Mueller (2000). Oceanside and
Thirty-Mile
Bank blind thrusts: implications for earthquake hazards in
coastal
southern California, Geology 28, 891–894.
Rockwell, T., and M. Murbach (1999). Holocene earthquake
history of the
Rose Canyon fault zone: final technical report submitted for
USGS
Grant No. 1434-95-G-2613, 37 pp.
Shearer, P. M. (1997). Improving local earthquake locations
using the L1-
norm and waveform cross-correlation: application to the
Whittier Nar-
rows, California, aftershock sequence, J. Geophys. Res. 102,
8269–
8283.
Shearer, P. M. (1998). Evidence from a cluster of small
earthquakes for a
fault at 18 km depth beneath Oak Ridge, southern California,
Bull.
Seism. Soc. Am. 88, 1327–1336.
Shearer, P. M. (2002). Parallel faults strands at 9-km depth
resolved on the
Imperial Fault, southern California, Geophys. Res. Lett. 29, no.
14,
1674.
Shearer, P. M., J. L. Hardebeck, L. Astiz, and K. B. Richards-
Dinger
(2003). Analysis of similar event clusters in aftershocks of the
1994
Northridge, California, earthquake, J. Geophys Res. 108, no.
B1,
2035, doi 10.1029/2001JB000685.
Southern California Earthquake Center Working Group on the
Probabilities
of Future Large Earthquakes in Southern California (SCECWG)
(1995). Seismic hazards in southern California: probable
earthquakes,
1994–2024, Bull. Seism. Soc. Am. 85, 379–439.
Walls, C., T. Rockwell, K. Mueller, Y. Bock, S. Williams, J.
Pfanner, J.
Dolan, and P. Feng (1998). Escape tectonics in the Los Angeles
met-
ropolitan region and implications for seismic risk, Nature 394,
356–
360.
Wells, D. L., and K. J. Coppersmith (1994). New empirical
relationships
among magnitude, rupture length, rupture area, and surface
displace-
ment, Bull. Seism. Soc. Am. 84, 974–1002.
Wilcox, R. E., T. P. Harding, and D. R. Seely (1973). Basic
wrench tec-
tonics, Am. Assoc. Petrol. Geol. Bull. 57, no. 1, 74–96.
Wright, T. L. (1991). Structural geology and tectonic evolution
of the Los
Angeles basin, In Active Margin Basins, K. T. Biddle (Editor),
Amer-
ican Association of Petroleum Geologists Memoir 52, 35–134.
Yeats, R. S. (1973). Newport–Inglewood fault zone, Los
Angeles basin,
California, Am. Assoc. Petrol. Geol. Bull. 57, no. 1, 117–135.
Yeats, R. S. (2001). Living with Earthquakes in California,
Oregon State
University Press, Corvallis, 406 pp.
Department of Environmental Health, Science, and Policy
University of California
Irvine, California 92697-7070
[email protected]
(L.B.G.)
Institute of Geophysics and Planetary Physics
Scripps Institution of Oceanography
University of California, San Diego
La Jolla, California 92093-0225
[email protected]
(P.M.S.)
Manuscript received 23 July 2003.
Paper Outline Due Thursday April 23
Posted on GWC Blackboard 3-5-20
Announced in class and illustrated on Thursday March 5, 2020
Required:
1. Two (2) reviewed articles with a person or persons as the
author. The U.S.G.S or the
EPA is not a person. The last pages of your article must
have a reference page. No
reference page or named person or multiple authors are
not acceptable.
2. Two but not more than five pages of text double or 1.5
spaced with normal margins like those in the sample outline.
3. Sources that are good various reviewed journals Science,
Nature, Scientific
American. These publications will have sources at the end
with references
within the text of the paper.
4. Sources to stay away from Science Direct, Geology.com,
National Geographic,
Newspapers, Magazines. If the article you chose has
advertisements then stay away
from these sources, or any with advertising ads.
What to hand in.
1. Your outline as a hard copy.
2. Complete printed out copies of your sources all clipped
together.
3. Must be clipped together with the outline. No folded corners
on the papers
as an attempt to keep them together. Use a clip.
Failure to provide a copy of your sources along with the outline
will result in a failing grade for this outline. No late
assignments will be accepted. No electronic files accepted.
Student Name Bud BennemanGeology 105 Spring 2020Paper Outline.docx

More Related Content

Similar to Student Name Bud BennemanGeology 105 Spring 2020Paper Outline.docx

Geology of Los Angeles AEG Final Paper
Geology of Los Angeles AEG Final PaperGeology of Los Angeles AEG Final Paper
Geology of Los Angeles AEG Final PaperEldon Gath
 
Earth_Science_&_Climate_Change_Somenath_Ganguly
Earth_Science_&_Climate_Change_Somenath_GangulyEarth_Science_&_Climate_Change_Somenath_Ganguly
Earth_Science_&_Climate_Change_Somenath_GangulySomenath Ganguly
 
Sea Level Changes as recorded in nature itself
Sea Level Changes as recorded in nature itselfSea Level Changes as recorded in nature itself
Sea Level Changes as recorded in nature itself
IJERA Editor
 
Sea level changes
Sea  level  changesSea  level  changes
Sea level changes
Pramoda Raj
 
impactos del cambio climatico en ecosistemas costeros
 impactos del cambio climatico en ecosistemas costeros impactos del cambio climatico en ecosistemas costeros
impactos del cambio climatico en ecosistemas costeros
Xin San
 
Paper 2
Paper 2Paper 2
Climate Change In Scotland
Climate Change In ScotlandClimate Change In Scotland
Climate Change In Scotland
Beth Salazar
 
Saltwater intrusion model pompano fl
Saltwater intrusion model pompano flSaltwater intrusion model pompano fl
Saltwater intrusion model pompano fl
Darrel Dunn
 
The Shifting Sands of Noosa
The Shifting Sands of Noosa The Shifting Sands of Noosa
The Shifting Sands of Noosa
Dr Brian Stockwell
 
Wathen 2011 Climatic Changer_Past 2,000 years
Wathen 2011 Climatic Changer_Past 2,000 yearsWathen 2011 Climatic Changer_Past 2,000 years
Wathen 2011 Climatic Changer_Past 2,000 yearsStephen Wathen
 
Review of Long Term Turbidity Trends with Possible Reasons around the World |...
Review of Long Term Turbidity Trends with Possible Reasons around the World |...Review of Long Term Turbidity Trends with Possible Reasons around the World |...
Review of Long Term Turbidity Trends with Possible Reasons around the World |...
CrimsonpublishersMedical
 
USGS: Can Shale Safely Host U.S. Nuclear Waste?
USGS: Can Shale Safely Host U.S. Nuclear Waste?USGS: Can Shale Safely Host U.S. Nuclear Waste?
USGS: Can Shale Safely Host U.S. Nuclear Waste?
Marcellus Drilling News
 
Diana Allen, SFU - Water Science Research: Challenges and Success Stories in ...
Diana Allen, SFU - Water Science Research: Challenges and Success Stories in ...Diana Allen, SFU - Water Science Research: Challenges and Success Stories in ...
Diana Allen, SFU - Water Science Research: Challenges and Success Stories in ...BC Water Science Symposium
 
West Antarctica and threat of a sea level rise disaster
West Antarctica and threat of a sea level rise disasterWest Antarctica and threat of a sea level rise disaster
West Antarctica and threat of a sea level rise disaster
colgan
 
XHAB1_Kudela__Roberts_2003
XHAB1_Kudela__Roberts_2003XHAB1_Kudela__Roberts_2003
XHAB1_Kudela__Roberts_2003phyto
 
Study: Fault activation by hydraulic fracturing in western Canada
Study: Fault activation by hydraulic fracturing in western CanadaStudy: Fault activation by hydraulic fracturing in western Canada
Study: Fault activation by hydraulic fracturing in western Canada
Marcellus Drilling News
 
Stratigraphy-SEQUENCES.pptx
Stratigraphy-SEQUENCES.pptxStratigraphy-SEQUENCES.pptx
Stratigraphy-SEQUENCES.pptx
SaadTaman
 
From the Arctic to the Tropics: The U.S. UNCLOS Bathymetric Mapping Program
From the Arctic to the Tropics: The U.S. UNCLOS Bathymetric Mapping ProgramFrom the Arctic to the Tropics: The U.S. UNCLOS Bathymetric Mapping Program
From the Arctic to the Tropics: The U.S. UNCLOS Bathymetric Mapping Program
Larry Mayer
 

Similar to Student Name Bud BennemanGeology 105 Spring 2020Paper Outline.docx (20)

Geology of Los Angeles AEG Final Paper
Geology of Los Angeles AEG Final PaperGeology of Los Angeles AEG Final Paper
Geology of Los Angeles AEG Final Paper
 
Earth_Science_&_Climate_Change_Somenath_Ganguly
Earth_Science_&_Climate_Change_Somenath_GangulyEarth_Science_&_Climate_Change_Somenath_Ganguly
Earth_Science_&_Climate_Change_Somenath_Ganguly
 
Sea Level Changes as recorded in nature itself
Sea Level Changes as recorded in nature itselfSea Level Changes as recorded in nature itself
Sea Level Changes as recorded in nature itself
 
Sea level changes
Sea  level  changesSea  level  changes
Sea level changes
 
impactos del cambio climatico en ecosistemas costeros
 impactos del cambio climatico en ecosistemas costeros impactos del cambio climatico en ecosistemas costeros
impactos del cambio climatico en ecosistemas costeros
 
Paper 2
Paper 2Paper 2
Paper 2
 
Letter commissioner maria damanaki 10 9-2012
Letter commissioner maria damanaki 10 9-2012Letter commissioner maria damanaki 10 9-2012
Letter commissioner maria damanaki 10 9-2012
 
Climate Change In Scotland
Climate Change In ScotlandClimate Change In Scotland
Climate Change In Scotland
 
Saltwater intrusion model pompano fl
Saltwater intrusion model pompano flSaltwater intrusion model pompano fl
Saltwater intrusion model pompano fl
 
The Shifting Sands of Noosa
The Shifting Sands of Noosa The Shifting Sands of Noosa
The Shifting Sands of Noosa
 
Wathen 2011 Climatic Changer_Past 2,000 years
Wathen 2011 Climatic Changer_Past 2,000 yearsWathen 2011 Climatic Changer_Past 2,000 years
Wathen 2011 Climatic Changer_Past 2,000 years
 
Review of Long Term Turbidity Trends with Possible Reasons around the World |...
Review of Long Term Turbidity Trends with Possible Reasons around the World |...Review of Long Term Turbidity Trends with Possible Reasons around the World |...
Review of Long Term Turbidity Trends with Possible Reasons around the World |...
 
USGS: Can Shale Safely Host U.S. Nuclear Waste?
USGS: Can Shale Safely Host U.S. Nuclear Waste?USGS: Can Shale Safely Host U.S. Nuclear Waste?
USGS: Can Shale Safely Host U.S. Nuclear Waste?
 
Diana Allen, SFU - Water Science Research: Challenges and Success Stories in ...
Diana Allen, SFU - Water Science Research: Challenges and Success Stories in ...Diana Allen, SFU - Water Science Research: Challenges and Success Stories in ...
Diana Allen, SFU - Water Science Research: Challenges and Success Stories in ...
 
West Antarctica and threat of a sea level rise disaster
West Antarctica and threat of a sea level rise disasterWest Antarctica and threat of a sea level rise disaster
West Antarctica and threat of a sea level rise disaster
 
XHAB1_Kudela__Roberts_2003
XHAB1_Kudela__Roberts_2003XHAB1_Kudela__Roberts_2003
XHAB1_Kudela__Roberts_2003
 
Study: Fault activation by hydraulic fracturing in western Canada
Study: Fault activation by hydraulic fracturing in western CanadaStudy: Fault activation by hydraulic fracturing in western Canada
Study: Fault activation by hydraulic fracturing in western Canada
 
Stratigraphy-SEQUENCES.pptx
Stratigraphy-SEQUENCES.pptxStratigraphy-SEQUENCES.pptx
Stratigraphy-SEQUENCES.pptx
 
Global warming
Global warmingGlobal warming
Global warming
 
From the Arctic to the Tropics: The U.S. UNCLOS Bathymetric Mapping Program
From the Arctic to the Tropics: The U.S. UNCLOS Bathymetric Mapping ProgramFrom the Arctic to the Tropics: The U.S. UNCLOS Bathymetric Mapping Program
From the Arctic to the Tropics: The U.S. UNCLOS Bathymetric Mapping Program
 

More from deanmtaylor1545

Assignment 1  Dealing with Diversity in America from Reconstructi.docx
Assignment 1  Dealing with Diversity in America from Reconstructi.docxAssignment 1  Dealing with Diversity in America from Reconstructi.docx
Assignment 1  Dealing with Diversity in America from Reconstructi.docx
deanmtaylor1545
 
Assignment 1 Why are the originalraw data not readily us.docx
Assignment 1 Why are the originalraw data not readily us.docxAssignment 1 Why are the originalraw data not readily us.docx
Assignment 1 Why are the originalraw data not readily us.docx
deanmtaylor1545
 
Assignment 1 Refer to the attached document and complete the .docx
Assignment 1 Refer to the attached document and complete the .docxAssignment 1 Refer to the attached document and complete the .docx
Assignment 1 Refer to the attached document and complete the .docx
deanmtaylor1545
 
Assignment 1 Remote Access Method EvaluationLearning Ob.docx
Assignment 1 Remote Access Method EvaluationLearning Ob.docxAssignment 1 Remote Access Method EvaluationLearning Ob.docx
Assignment 1 Remote Access Method EvaluationLearning Ob.docx
deanmtaylor1545
 
Assignment 1 Please read ALL directions below before startin.docx
Assignment 1 Please read ALL directions below before startin.docxAssignment 1 Please read ALL directions below before startin.docx
Assignment 1 Please read ALL directions below before startin.docx
deanmtaylor1545
 
Assignment 1 Inmates Rights and Special CircumstancesCriteria.docx
Assignment 1 Inmates Rights and Special CircumstancesCriteria.docxAssignment 1 Inmates Rights and Special CircumstancesCriteria.docx
Assignment 1 Inmates Rights and Special CircumstancesCriteria.docx
deanmtaylor1545
 
Assignment 1 Go back through the business press (Fortune, The Ec.docx
Assignment 1 Go back through the business press (Fortune, The Ec.docxAssignment 1 Go back through the business press (Fortune, The Ec.docx
Assignment 1 Go back through the business press (Fortune, The Ec.docx
deanmtaylor1545
 
Assignment 1 Discussion—Environmental FactorsIn this assignment, .docx
Assignment 1 Discussion—Environmental FactorsIn this assignment, .docxAssignment 1 Discussion—Environmental FactorsIn this assignment, .docx
Assignment 1 Discussion—Environmental FactorsIn this assignment, .docx
deanmtaylor1545
 
Assignment 1 1. Using a Microsoft Word document, please post one.docx
Assignment 1 1. Using a Microsoft Word document, please post one.docxAssignment 1 1. Using a Microsoft Word document, please post one.docx
Assignment 1 1. Using a Microsoft Word document, please post one.docx
deanmtaylor1545
 
Assignment 1  Dealing with Diversity in America from Reconstructi.docx
Assignment 1  Dealing with Diversity in America from Reconstructi.docxAssignment 1  Dealing with Diversity in America from Reconstructi.docx
Assignment 1  Dealing with Diversity in America from Reconstructi.docx
deanmtaylor1545
 
Assignment 1  Due Monday 92319 By using linear and nonlinear .docx
Assignment 1  Due Monday 92319 By using linear and nonlinear .docxAssignment 1  Due Monday 92319 By using linear and nonlinear .docx
Assignment 1  Due Monday 92319 By using linear and nonlinear .docx
deanmtaylor1545
 
Assignment 1This assignment is due in Module 8. There are many v.docx
Assignment 1This assignment is due in Module 8. There are many v.docxAssignment 1This assignment is due in Module 8. There are many v.docx
Assignment 1This assignment is due in Module 8. There are many v.docx
deanmtaylor1545
 
Assignment 1TextbookInformation Systems for Business and Beyond.docx
Assignment 1TextbookInformation Systems for Business and Beyond.docxAssignment 1TextbookInformation Systems for Business and Beyond.docx
Assignment 1TextbookInformation Systems for Business and Beyond.docx
deanmtaylor1545
 
ASSIGNMENT 1TASK FORCE COMMITTEE REPORTISSUE AND SOLUTI.docx
ASSIGNMENT 1TASK FORCE COMMITTEE REPORTISSUE AND SOLUTI.docxASSIGNMENT 1TASK FORCE COMMITTEE REPORTISSUE AND SOLUTI.docx
ASSIGNMENT 1TASK FORCE COMMITTEE REPORTISSUE AND SOLUTI.docx
deanmtaylor1545
 
Assignment 1Select one of these three philosophers (Rousseau, Lo.docx
Assignment 1Select one of these three philosophers (Rousseau, Lo.docxAssignment 1Select one of these three philosophers (Rousseau, Lo.docx
Assignment 1Select one of these three philosophers (Rousseau, Lo.docx
deanmtaylor1545
 
Assignment 1Scenario 1You are developing a Windows auditing pl.docx
Assignment 1Scenario 1You are developing a Windows auditing pl.docxAssignment 1Scenario 1You are developing a Windows auditing pl.docx
Assignment 1Scenario 1You are developing a Windows auditing pl.docx
deanmtaylor1545
 
Assignment 1Research by finding an article or case study discus.docx
Assignment 1Research by finding an article or case study discus.docxAssignment 1Research by finding an article or case study discus.docx
Assignment 1Research by finding an article or case study discus.docx
deanmtaylor1545
 
Assignment 1Positioning Statement and MottoUse the pro.docx
Assignment 1Positioning Statement and MottoUse the pro.docxAssignment 1Positioning Statement and MottoUse the pro.docx
Assignment 1Positioning Statement and MottoUse the pro.docx
deanmtaylor1545
 
ASSIGNMENT 1Hearing Versus ListeningDescribe how you le.docx
ASSIGNMENT 1Hearing Versus ListeningDescribe how you le.docxASSIGNMENT 1Hearing Versus ListeningDescribe how you le.docx
ASSIGNMENT 1Hearing Versus ListeningDescribe how you le.docx
deanmtaylor1545
 
assignment 1Essay Nuclear ProliferationThe proliferation of.docx
assignment 1Essay Nuclear ProliferationThe proliferation of.docxassignment 1Essay Nuclear ProliferationThe proliferation of.docx
assignment 1Essay Nuclear ProliferationThe proliferation of.docx
deanmtaylor1545
 

More from deanmtaylor1545 (20)

Assignment 1  Dealing with Diversity in America from Reconstructi.docx
Assignment 1  Dealing with Diversity in America from Reconstructi.docxAssignment 1  Dealing with Diversity in America from Reconstructi.docx
Assignment 1  Dealing with Diversity in America from Reconstructi.docx
 
Assignment 1 Why are the originalraw data not readily us.docx
Assignment 1 Why are the originalraw data not readily us.docxAssignment 1 Why are the originalraw data not readily us.docx
Assignment 1 Why are the originalraw data not readily us.docx
 
Assignment 1 Refer to the attached document and complete the .docx
Assignment 1 Refer to the attached document and complete the .docxAssignment 1 Refer to the attached document and complete the .docx
Assignment 1 Refer to the attached document and complete the .docx
 
Assignment 1 Remote Access Method EvaluationLearning Ob.docx
Assignment 1 Remote Access Method EvaluationLearning Ob.docxAssignment 1 Remote Access Method EvaluationLearning Ob.docx
Assignment 1 Remote Access Method EvaluationLearning Ob.docx
 
Assignment 1 Please read ALL directions below before startin.docx
Assignment 1 Please read ALL directions below before startin.docxAssignment 1 Please read ALL directions below before startin.docx
Assignment 1 Please read ALL directions below before startin.docx
 
Assignment 1 Inmates Rights and Special CircumstancesCriteria.docx
Assignment 1 Inmates Rights and Special CircumstancesCriteria.docxAssignment 1 Inmates Rights and Special CircumstancesCriteria.docx
Assignment 1 Inmates Rights and Special CircumstancesCriteria.docx
 
Assignment 1 Go back through the business press (Fortune, The Ec.docx
Assignment 1 Go back through the business press (Fortune, The Ec.docxAssignment 1 Go back through the business press (Fortune, The Ec.docx
Assignment 1 Go back through the business press (Fortune, The Ec.docx
 
Assignment 1 Discussion—Environmental FactorsIn this assignment, .docx
Assignment 1 Discussion—Environmental FactorsIn this assignment, .docxAssignment 1 Discussion—Environmental FactorsIn this assignment, .docx
Assignment 1 Discussion—Environmental FactorsIn this assignment, .docx
 
Assignment 1 1. Using a Microsoft Word document, please post one.docx
Assignment 1 1. Using a Microsoft Word document, please post one.docxAssignment 1 1. Using a Microsoft Word document, please post one.docx
Assignment 1 1. Using a Microsoft Word document, please post one.docx
 
Assignment 1  Dealing with Diversity in America from Reconstructi.docx
Assignment 1  Dealing with Diversity in America from Reconstructi.docxAssignment 1  Dealing with Diversity in America from Reconstructi.docx
Assignment 1  Dealing with Diversity in America from Reconstructi.docx
 
Assignment 1  Due Monday 92319 By using linear and nonlinear .docx
Assignment 1  Due Monday 92319 By using linear and nonlinear .docxAssignment 1  Due Monday 92319 By using linear and nonlinear .docx
Assignment 1  Due Monday 92319 By using linear and nonlinear .docx
 
Assignment 1This assignment is due in Module 8. There are many v.docx
Assignment 1This assignment is due in Module 8. There are many v.docxAssignment 1This assignment is due in Module 8. There are many v.docx
Assignment 1This assignment is due in Module 8. There are many v.docx
 
Assignment 1TextbookInformation Systems for Business and Beyond.docx
Assignment 1TextbookInformation Systems for Business and Beyond.docxAssignment 1TextbookInformation Systems for Business and Beyond.docx
Assignment 1TextbookInformation Systems for Business and Beyond.docx
 
ASSIGNMENT 1TASK FORCE COMMITTEE REPORTISSUE AND SOLUTI.docx
ASSIGNMENT 1TASK FORCE COMMITTEE REPORTISSUE AND SOLUTI.docxASSIGNMENT 1TASK FORCE COMMITTEE REPORTISSUE AND SOLUTI.docx
ASSIGNMENT 1TASK FORCE COMMITTEE REPORTISSUE AND SOLUTI.docx
 
Assignment 1Select one of these three philosophers (Rousseau, Lo.docx
Assignment 1Select one of these three philosophers (Rousseau, Lo.docxAssignment 1Select one of these three philosophers (Rousseau, Lo.docx
Assignment 1Select one of these three philosophers (Rousseau, Lo.docx
 
Assignment 1Scenario 1You are developing a Windows auditing pl.docx
Assignment 1Scenario 1You are developing a Windows auditing pl.docxAssignment 1Scenario 1You are developing a Windows auditing pl.docx
Assignment 1Scenario 1You are developing a Windows auditing pl.docx
 
Assignment 1Research by finding an article or case study discus.docx
Assignment 1Research by finding an article or case study discus.docxAssignment 1Research by finding an article or case study discus.docx
Assignment 1Research by finding an article or case study discus.docx
 
Assignment 1Positioning Statement and MottoUse the pro.docx
Assignment 1Positioning Statement and MottoUse the pro.docxAssignment 1Positioning Statement and MottoUse the pro.docx
Assignment 1Positioning Statement and MottoUse the pro.docx
 
ASSIGNMENT 1Hearing Versus ListeningDescribe how you le.docx
ASSIGNMENT 1Hearing Versus ListeningDescribe how you le.docxASSIGNMENT 1Hearing Versus ListeningDescribe how you le.docx
ASSIGNMENT 1Hearing Versus ListeningDescribe how you le.docx
 
assignment 1Essay Nuclear ProliferationThe proliferation of.docx
assignment 1Essay Nuclear ProliferationThe proliferation of.docxassignment 1Essay Nuclear ProliferationThe proliferation of.docx
assignment 1Essay Nuclear ProliferationThe proliferation of.docx
 

Recently uploaded

Instructions for Submissions thorugh G- Classroom.pptx
Instructions for Submissions thorugh G- Classroom.pptxInstructions for Submissions thorugh G- Classroom.pptx
Instructions for Submissions thorugh G- Classroom.pptx
Jheel Barad
 
Sha'Carri Richardson Presentation 202345
Sha'Carri Richardson Presentation 202345Sha'Carri Richardson Presentation 202345
Sha'Carri Richardson Presentation 202345
beazzy04
 
June 3, 2024 Anti-Semitism Letter Sent to MIT President Kornbluth and MIT Cor...
June 3, 2024 Anti-Semitism Letter Sent to MIT President Kornbluth and MIT Cor...June 3, 2024 Anti-Semitism Letter Sent to MIT President Kornbluth and MIT Cor...
June 3, 2024 Anti-Semitism Letter Sent to MIT President Kornbluth and MIT Cor...
Levi Shapiro
 
Synthetic Fiber Construction in lab .pptx
Synthetic Fiber Construction in lab .pptxSynthetic Fiber Construction in lab .pptx
Synthetic Fiber Construction in lab .pptx
Pavel ( NSTU)
 
Embracing GenAI - A Strategic Imperative
Embracing GenAI - A Strategic ImperativeEmbracing GenAI - A Strategic Imperative
Embracing GenAI - A Strategic Imperative
Peter Windle
 
TESDA TM1 REVIEWER FOR NATIONAL ASSESSMENT WRITTEN AND ORAL QUESTIONS WITH A...
TESDA TM1 REVIEWER  FOR NATIONAL ASSESSMENT WRITTEN AND ORAL QUESTIONS WITH A...TESDA TM1 REVIEWER  FOR NATIONAL ASSESSMENT WRITTEN AND ORAL QUESTIONS WITH A...
TESDA TM1 REVIEWER FOR NATIONAL ASSESSMENT WRITTEN AND ORAL QUESTIONS WITH A...
EugeneSaldivar
 
Unit 2- Research Aptitude (UGC NET Paper I).pdf
Unit 2- Research Aptitude (UGC NET Paper I).pdfUnit 2- Research Aptitude (UGC NET Paper I).pdf
Unit 2- Research Aptitude (UGC NET Paper I).pdf
Thiyagu K
 
The geography of Taylor Swift - some ideas
The geography of Taylor Swift - some ideasThe geography of Taylor Swift - some ideas
The geography of Taylor Swift - some ideas
GeoBlogs
 
678020731-Sumas-y-Restas-Para-Colorear.pdf
678020731-Sumas-y-Restas-Para-Colorear.pdf678020731-Sumas-y-Restas-Para-Colorear.pdf
678020731-Sumas-y-Restas-Para-Colorear.pdf
CarlosHernanMontoyab2
 
BÀI TẬP BỔ TRỢ TIẾNG ANH GLOBAL SUCCESS LỚP 3 - CẢ NĂM (CÓ FILE NGHE VÀ ĐÁP Á...
BÀI TẬP BỔ TRỢ TIẾNG ANH GLOBAL SUCCESS LỚP 3 - CẢ NĂM (CÓ FILE NGHE VÀ ĐÁP Á...BÀI TẬP BỔ TRỢ TIẾNG ANH GLOBAL SUCCESS LỚP 3 - CẢ NĂM (CÓ FILE NGHE VÀ ĐÁP Á...
BÀI TẬP BỔ TRỢ TIẾNG ANH GLOBAL SUCCESS LỚP 3 - CẢ NĂM (CÓ FILE NGHE VÀ ĐÁP Á...
Nguyen Thanh Tu Collection
 
1.4 modern child centered education - mahatma gandhi-2.pptx
1.4 modern child centered education - mahatma gandhi-2.pptx1.4 modern child centered education - mahatma gandhi-2.pptx
1.4 modern child centered education - mahatma gandhi-2.pptx
JosvitaDsouza2
 
"Protectable subject matters, Protection in biotechnology, Protection of othe...
"Protectable subject matters, Protection in biotechnology, Protection of othe..."Protectable subject matters, Protection in biotechnology, Protection of othe...
"Protectable subject matters, Protection in biotechnology, Protection of othe...
SACHIN R KONDAGURI
 
Unit 8 - Information and Communication Technology (Paper I).pdf
Unit 8 - Information and Communication Technology (Paper I).pdfUnit 8 - Information and Communication Technology (Paper I).pdf
Unit 8 - Information and Communication Technology (Paper I).pdf
Thiyagu K
 
2024.06.01 Introducing a competency framework for languag learning materials ...
2024.06.01 Introducing a competency framework for languag learning materials ...2024.06.01 Introducing a competency framework for languag learning materials ...
2024.06.01 Introducing a competency framework for languag learning materials ...
Sandy Millin
 
The basics of sentences session 5pptx.pptx
The basics of sentences session 5pptx.pptxThe basics of sentences session 5pptx.pptx
The basics of sentences session 5pptx.pptx
heathfieldcps1
 
aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa
aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa
aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa
siemaillard
 
How to Make a Field invisible in Odoo 17
How to Make a Field invisible in Odoo 17How to Make a Field invisible in Odoo 17
How to Make a Field invisible in Odoo 17
Celine George
 
Phrasal Verbs.XXXXXXXXXXXXXXXXXXXXXXXXXX
Phrasal Verbs.XXXXXXXXXXXXXXXXXXXXXXXXXXPhrasal Verbs.XXXXXXXXXXXXXXXXXXXXXXXXXX
Phrasal Verbs.XXXXXXXXXXXXXXXXXXXXXXXXXX
MIRIAMSALINAS13
 
Model Attribute Check Company Auto Property
Model Attribute  Check Company Auto PropertyModel Attribute  Check Company Auto Property
Model Attribute Check Company Auto Property
Celine George
 
Digital Tools and AI for Teaching Learning and Research
Digital Tools and AI for Teaching Learning and ResearchDigital Tools and AI for Teaching Learning and Research
Digital Tools and AI for Teaching Learning and Research
Vikramjit Singh
 

Recently uploaded (20)

Instructions for Submissions thorugh G- Classroom.pptx
Instructions for Submissions thorugh G- Classroom.pptxInstructions for Submissions thorugh G- Classroom.pptx
Instructions for Submissions thorugh G- Classroom.pptx
 
Sha'Carri Richardson Presentation 202345
Sha'Carri Richardson Presentation 202345Sha'Carri Richardson Presentation 202345
Sha'Carri Richardson Presentation 202345
 
June 3, 2024 Anti-Semitism Letter Sent to MIT President Kornbluth and MIT Cor...
June 3, 2024 Anti-Semitism Letter Sent to MIT President Kornbluth and MIT Cor...June 3, 2024 Anti-Semitism Letter Sent to MIT President Kornbluth and MIT Cor...
June 3, 2024 Anti-Semitism Letter Sent to MIT President Kornbluth and MIT Cor...
 
Synthetic Fiber Construction in lab .pptx
Synthetic Fiber Construction in lab .pptxSynthetic Fiber Construction in lab .pptx
Synthetic Fiber Construction in lab .pptx
 
Embracing GenAI - A Strategic Imperative
Embracing GenAI - A Strategic ImperativeEmbracing GenAI - A Strategic Imperative
Embracing GenAI - A Strategic Imperative
 
TESDA TM1 REVIEWER FOR NATIONAL ASSESSMENT WRITTEN AND ORAL QUESTIONS WITH A...
TESDA TM1 REVIEWER  FOR NATIONAL ASSESSMENT WRITTEN AND ORAL QUESTIONS WITH A...TESDA TM1 REVIEWER  FOR NATIONAL ASSESSMENT WRITTEN AND ORAL QUESTIONS WITH A...
TESDA TM1 REVIEWER FOR NATIONAL ASSESSMENT WRITTEN AND ORAL QUESTIONS WITH A...
 
Unit 2- Research Aptitude (UGC NET Paper I).pdf
Unit 2- Research Aptitude (UGC NET Paper I).pdfUnit 2- Research Aptitude (UGC NET Paper I).pdf
Unit 2- Research Aptitude (UGC NET Paper I).pdf
 
The geography of Taylor Swift - some ideas
The geography of Taylor Swift - some ideasThe geography of Taylor Swift - some ideas
The geography of Taylor Swift - some ideas
 
678020731-Sumas-y-Restas-Para-Colorear.pdf
678020731-Sumas-y-Restas-Para-Colorear.pdf678020731-Sumas-y-Restas-Para-Colorear.pdf
678020731-Sumas-y-Restas-Para-Colorear.pdf
 
BÀI TẬP BỔ TRỢ TIẾNG ANH GLOBAL SUCCESS LỚP 3 - CẢ NĂM (CÓ FILE NGHE VÀ ĐÁP Á...
BÀI TẬP BỔ TRỢ TIẾNG ANH GLOBAL SUCCESS LỚP 3 - CẢ NĂM (CÓ FILE NGHE VÀ ĐÁP Á...BÀI TẬP BỔ TRỢ TIẾNG ANH GLOBAL SUCCESS LỚP 3 - CẢ NĂM (CÓ FILE NGHE VÀ ĐÁP Á...
BÀI TẬP BỔ TRỢ TIẾNG ANH GLOBAL SUCCESS LỚP 3 - CẢ NĂM (CÓ FILE NGHE VÀ ĐÁP Á...
 
1.4 modern child centered education - mahatma gandhi-2.pptx
1.4 modern child centered education - mahatma gandhi-2.pptx1.4 modern child centered education - mahatma gandhi-2.pptx
1.4 modern child centered education - mahatma gandhi-2.pptx
 
"Protectable subject matters, Protection in biotechnology, Protection of othe...
"Protectable subject matters, Protection in biotechnology, Protection of othe..."Protectable subject matters, Protection in biotechnology, Protection of othe...
"Protectable subject matters, Protection in biotechnology, Protection of othe...
 
Unit 8 - Information and Communication Technology (Paper I).pdf
Unit 8 - Information and Communication Technology (Paper I).pdfUnit 8 - Information and Communication Technology (Paper I).pdf
Unit 8 - Information and Communication Technology (Paper I).pdf
 
2024.06.01 Introducing a competency framework for languag learning materials ...
2024.06.01 Introducing a competency framework for languag learning materials ...2024.06.01 Introducing a competency framework for languag learning materials ...
2024.06.01 Introducing a competency framework for languag learning materials ...
 
The basics of sentences session 5pptx.pptx
The basics of sentences session 5pptx.pptxThe basics of sentences session 5pptx.pptx
The basics of sentences session 5pptx.pptx
 
aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa
aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa
aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa
 
How to Make a Field invisible in Odoo 17
How to Make a Field invisible in Odoo 17How to Make a Field invisible in Odoo 17
How to Make a Field invisible in Odoo 17
 
Phrasal Verbs.XXXXXXXXXXXXXXXXXXXXXXXXXX
Phrasal Verbs.XXXXXXXXXXXXXXXXXXXXXXXXXXPhrasal Verbs.XXXXXXXXXXXXXXXXXXXXXXXXXX
Phrasal Verbs.XXXXXXXXXXXXXXXXXXXXXXXXXX
 
Model Attribute Check Company Auto Property
Model Attribute  Check Company Auto PropertyModel Attribute  Check Company Auto Property
Model Attribute Check Company Auto Property
 
Digital Tools and AI for Teaching Learning and Research
Digital Tools and AI for Teaching Learning and ResearchDigital Tools and AI for Teaching Learning and Research
Digital Tools and AI for Teaching Learning and Research
 

Student Name Bud BennemanGeology 105 Spring 2020Paper Outline.docx

  • 1. Student Name: Bud Benneman Geology 105 Spring 2020 Paper Outline The Newport–Inglewood fault zone (NIFZ) of southern California. I. The Newport–Inglewood fault zone (NIFZ) was first identified as a significant threat to southern California residents in 1933 when it generated the Magnitude 6.3 Long Beach Earthquake, killing 115 people. A. The Newport Inglewood fault is located in southern Los Angeles County in the city of Inglewood and transverses south to Newport Beach in Orange County where it becomes an off shore fault. B. The NIFZ then connects to the Rose Canyon Fault none, becomes a landward fault in San Diego County. C. This is a stress reliever Strike-Slip Fault Zone associated with the San Andréas Fault Zone. D. Ground water basins in Los Angeles County may be used as predictors of fault movement due to sudden changes in ground water level. E. In southern California, tectonic deformation between the Pacific and North American plates is accommodated primarily by a zone of strike-slip faults, II. This fault is important to the exploration of Oil. A. The NIFZ has been studied extensively in the Los Angeles basin by petroleum geologists.
  • 2. B. The NIFZ overlies a major tectonic boundary separating eastern continental basement rocks of granitic and associated metamorphic rocks of Santa Catalina Island schist. C. The NIFZ follows along a former Mesozoic subduction zone. D. This fault is associated with several oil basins including Newport Beach, Huntington Beach, Seal Beach, Long Beach, and Signal Hill. III. Earthquake Potential for Los Angeles, Orange and San Diego Counties. A. The epicenter for the 1933 Long Beach Quake was located in Newport Beach just south of the Santa Ana River discharge into the Pacific Ocean. B. Orange County in 1933 consisted of small farm-towns, cattle ranching and agricultural fields. C. The closest city of significant development was Lang Beach, which was devastated by the 1933 6,3 quake. D. Today cities of Newport Beach, Huntington Beach, Costa Mesa, are large population centers, which have replaced agriculture as the primary economic sector. E. The rate of ground water basin contraction is important in determining possible earthquake releases in the Los Angeles Basin. F. Many Orange and Los Angeles county buildings are located along high risk development zones during earthquakes. IV. Earthquake Monitoring and Prediction a. The location of the 2000 cluster is southwest of an area of active faulting from 1982 to 1990 along this zone and reported strike-slip offshore fault between Newport Beach and the San Joaquin Hills was traced from seismic activity.
  • 3. b. Analyzed microseismicity from a cluster of epicenters between 2 and 2.5 km along the fault zone suggest potential earthquakes of higher magnitude are possible. c. A possible 7.5 magnitude earthquake could result from rupture of the entire fault. The 6.5-km depth of the Newport Beach seismicity cluster does not provide information on the geometry or interaction between the strike-fault zone and Oceanside thrust faults. d. Los Angeles groundwater basin contraction could cause faults such as the Whittier E. Oil companies and water basin studies using seismic studies and human induced micro-seismic studies have identified several fault structures such as blind thrust faults in the Los Angeles basin. Blind thrust along with strike slip fault structures suggest potential earthquakes of higher magnitude are possible for the Los Angeles and Orange County basins. IV. Conclusion a. The Newport Inglewood Fault is a major stress reliever for the San Andreas Fault Zone. b. This fault zone is tied in between the North American Plate and Pacific Plates Strike Slip Zones and follows a Mesozoic Subduction zone. c. The location of the NIFZ has potential for a large magnitude earthquake due to it being sandwiched between the eastern San Andreas and Pacific Margin plate Boundary.
  • 4. Effects of episodic fluid flow on hydrocarbon migration in the Newport-Inglewood Fault Zone, Southern California B. JUNG1, G. GARVEN 2 AND J. R. BOLES3 1Department of Earth Sciences, Uppsala University, Uppsala, Sweden; 2Department of Earth and Ocean Sciences, Tufts University, Medford, MA, USA; 3Department of Earth Science, University of California, Santa Barbara, CA, USA ABSTRACT Fault permeability may vary through time due to tectonic deformations, transients in pore pressure and effective stress, and mineralization associated with water-rock reactions. Time-varying permeability will affect subsurface fluid migration rates and patterns of petroleum accumulation in densely faulted sedimentary basins such as those associated with the borderland basins of Southern California. This study explores the petroleum fluid dynamics of this migration. As a multiphase flow and petroleum migration case study on the role of faults, computational models for both episodic and continuous hydrocarbon migration are constructed to investigate large-scale fluid
  • 5. flow and petroleum accumulation along a northern section of the Newport-Inglewood fault zone in the Los Angeles basin, Southern California. The numerical code solves the governing equations for oil, water, and heat transport in heterogeneous and anisotropic geologic cross sections but neglects flow in the third dimension for practical applications. Our numerical results suggest that fault permeability and fluid pressure fluctuations are cru- cial factors for distributing hydrocarbon accumulations associated with fault zones, and they also play important roles in controlling the geologic timing for reservoir filling. Episodic flow appears to enhance hydrocarbon accu- mulation more strongly by enabling stepwise build-up in oil saturation in adjacent sedimentary formations due to temporally high pore pressure and high permeability caused by periodic fault rupture. Under assumptions that fault permeability fluctuate within the range of 1–1000 millidarcys (10�15–10�12 m2) and fault pressures fluctuate within 10–80% of overpressure ratio, the estimated oil volume in the Inglewood oil field (approximately 450 mil- lion barrels oil equivalent) can be accumulated in about 24 000 years, assuming a seismically induced fluid flow event occurs every 2000 years. This episodic petroleum migration model could be more geologically important than a continuous-flow model, when considering the observed
  • 6. patterns of hydrocarbons and seismically active tectonic setting of the Los Angeles basin. Key words: episodic fluid flow, fluid flow in faults, multiphase flow in siliciclastic sedimentary basins, petroleum migration Received 21 May 2013; accepted 16 October 2013 Corresponding author: Byeongju Jung, Department of Earth Sciences, Uppsala University, Gl227 Geocentrum, Villav€agen 16B, 753 36 Uppsala, Sweden. Email: [email protected] Tel: +46 018 471 2264. Fax: +1 617 627 3584. Geofluids (2014) 14, 234–250 INTRODUCTION Large-scale faults in sedimentary basins have become increasingly studied due to their important role in convey- ing and compartmentalizing hydrocarbons (Aydin 2000; Boles et al. 2004; Karlsen & Skeie 2006; Kroeger et al. 2009; Zhang et al. 2009; Gong et al. 2011). The hydro- mechanical properties of faults in active continental mar- gins are strongly affected by tectonic deformation, so
  • 7. considering the fluid dynamics of faults will likely improve our understanding of petroleum migration (Reynolds & Lister 1987; Blanpied et al. 1992; Sibson 1994; Appold & Garven 2000; Yamaguchi et al. 2011). For example, there is abundant geological evidence, at both macroscopic and microscopic scales, that faults focus fluid flow over long periods of time but later are sealed by mechanical compac- tion and chemical reactions causing mineral precipitation (Eichhubl & Boles 2000; Caine et al. 2010; Faulkner et al. 2010). The hydrologic activity of fault zones may also depend highly on earthquakes, which in turn may induce periodic fluctuations in pore fluid pressure and fault per- meability (Evans 1992; Sibson 1994). Furthermore, large- © 2013 John Wiley & Sons Ltd Geofluids (2014) 14, 234–250 doi: 10.1111/gfl.12070 scale fault zones may affect regional hydrocarbon migration by regulating the spatial distribution of overpressure in the
  • 8. subsurface (Sperrevik et al. 2002; Fisher et al. 2003; Sork- habi & Tsuji 2005). Laboratory experiments on fractured rock show that active shear faults are more permeable than the adjacent country rock by two to three orders in magni- tude. These faults then become less permeable when deac- tivated (Aydin 2000). The mechanism of recurring fluid pressure build-up, hy- drofracturing, fluid surge, and fault sealing can also be a potential means for hydrocarbon migration (Bradley 1975; Walder & Nur 1984; Mandl & Harkness 1987). For exam- ple, field observations of brecciated rocks and hydrother- mal veins from the Stillwater fault zone in Nevada indicate that petroleum migration was not completed as one single flow event, but rather accumulation too place over many episodes of oil flow during the deformation history (Caine et al. 2010). Geochemical evidence from hydrocarbon con- densates found in the South China basin also support the notion that petroleum migration occurs simultaneously
  • 9. with episodes of hydrothermal fluid flow (Guo et al. 2011). We further hypothesize that the hydrodynamic effects of multiphase flow are more effective for long-distance trans- port, during periods of strong overpressuring associated with episodes of seismically controlled fluid flow. Episodic flow associated with large faults may also be more effective than long-term continuous or steady flow of hydrocarbons, as might be envisioned for a slowly subsiding sedimentary basin. To test this hypothesis, we conduct 2D finite ele- ment simulations for multiphase flow in geologically com- plex cross sections through a basin. We introduce two migration scenarios of continuous flow and of episodic flow, which likely account for the current distribution of hydrocarbon pools such as the Inglewood oil field. The continuous models assume constant fault permeability and fluid pressure throughout the simulation time, while those conditions are time varying in the episodic models. We first
  • 10. compared the fluid pressure, subsurface temperature, and petroleum saturations from these models and then per- formed sensitivity studies on the fault permeability and the frequency of episodic flow pulses to understand how these permeability transients might affect overall hydrocarbon migration and accumulation patterns in a faulted sedimen- tary basin. GEOLOGIC SETTING The Los Angeles (LA) basin is one of the most prolific hydrocarbon-producing areas on Earth, and it hosts histor- ically giant oil fields. From the geological survey, it was recognized that the upper Miocene formations contain organic-rich sediments and play an important role as the primary hydrocarbon source rock. Most of the hydrocar- bons were thermally matured in the central part of the basin (the central syncline area) and have migrated to the edges, which are laterally confined by regional-scale fault zones (Biddle 1991; Jeffrey et al. 1991). The Newport-
  • 11. Inglewood fault zone is one of the major regional fault sys- tems that structurally border the southwestern side of the LA basin, and numerous hydrocarbon reservoirs are closely associated with the fault structure (Fig. 1A). In the Ingle- (A) (B) Fig. 1. Faults and oil fields in the Los Angeles basin (after Wright 1991): (A) LA basin, (B) Inglewood oil field. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 Effects of episodic fluid flow 235 wood oil field located on the northern part of the fault zone, the main production area exists at the intersection between the Newport-Inglewood and Sentous faults and is elongated along the trends of both faults (Fig. 1B). Geo- chemical indicators and biomarkers also suggest the hydro- carbons discovered in this area have mostly migrated approximately 10–15 km from the deep basin center with minor mixing of indigenous petroleum (Kaplan et al.
  • 12. 2000). Tectonic deformation and subsequent seismic activ- ity make this region attractive for studying the relationship between hydrocarbon migration and active fault structures in young sedimentary basins. The Newport-Inglewood fault zone consists of a series of en echelon strike slip faults that were reactivated during the late Pliocene transpressional deformation (Pasadena Orogeny) and transformed into more complicated anticline structures containing normal and reverse faults (Fig. 2). Hydrocarbon reservoirs exist in multiple sedimentary for- mations but are more concentrated in Pliocene strata con- taining a high proportion of sandstone. Many productive petroleum reservoirs align with the trend of the Newport- Inglewood fault zone, which extends and merges with the Central Basin d�ecollement at about 10 km depth (Shaw & Suppe 1996). The permeability of these sandstones ranges from 10’s to 1000’s millidarcys (10�14–10�12 m2) for sandy Plio- cene formations (Hesson & Olilang 1990). The thickness
  • 13. of sediments in the central syncline is approximately 10 km and becomes gradually thinner toward the north- ern and southwestern edges of the basin (Blake 1991; Fig. 3). Figure 4A shows an outcrop of channel-fill sand- stones (Sespe Formation) in the adjacent Santa Barbara area. The faults in these rocks are filled with carbonate mineral precipitates and lithified hydrocarbons, which is strong evidence that the fault zone provided active chan- nels for hydrocarbon fluids, but later these were sealed by subsequent reactions involving diagenetic-hydrother- mal mineralization as fluids cooled or the pressure rap- idly dropped (Fig. 4B). Five separate hydrogeologic units were considered here: middle and pper Miocene, lower and upper Pliocene, and Pleistocene formations (Fig. 3). The middle Miocene unit (Topanga Formation) consists of medium to coarse sandstones with intercalated shale layers. The upper Miocene sediments (Puente Formation) are mostly siltstone and silty sandstone with interbedded
  • 14. pelagic mudstone and shale layers (nodular shales in some areas) that contain high organic carbon contents (10–16%). This formation is often considered to be the primary source rock for petroleum generation (Jeffrey et al. 1991). The lower Pliocene unit (Repetto forma- tion) serves as a major reservoir for hydrocarbon accu- mulation. The Repetto consists of fine to coarse sandstones with interbedded siltstone and shale layers that have relatively high permeability of 10–100 md (10�14–10�13 m2) (Higgins & Chapman 1984). The geology of the upper Pliocene unit (Pico formation) is very similar to the Repetto formation, consisting of inter- bedded sandstone and siltstone, but with slightly lower permeability. The Pleistocene unit (San Pedro formation) consists of relatively uniform and highly permeable sand layers interbedded with minor gravel, silt, and shale lay- ers (Olson 1978). We propose that north–south transpressional tectonic
  • 15. stresses pressurized the basin continuously from the late Pliocene to the present. A coupled mechanism of tec- tonic loading, pore fluid pressure build-up, fault instabil- ity, and fluid flow may have induced episodic fluid flow events (Sibson 1994). Elevated effective stress and pore pressure by tectonic loading compact the sedimentary rocks during the interseismic periods, making the fault zone mechanically unstable. When ruptured, high perme- ability and low pore pressure temporarily create focused fluid flow in the fault zone. The fault stays open for a relatively short period of time due to hydromechanical compaction as the pore pressure decreases, and then sealed further by hydrothermal mineral precipitation (Fig. 5). The continuous tectonic stress and rebuilding of pore pressure cause this cycle to repeat after the fault zone becomes sealed. Field observations of petroleum and carbonate mineral deposits in other siliciclastic faults suggest an episodic nature of injected fluids (Eichhubl &
  • 16. Boles 2000; Garden et al. 2001), and these episodic flow patterns may affect the geohydrologic controls on hydro- carbon accumulation. We modeled this dynamic aspect of fault zone to understand its effects on petroleum migra- tion and entrapment. Fig. 2. Cross section of the Newport-Inglewood fault zone along the tran- sect X–X′ (courtesy of Plains Exploration and Production Company, 2008). © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 236 B. JUNG et al. MULTIPHASE FLOW MODEL To model petroleum migration in the basin, we need to predict fluid saturation and pressure for both the wetting phase (formation water) and nonwetting phase (oil). The governing equations for multiphase fluid flow can be derived from the fluid mass conservation equations, as written by Bear (1972): (A) (B)
  • 17. Fig. 4. Outcrop pictures of organic-rich sedimentary rocks in California: (A) channel-fill sandstones in nonmarine Sespe Formation (Oligocene) from Old San Marcos road, Santa Barbara County, CA, (B) tar-filled fault breccia in Monterey Formation at Arroyo Burro beach, Santa Barbara County, CA. Fig. 3. Lithostratigraphy of the Los Angeles basin, modified from Blake (1991). © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 Effects of episodic fluid flow 237 / @ðqwSwÞ @t ¼ �O � ðqwvwÞ þ qwmw ð1Þ / @ðqnSnÞ @t ¼ / @ðqnð1 � SwÞÞ @t ¼ �O � ðqnvnÞ þ qnmn ð2Þ where / is the effective porosity of a formation, and S is the saturation of the phase: the subscripts n and w denote
  • 18. the nonwetting (liquid petroleum) and wetting (water) phases, respectively. Additionally, q is mass density of the fluid, m is a source/sink term, and v is the Darcy velocity (specific discharge) of each fluid phase, expressed as fol- lows: vw ¼ �kwkðOqw � qwgÞ ð3Þ vn ¼ �knkðOqn � qngÞ ¼ �knðOqw þ Oqc � qgÞ ð4Þ where k is the intrinsic permeability tensor, p is fluid pres- sure, g is a gravitational vector (g = (0, 0, �g)), and pc is capillary pressure (pc = pn – pw). The parameter, k, is the mobility coefficient and defined by the ratio of relative per- meability (kr) and dynamic viscosity (l) as: kw ¼ krw=lw; kn ¼ krn=ln ð5Þ After a few steps of algebraic manipulation, the pressure and saturation equations can be decoupled. If slightly com- pressible fluids are assumed, the final form of average pres- sure and saturation equations can be written as follows (Geiger et al. 2004): /ct @ �P
  • 19. @t ¼ O � kfktO�P � 0:5ðkw � knÞOpc� ðkwqw þ knqnÞggþmt ð6Þ / @Sn @t ¼ O � ½fnvt � �kkfOqc þ ðpw � pnÞgg� � mt ¼ 0 ð7Þ where ct is bulk compressibility of a medium, and mt is a source/sink term. vt is the sum of water and petroleum velocites (vt = vw + vn). The average pressure �P is an arith- metic mean of the water and petroleum pressure, and f is a fractional flow coefficient that is also defined for simplicity: fw ¼ kw kt ; fn ¼ kn kt ; and �k ¼ kwkn kt ð8Þ The first, second, and third terms in the right-hand side
  • 20. of the pressure equation (Eq. 6) represent advection, capil- larity, and buoyancy flow terms, respectively. Conventional multiphase flow equations were decoupled in terms of average pressure and petroleum saturation. These equations were solved using the implicit-pressure explicit-saturation (so called ‘IMPES’) technique (Helmig 1997; Huber & Helmig 2000; Class et al. 2002; Reichen- berger et al. 2006), a technique that produces solutions faster than those requiring time-consuming nonlinear itera- tions. Solution s to the pressure and saturation equations were computed using a hybrid numerical method called FEFVM suggested by Geiger et al. (2004, 2006). This method applies a finite element method (FEM) for com-
  • 21. puting average pressure and then a finite volume method (FVM) for computing fluid saturation. A fully upwind for- mulation and total velocity diminishing (TVD) method (Harten 1997) were also used for solutions to avoid both numerical dispersion and spurious oscillation at the satura- tion front. Capillary pressure and relative permeability models devel- oped by van Genuchten (1980) were used for describing two-phase fluid–solid interaction in the porous media. The van Genuchten model has been widely used in a multiphase flow modeling and well known for providing stable numerical solutions when applying continuous capil- lary pressure functions for a whole saturation interval.
  • 22. Stress or temperature dependent parameters were not included in the numerical formulation. Petroleum density and viscosity were computed using empirical equations suggested by Glasø (1980) and Eng- land et al. (1987). Liquid petroleum density generally decreases with increasing pressure because the solubility of gaseous component increases. The dynamic viscosity of oil Fig. 5. Coupled hydromechanic and hydrother- mal processes for episodic fluid flow (as known as fault-valve mechanism), adopted from Sibson (1994). The parameter and variables: pf is fluid pressure; s is shear stress along the fault; C is
  • 23. the cohesive strength of the fault; ls is the sta- tic coefficient of rock friction, and rn is the nor- mal stress on the fault. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 238 B. JUNG et al. varies between 5 9 10�4 and 5 9 10�2 Pa�s, and generally decreases exponentially with increasing temperature (Eng- land et al. 1987). Fluid properties of the formation water were obtained from a set of state equations proposed by Phillips et al. (1981, 1983) and Watson et al. (1980), which consider the effects of temperature, pressure, and salinity/NaCl concentration on water density.
  • 24. MODEL CONFIGURATION The models, in this research, are based on the cross section in Fig. 2, which illustrates the present geology of the New- port-Inglewood fault zone and associated oil fields (e.g., Inglewood oil field) along the transect X–X′ in Fig. 1B. This profile, which consists of anticline folds with heavily faulted rocks, was chosen because it represents the typical geologic settings of oil reservoirs along the NIFZ. The numerical grid was discretized into 7655 triangular ele- ments using a Delaunay triangulation method for detailed rendering of geological structures (de Berg et al. 2008; Fig. 6). The Newport-Inglewood fault zone and surround- ing areas were divided into smaller elements to increase the
  • 25. resolution of the numerical solution. Because this model is only two-dimensional, a character- istic fault length or effective flow field length was intro- duced for petroleum volume calculation. The total length of Inglewood Oil Field is about 5 km, so we assume that 20% of the total fault length (Approximately 1 km) in pro- file is available for petroleum migration. The width of fault zone in the numerical grid is approximately 50 m. Model parameters used in the simulations are listed in Table 1. Permeability and porosity of each hydrogeologic unit were obtained from several publications (Yerkes 1972; Olson 1978; Higgins & Chapman 1984; Hesson & Olilang
  • 26. 1990; Nishikawa et al. 2009) and chosen within the ranges considered to be representative for these local formations. Fault permeability was not available from any local field measurements, so it was systematically varied or dynami- cally changed as part of model parameter sensitivity analy- sis. Anisotropic permeability ratio values of up to 100:1 (kx/kz), typical for a regional-scale flow, were chosen for most formations except the Pleistocene sediments, which are known to be the most permeable and yet not fully con- solidated. A thermal dispersivity a approximately 100 m was assumed for all formations, reflecting the longitudinal solute dispersivity value for a regional groundwater flow system (de Marsily 1986; Gelhar et al. 1992). Matrix ther-
  • 27. mal conductivity values of 3.0 W m�1°C were assigned to most sandstone-dominant units and values of 2.5– 2.8 W m�1°C were assigned to the units having high con- tent of siltstone and interbedded shale (Blackwell & Steele 1989). A specific heat capacity value of 750 J kg�1°C, typi- cal of sandstone and shale, was assigned for all hydrogeo- logic units and faults (Sabins 1997). Formation of water salinity in the LA basin usually ranges from 20 000 to 34 000 ppm TDS (Hesson & Olilang 1990; California Department of Conservation 1992) for most oil fields. The groundwater salinity was set to 25 000 ppm TDS for the entire model profile. Capillary pressure in porous sandstone is generally <0.1 bar (approximately 0.01 MPa) but may increase to tens of bars in source rocks with clay grain size (Ingebrit-
  • 28. sen et al. 2006). Capillarity model parameters were chosen within the range of typical rock types obtained from other publications (Levorsen 1967; Wendebourg 1994; Bloom- field et al. 2001). Typically, the capillary pressure of more permeable formations exhibit lower values, but sharp increases occur near the irreducible water saturation point (Swr). The sum of water and petroleum relative permeabil- ity is usually less than one when both phases are mobile. Initially, hydrostatic conditions and conductive thermal profile were assumed throughout the basin. The levels of overpressure were chosen considering that the pore pres- sure in the Southern California faults may approach
  • 29. lithostatic pressure due to mechanical compaction (Gra- tier et al. 2002, 2003). The values of overpressure ratio, fault permeability, and the period of episodic flow pulses are presented in Table 2. The over pressure ratio (k*) in a sedimentary basin can be defined as follows (Wang et al. 2006) Fig. 6. Model numerical grid, boundary conditions and hydrostratigraphy of the Newport-Inglewood fault zone based on the cross section along the transect X–X’ (Fig. 1b). The upper margin is the prescribed pressure head of 200 m and the isothermal boundary condition of 4°C. The left and right margins of the grid were assigned to be hydrostatic for pressure and ther-
  • 30. mally insulated (no flow) for heat. The bottom margin were assigned as no flow and a constant temperature of 160°C. Overpressure is applied to the right end of the fault boundary (marked as a white arrow) as a prescribed pressure condition, and the petroleum saturation at this boundary is con- stant (Sn = 0.6). © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 Effects of episodic fluid flow 239 k� ¼ P � PF PL � PF � 100 ð9Þ
  • 31. where P is pore fluid pressure in the formation, PF is hydrostatic pressure, and PL is lithostatic pressure. Porosity of the fault also changes dynamically in the range of 0.1– 0.3, accompanying with the permeability variation. In middle Miocene to early Pleistocene, the basin is still under a shallow marine environment (Wright 1991), so the prescribed pressure head of 200 m and the isothermal boundary condition of 4°C were assigned to the top boundary of the grid. The temperature value was chosen within the ranges of estimated Miocene shallow seawater temperature (0–6°C) (Zachos et al. 2001). Boundary con- ditions along the bottom margin of the grid were assigned as no fluid flow (impermeable) and a constant
  • 32. temperature of 160°C, based on geothermal gradients reported in this area of 35–40°C km�1 (Higgins & Chap- man 1984; Jeffrey et al. 1991). The left and right mar- gins of the grid were assigned to be hydrostatic for pressure and thermally insulated for heat. Petroleum was injected through the fault boundary at the right boundary (white arrow in Fig. 6), and this condition is physically possible only when we assume that most of petroleum was generated in the deep basin center and migrated through the fault zone. CONTINUOUS FLOW MODEL First, we considered a continuous hydrocarbon migration scenario through the Newport-Inglewood fault zone, assuming a slightly overpressured sedimentary basin envi-
  • 33. ronment with constant fault zone permeability. The level of overpressure at the fault boundary (marked as a white arrow in Fig. 6) was held constant during the simulation, and the fault zone was considered as a channel for releas- ing overpressure of the basin center as well as a conduit for petroleum migration. For Model 1, the vertical fault per- meability kz was set to 10 md (10 �14 m2), which is within the range of kx for the Pliocene unit and higher than that of Upper Miocene unit. The permeability value was inferred from the fact that a fault zone may be more per- meable than its host rock, especially when the fault was ruptured by shear stresses and brecciated to include sur-
  • 34. rounding rocks (Aydin 2000). The change of fault perme- ability during deformation depends on fault rock clay content and burial history. Laboratory experimental data show permeability increasing over 2–4 orders of magnitude (Fisher & Knipe 2001). Fluid pressure gradients in a compacting sedimentary basin usually vary between hydrostatic (approximately 10.1 MPa km�1) and lithostatic pressure (approximately 23.3 MPa km�1), and the values used in the simulations Table 1 Hydrogeologic model parameters. Parameter Topanga Puente Repetto Pico San Pedro NIFZ kx, Horizontal permeability (md) 3.2 0.13 18 50 50 1–1000* kx/kz, Anisotropy 100 100 100 100 100 0.1 /, Porosity 0.16 0.17 0.23 0.33 0.30 0.20
  • 35. ct, Bulk compressibility (Pa �1) 1 9 10�10 3 9 10�9 1 9 10�10 1 9 10�10 1 9 10�10 3 9 10�9 e, Thermal dispersivity (m) 100 100 100 100 100 100 k, Bulk thermal conductivity (W m�1°C) 3.0 2.5 2.7 2.8 2.8 3.0 c, Specific heat of matrix (J kg�1 °C) 750 750 750 750 750 750 a, van Genuchten coefficient (m�1) 3.0 9 10�4 4.0 9 10�4 2.5 9 10�3 2.0 9 10�3 2.0 9 10�3 1.0 9 10�2 n, van Genuchten coefficient 5.0 4.0 6.0 4.5 4.3 8.0 Swr, Irreducible wetting phase saturation (%) 10 18 12 12 12 12 Snr, Irreducible nonwetting phase saturation (%) 6 10 7 7 7 3 *Fault zone permeabilities are varied depending on model settings described in Table 2. Permeability and porosity values were selected within the ranges from a technical report written by the California Department of Conservation (1991) and unpublished data from Plains Exploration and Production Company. 1 md = 10�15 m2. Capillarity model parameters are obtained from publications (see the text). Table 2 Petroleum migration scenarios for the Newport-
  • 36. Inglewood fault zone. Model Scenario Overpressure ratio (%) Fault permeability (md)* Period of fluid pulse (years) Petroleum saturation 1 Continuous 10 10 0.6 2 Continuous 10 5 0.6 3 Continuous 10 20 0.6 4 Continuous 10 50 0.6 5 Episodic 10–80 1–1000 3000 0.6 6 Episodic 10–80 1–1000 2000 0.6 7 Episodic 10–80 1–1000 1000 0.6 8 Episodic 10–80 1–1000 500 0.6 *1 md = 10�15 m2. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 240 B. JUNG et al. were selected based on the current LA basin fluid pressure
  • 37. gradient of 9.9–12.6 MPa km�1 (Berry 1973; McCulloh 1979). These values can be converted to �1.5 to 19.0% in overpressure ratio. These relatively low-pressure gradients, despite the high subsidence and sedimentation rates, seem to originate from the high content of coarse-grained sedi- ments that can dissipate pressure very effectively in the LA basin (Hayba & Bethke 1995). We assumed the whole domain was initially hydrostatic, and then, the lower right end of the fault on the boundary of the mesh (marked by the white arrow in Fig. 6) was set to 10% overpressure. Petroleum saturation at the inlet of the fault zone was maintained as a constant value of 0.6, which is taken from the average value of Miocene strata in the Inglewood oil field (California Department of Conser-
  • 38. vation 1992). Figure 7 illustrates the pore pressure change in terms of hydraulic head (h) of the continuous flow model, assuming a constant fault permeability of 10 md (10�14 m2) and overpressure ratio of 10% during the whole simulation time (Model 1). The overpressured areas are sustained in the lower part of the fault zone and the Miocene formations, but the Pliocene and Pleistocene formations remain largely hydrostatic (colored in white). The high pore fluid pressure change is shown in the lower part of the fault. High fluid pressure in the fault zone is confined in the Upper Mio- cene formation due to its low permeability. The pressure pattern reaches the equilibrium state approximately after
  • 39. 40 000 years. Subsurface temperature under the continu- ous flow model displays dominantly conductive transport patterns with nearly horizontal isotherms, except for a slight elevation along the lower part of the fault zone (Fig. 8A). Elevated temperatures are the result of upward fluid flow into the fault zone, and contrast in thermal con- ductivity. The petroleum velocity distribution of the domain (Fig. 8B) shows that high velocity (approximately 0.2 m year�1) area appears at the lower part of the fault zone, and it gradually decreases moving up along the fault. The petroleum velocity at the upper part of the fault is <0.03 m year�1. Petroleum migrates mainly through the fault zone, and
  • 40. the upper Miocene formation provides an effective seal for the migration (Fig. 9A). Highly saturated areas (warmer color) are found in the boundaries between the fault zone and the upper Miocene formation because petroleum dis- charge away from the fault is retarded by the permeability contrast. Petroleum continues to rise due to buoyancy and saturates the tributary fault structures such as the Sentous fault before smearing into the Pliocene formation near the fault zone, and eventually vents on the sea floor (Fig. 9C, D). EFFECT OF FAULT PERMEABILITY Effects of fault zone permeability on petroleum distribu- tion were investigated by comparing four continuous
  • 41. flow models (Models 1–4). The parameters assumed are tabulated in Table 2. For each model scenario, the per- meability and pore pressure in the fault zone are assumed to be constant. We also compared total accumulation time required to reach the current estimate of total oil reserve in the Inglewood oil field: approximately 450 million barrels of oil equivalent (MMBOE = 106 barrels) (California Department of Conservation 1992). The total petroleum volume in a unit of barrel of oil equivalent (BOE), computed to provide a base of comparison, was obtained by multiplying petroleum saturation, porosity, and characteristic length (1 km) for considering the vol- ume as a three-dimensional domain. The total petroleum
  • 42. volume within the domain can be calculated using petro- leum saturation, porosity, finite elemental area, and char- acteristic length: PV ¼ XN i¼1 Sn/Ai ! � Lc ð10Þ where PV is the total petroleum volume at a specified time step, N the number of finite elements, Sn the petroleum saturation, / is porosity, Ai is the area of the i-th finite ele- ment, and Lc the characteristic length. The temporal variation of total petroleum volume of
  • 43. four continuous models are depicted for showing the rela- tionship of fault permeability and flow simulation time required for the total volume of 450 million barrels Fig. 7. Hydraulic head change (Dh = h – h0, h0 is hydrostatic pressure) of the Newport-Inglewood fault zone applied with the continuous petroleum migration model having constant fault zone permeability of 10 md (10�14 m2) (Model 1) at 100 000 years after starting simulation. Overpres- sure appears along the fault zone. The domain was initially hydrostatic and then the 10% overpressure was applied at the lower end of the fault boundary. Numbers are in meter.
  • 44. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 Effects of episodic fluid flow 241 (Fig. 10). Not surprisingly, the results indicate that higher fault permeability results in faster accumulation. For exam- ple, the reservoir takes about 62 000 years to fill when the fault permeability is 10 md (10�14 m2), but this time per- iod reduces to about 34 000 years when the permeability increases to 50 md. It is also worth noting that the filling time increases with decreasing fault permeability almost in linear proportion. The spatial patterns of petroleum satura- tion are also affected significantly by fault permeability. Figure 11 shows petroleum saturations for Models 1–4
  • 45. when the total volume is equal to 450 million barrels. In the high fault permeability model (Model 4, Fig. 11D), petroleum migrates more vigorously laterally to invade adjacent formations through fingering and channeling (A) (B) Fig. 8. (A) Subsurface temperature (°C) and (B) petroleum velocity of the fault zone (Model 1) at 100 000 years after starting simulation. The top margin of the grid was set as the isothermal boundary condition of 4°C, and the bottom margin were assigned as a constant tempera- ture of 160°C, based on geothermal gradients reported in this area of 35–40°C km�1. The left
  • 46. and right sides were assumed to be thermally insulated for heat. (A) (B) (C) (D) Fig. 9. Evolution of petroleum saturation for Model 1 at (A) 40 000 years, (B) 60 000 years, (C) 80 000 years, and (D) 100 000 years of simulation time showing the advancement of hydrocarbons through the fault zone. Petroleum invades into the upper Miocene units because of the relatively high fluid pressure in the lower part of the fault. The outline of this magnified area is marked with dashed lines in Fig. 6.
  • 47. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 242 B. JUNG et al. effects and accumulates significantly enhanced pools in the Pliocene sediments compared with that of low permeabil- ity cases. In the low permeability model (Fig. 11A, Model 2), however, accumulation tends to be more abundant in the upper Miocene sediments and less in the Pliocene unit. Fig. 10. Effect of fault permeability on flow duration needed to accumulate the current known volume of petroleum in the Inglewood
  • 48. oil field (1 MMBOE = 106 barrels). Flow dura- tion assumes that 20% of the total fault length in transverse is available for petroleum migra- tion and therefore characteristic or effective flow length of about 1 km. (A) (B) (C) (D) Fig. 11. Comparison of continuous flow models showing the effects of fault permeability on petroleum migration and distribution, indicated by saturation. Numbers in parenthesis indicate times needed for accumulating total oil reverse in Inglewood field (see Fig. 10).
  • 49. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 Effects of episodic fluid flow 243 EPISODIC FLOW MODEL Models 5 through 8 were designed to investigate the effects of episodic fluid flow on petroleum migration, based on a fault-valve mechanism (Wright 1991), so a repeating pattern of changing fault permeability and pore fluid pressure was assigned (Fig. 12). A sharp increase in permeability from k = 1 to 1000 md (10�15 to 10�12 m2) denotes fault rupturing caused by tectonic stresses or due to hydrofracturing by fluid pressure accumulation (Fisher & Knipe 2001). Elevated fault permeability and accompa-
  • 50. nied focused fluid flow are, however, considered to be short-lived (<100 years) events (Sibson 1994), so the fault will be closed hydromechanically as the pore pressure decreases and will be assumed to sealed by mineral precipi- tation (Dieterich & Kilgore 1996; Scholz 2002; Johnson & Jia 2005). Recent field measurements of time-varying permeability in faults after large earthquakes also suggest that transient fluid pulses play a major role for water circu- lation in deep fault zones (Xue et al. 2013). To simulate these sealing effects, the fault zone perme- ability was decreased from 1000 to 10 md during the first 100 years after rupturing, and then, it continued to be reduced to 1 md until the next pulse occurs 2000 years
  • 51. later. We assumed that the hydromechanical compaction causing the first stage of permeability drop is abrupt and much faster than the second stage of drop dominated by hydrothermal precipitation. We also assume that the fault ruptured at overpressure ratio of 80%, and pore fluid pressure dropped to 10% over- pressure level for the episodic flow model. The fluid pressure gradually builds back up to 80% during the sealing period. Qualitative behavior of these fluctuating pore pressure pat- terns was first suggested by Sibson (see review by Roberts & Nunn 1995), and the pressure value at the fault rupture was chosen from work by Gratier et al. (2003) and Appold et al.
  • 52. (2007). Petrologic evidence of multiple pulses of petroleum flow in fault zones is reported in publications that deal with Neogene basins in southern California (Boles & Ramseyer 1987; Eichhubl & Boles 2000; Perez & Boles 2007), but there are a lack of articles providing a quantitative estimation of possible flow duration. In this study, the period of epi- sodic flow pulse was chosen within the range of 500– 2000 years based on Myers’s research on the Puente Hill blind fault in the LA basin (Myers et al. 2003), where he estimated the recurrence period of earthquakes in this fault zone to be approximately 1700–3200 years, assuming a Richter magnitude of M = 6.0–7.5. We assume that earth- quakes in that range could impose significant hydromechan-
  • 53. ical impacts on the adjacent aquifers. Fig. 12. Fault permeability and pore fluid pres- sure settings for the episodic flow model (Model 6). Fault permeability changes applied to the whole fault zone, and pore pressure changes applied to the end of the fault on the boundary elements (marked by the white arrow in Fig. 6) © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 244 B. JUNG et al. When the fault zone is sealed, pore fluid pressure builds up in the lower part of the fault so it cannot reach the upper formations due to the low fault permeability (Fig. 13A).
  • 54. After the fault is open, previously accumulated high pressure is released, and a temporary focus of fluid flow is created in the fault zone (Fig. 13B). The released overpressure per- vades the Miocene formations and the center of the anticline structure. This strong and episodic pressure anomaly may act as a strong transient driving force for both groundwater and petroleum. Furthermore, a subsurface heat anomaly is also detected along the fault zone due to the increased flow rate and heat advection (Fig. 14). Temperature elevation in the fault zone can facilitate the petroleum migration by decreasing viscosity of the fluid. However, the heat anomaly disappears when the fault is sealed, and isotherms become
  • 55. horizontal and conductive. The pattern of petroleum migration in the early stages is similar to that of continuous migration, which mostly fol- lows the fault surface (Fig. 15A), but some distinctive and different patterns appear in later stages. In the episodic model, petroleum migration seems to be more dispersed than in the continuous model and accumulates in broader areas in the middle of the shallow Pliocene sediments. Petroleum migrates from the fault structures themselves and extends into the adjacent sedimentary formations. Effective spreading of petroleum to the adjacent rocks is due to the high permeability, Darcy flow rates, and fluid pressure concentrated in this area during the episodic flow
  • 56. events. Especially, the high fluid pressure facilitates petro- leum to overcome entry pressure of reservoir rocks sur- rounding the fault zone. EFFECT OF EPISODIC FLOW FREQUENCY To further investigate the effect of frequency of episodic fluid flow on petroleum accumulation, the temporal varia- tion of total petroleum volumes for Models 5 to 8 can be compared graphically (Fig. 16). The total volume of petro- (A) (B) Fig. 13. Hydraulic head change (Δh = h – h0) for the episodic flow model (Model 6): (A) just before the fault opening due to fault rupture, and (B) 50 years after the fault opening. h0 is
  • 57. the hydrostatic pressure. Numbers are in meter. (A) (B) Fig. 14. Subsurface temperature for the epi- sodic flow model (Model 6): (A) before the fault opening, and (B) 50 years after the fault open. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 Effects of episodic fluid flow 245 leum increases with flow frequency. If we set the known 450 million barrels as the comparison base, the flow model with a shorter period of episodic flow pulses reaches the base faster than the model with a longer period. For exam-
  • 58. ple, it takes approximately 5700 years if the period of epi- sodic pulse is 500 years (Model 8) but takes more than 24 100 years if the period is 3000 years (Model 5). Petro- leum distributions of Models 5 to 8 at the comparison base were compared in Fig. 17. The overall patterns of petroleum distributions from four models look similar. (A) (B) (C) (D) Fig. 15. Evolution of petroleum saturation for the episodic flow model (Model 6) at four time steps. Fig. 16. Comparison of total petroleum volume changes of episodic flow models, showing the
  • 59. effect of fault permeability on recharging (fill- ing) times needed to accumulate the observed volume of total oil reserve in the Inglewood oil field. Filling times (durations) assume that only 20% of the fault zone is available for petroleum migration along the length of the oil field. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 246 B. JUNG et al. They all caused significant accumulations in the Pliocene unit and channeled flow in the middle Miocene unit, but the shorter period scenario (Model 8, Fig. 17D) enables
  • 60. more pervasive accumulations in the center of the Pliocene unit, in the Sentous fault, and in the upper part of the Newport-Inglewood fault zone. It also caused relatively smaller intrusion into the Middle Miocene unit, which probably means that the episodic migration having a shorter period is more active in transferring petroleum upward because the fault zone is open more frequently and assumed to be sealed for less time. The episodic flow model having the fault k range of 1– 1000 md (10�15 to 10�12 m2) and a pulse interval of 3000 years (Model 5) can be compared with a continuous model (Model 4) having k values of 50 md to demonstrate differences between continuous and episodic flow models.
  • 61. These two cases were selected because their simulation times required for fill the reservoir were most similar although the pattern of the driving force is different. The episodic model (Model 5) takes 24 100 years to accumu- late the petroleum at the level of comparison point (approximately 450 million barrels). This value is faster than the elapsed time of the continuous model, which is 34 000 years, having fault permeability of 50 md. When we consider the relatively short sum of open fault duration (<1400 years) in Model 5, it could be said that the epi- sodic flow can drive migration of petroleum more effi- ciently compared to the continuous models with constant fault permeability higher than 50 md.
  • 62. For the given amount of total petroleum volume, epi- sodic fault pressure-driven migration generates broader and more highly saturated petroleum accumulations in the Pli- ocene sediments and also showed more dispersed patterns in the Middle Miocene unit. On the other hand, the con- tinuous migration made relatively low-saturated accumula- tions simply following the fault structures. Petroleum in the continuous model also tends to stay in the lower part of the fault zone and rise slowly by the forces of buoyancy. CONCLUSIONS To understand the dynamics of petroleum migration in active margin basins, both continuous and episodic fluid migration models were constructed for the Newport-Ingle-
  • 63. wood fault zone in the Los Angeles basin. The effects of time-varying fault permeability on hydrocarbon accumula- tion rate and pattern are also quantitatively investigated using multiphase numerical modeling. In continuous flow models, the petroleum migration rate in the fault zone ranges 0.03–0.2 m year�1, and the (A) (B) (C) (D) Fig. 17. Comparison of episodic flow models showing the effects of flow frequency on petro- leum migration and distribution. Numbers in parenthesis indicate times needed for accumu-
  • 64. lating total oil reverse in Inglewood field (see Fig. 16). © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 Effects of episodic fluid flow 247 duration of time needed to form the Inglewood oil field (approximately 450 million barrels) is approximately 34 000–62 000 years depending on the fault permeability assigned varying 50–5 md. High fault permeability also allows more petroleum accumulation and channeling in the Pliocene formation because larger overpressures can propa- gate further and reach at the upper part of the fault. In episodic models, the total oil volume in the reservoir
  • 65. can be accumulated in about 24 000 years under assump- tions that fault permeability fluctuates between 1–1000 md and seismically induced flow pulse occurs every 2000 years. Episodic migration forms larger areas of oil accumulation with highly petroleum saturation accumulations in a broader region surrounding the fault zone. Considering the observed pattern of hydrocarbon accumulation in this region, the episodic model could be more geologically meaningful than continuous migration model. The frequency of episodic flow affects petroleum migra- tion and distribution patterns in the fault zone. Faster and broader accumulations appear in the Pliocene sediments if
  • 66. the pulse period becomes shorter. The frequent episodic fluid pulse also forms accumulations with higher petroleum saturation and flow channels in adjacent formations. ACKNOWLEDGEMENTS This research was supported by a grant from the U.S. Department of Energy – Basic Energy Sciences (Grant number: DE-FG02-96ER14620), The GDL Foundation also provided a Ph.D. scholarship to the first author, for which he is extremely grateful. REFERENCES Appold MS, Garven G (2000) Reactive flow models of ore formation in the Southeast Missouri district. Economic Geology, 95, 1605–26.
  • 67. Appold MS, Garven G, Boles JR, Eichhubl P (2007) Numerical modeling of the origin of calcite mineralization in the Refugio- Carneros fault, Santa Barbara Basin, California. Geofluids, 7, 79– 95. Aydin A (2000) Fractures, faults, and hydrocarbon entrapment, migration and flow. Marine and Petroleum Geology, 17, 797– 814. Bear J (1972) Dynamics of Fluids in Porous Media. Dover Publications Inc., New York. de Berg M, Cheong O, van Kreveld M, Overmars M (2008) Computational Geometry, 3rd edn. Springer-Verlag, Berlin, 386 p. Berry FAF (1973) High fluid potentials in California Coast Ranges and their tectonic significance. American Association of
  • 68. Petroleum Geologists Bulletin, 57, 1219–45. Biddle KT (1991) The Los Angeles basin: an overview. In: Active Margin Basins (ed. Biddle KT), pp. 1–24. AAPG Memoir 52, Tulsa, OK. Blackwell DD, Steele JL (1989) Thermal conductivity of sedimentary rocks: measurement and significance. In: Thermal History of Sedimentary Basins: Methods and Case Histories (eds Naeser ND, McCulloh TH), pp. 13–36. Springer-Verlag, New York. Blake GH (1991) Review of the Neogene biostratigraphy and stratigraphy of the Los Angeles Basin and implications for basin evolution. In: The Los Angeles Basin: An Overview (ed. Biddle KT), pp. 135–84. AAPG Memoir 52, Tulsa, OK. Blanpied ML, Lockner DA, Byerlee JD (1992) An earthquake mechanism based on rapid sealing of faults. Nature, 358, 574–6.
  • 69. Bloomfield JP, Gooddy DC, Bright MI, Williams PJ (2001) Pore- throat size distributions in Permo-Triassic sandstones from the United Kingdom and some implications for contaminant hydrogeology. Hydrogeology Journal, 9, 219–30. Boles JR, Ramseyer K (1987) Diagenetic carbonate in Miocene sandstone reservoir, San Joaquin basin, California. American Association of Petroleum Geologists Bulletin, 71, 1475–87. Boles JR, Eichhubl P, Garven G, Chen J (2004) Evolution of a hydrocarbon migration pathway along basin-bounding faults: evidence from fault cement. American Association of Petroleum Geologists Bulletin, 88, 947–70. Bradley JS (1975) Abnormal formation pressure. American Association of Petroleum Geologists Bulletin, 59, 957–73. Caine JS, Bruhn RL, Forster CB (2010) Internal structure, fault rocks, and inferences regarding deformation, fluid flow, and
  • 70. mineralization in the seismogenic Stillwater normal fault, Dixie Valley, Nevada. Journal of Structural Geology, 32, 1576–89. California Department of Conservation (1992) California Oil and Gas Fields: Volume II – Southern, Central Coastal, and Offshore California Oil and Gas Fields. California Department of Conservation, Division of Oil, Gas, and Geothermal Resources, Sacramento, CA. Class H, Helmig R, Bastian P (2002) Numerical simulation of non-isothermal multiphase multicomponent processes in porous media. 1. An efficient solution technique. Advances in Water Resources, 25, 533–50. Dieterich JH, Kilgore BD (1996) Imaging surface contacts: power law contact distributions and contact stresses in quartz, calcite, glass and acrylic plastic. Tectonophysics, 256, 219–39.
  • 71. Eichhubl P, Boles JR (2000) Focused fluid flow along faults in the Monterey Formation, coastal California. Geological Society of America Bulletin, 112, 1667–79. England WA, Mackenzie AS, Mann DM, Quigley TM (1987) The movement and entrapment of petroleum fluids in the subsurface. Journal of the Geological Society, London, 144, 327– 47. Evans B (1992) Greasing the fault. Nature, 358, 544–5. Faulkner DR, Jackson CAL, Lunn RJ, Schlische RW, Shipton ZK, Wibberley CAJ, Withjack MO (2010) A review of recent developments concerning the structure, mechanics and fluid flow properties of fault zones. Journal of Structural Geology, 32, 1557–75.
  • 72. Fisher QJ, Knipe RJ (2001) The permeability of faults within siliciclastic petroleum reservoirs of the North Sea and Norwegian Continental Shelf. Marine and Petroleum Geology, 18, 1063–81. Fisher QJ, Casey M, Harris SD, Knipe RJ (2003) Fluid-flow properties of faults in sandstone: the importance of temperature history. Geology, 31, 965–8. Garden IR, Guscott SC, Burley SD, Foxford KA, Walsh JJ, Marshall J (2001) An exhumed palaeo-hydrocarbon migration fairway in a faulted carrier system, Entrada Sandstone of SE Utah, USA. Geofluids, 1, 195–213. Geiger S, Burri A, Roberts S, Matthai SK, Zoppou C (2004) Combining finite element and finite volume methods for © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 248 B. JUNG et al.
  • 73. efficient multiphase flow simulations in highly heterogeneous and structurally complex geologic media. Geofluids, 4, 284–99. Geiger S, Driesner T, Heinrich CA, Matthai SK (2006) Multiphase thermohaline convection in the earth’s crust: I. A new finite element – finite volume solution technique combined with a new equation of state for NaCl-H2O. Transport in Porous Media, 63, 399–434. Gelhar LW, Welty C, Rehfeldt KR (1992) A critical review of data on field-scale dispersion in aquifers. Water Resources Research, 28, 1955–74. van Genuchten MT (1980) A closed-form equation for predicting the hydraulic conductivity of unsaturated soils. Soil Science Society of America Journal, 44, 892–8. Glasø Ø (1980) Generalized pressure-volume-temperature
  • 74. correlations. Journal of Petroleum Technology, 32, 785–95. Gong ZS, Huang LF, Chen PH (2011) Neotectonic controls on petroleum accumulations, offshore China. Journal of Petroleum Geology, 34, 5–28. Gratier JP, Favreau P, Renard F (2002) Fluid pressure evolution during the earthquake cycle controlled by fluid flow and pressure solution crack sealing. Earth, Planets and Space, 54, 1139–46. Gratier JP, Favreau P, Renard F (2003) Modeling fluid transfer along California faults when integrating pressure solution crack sealing and compaction processes. Journal of Geophysical Research, 108, 1–25. Guo X, He S, Liu K, Cao F, Shi H, Zhu J (2011) Condensates in the PY30-1 structure, Panyu Uplift, Pearl River Mouth Basin, South China Sea: evidence for hydrothermal activity associated
  • 75. with petroleum migration and accumulation. Journal of Petroleum Geology, 34, 217–32. Harten A (1997) High resolution schemes for hyperbolic conservation laws. Journal of Computational Physics, 135, 260– 78. Hayba DO, Bethke CM (1995) Timing and velocity of petroleum migration in the Los Angeles Basin. Journal of Geology, 103, 33–49. Helmig R (1997) Multiphase Flow and Transport Processes in the Subsurface: A Contribution to the Modeling of Hydrosystems. Springer, Berlin. 367 pp. Hesson HB, Olilang HR (1990) Seal Beach Oil Field: Alamitos and Marine Areas TR39. Division of Oil and Gas, California Department of Conservation, Sacramento, CA. Higgins CT, Chapman RH (1984) Geothermal energy at Long Beach Naval Shipyard and Naval Station and at Seal Beach Naval Weapons Station, California. DMG Open-File Report
  • 76. 84-32, California Department of Conservation, Division of Mines and Geology, Sacramento, CA. Huber R, Helmig R (2000) Node-centered finite volume discretizations for the numerical simulation of multiphase flow in heterogeneous porous media. Computational Geosciences, 4, 141–64. Ingebritsen SE, Neuzil CE, Sanford WE (2006) Groundwater in Geologic Processes, 2nd edn. Cambridge University Press, Cambridge. 536 pp. Jeffrey AWA, Alimi HM, Jenden PD (1991) Geochemistry of Los Angeles basin and gas system. In: Active Margin Basins (ed. Biddle KT), pp. 197–220. AAPG Memoir 52, Tulsa, OK. Johnson PA, Jia X (2005) Nonlinear dynamics, granular media and dynamic earthquake triggering. Nature, 437, 871–4. Kaplan IR, Alimi MH, Hein C, Jeffrey A, Lafferty MR,
  • 77. Mankiewicz MP (2000) The geochemistry of hydrocarbons and potential source rocks from the Los Angeles and Ventura basins, data synthesis and text, v.I. In: Collection of Papers Written in the Mid-to-Late 1980’s and in 1997 by Staff Members of Global Geochemistry Corporation about the Oil, Gas, and Source Rock Investigations Carried Out in the San Joaquin, Santa Maria, Santa Barbara, Ventura, and Los Angeles Basins, California (ed.Kaplan IR), pp. 1–238. Pacific Section – AAPG, Bakersfield, CA. Karlsen DA, Skeie JE (2006) Petroleum migration, faults and overpressure, Part I: calibrating basin modelling using petroleum in traps – a review. Journal of Petroleum Geology, 29, 227–55. Kroeger KF, di Primio R, Horsfield B (2009) Hydrocarbon flow modeling in complex structures (Mackenzie Basin, Canada). American Association of Petroleum Geologists Bulletin, 93, 1209–34.
  • 78. Levorsen AI (1967) Geology of Petroleum, 2nd edn. W. H. Freeman, San Francisco, CA. Mandl G, Harkness RM (1987) Hydrocarbon migration by hydraulic fracturing. In: Deformation of Sediments and Sedimentary Rocks 29 (eds Jones ME, Preston RMF), pp. 39– 53. Geological Society of London Special Publication, London. de Marsily G (1986) Quantitative Hydrogeology: Groundwater Hydrology for Engineers. Academic Press, Orlando, FL, 440 p. McCulloh TH (1979) Implication for petroleum appraisal. In: Geologic Studies of the Point Conception Deep Stratigraphic Test Well OCS-CAL 78-164 No. 1 Outer Continental Shelf Southern California, United States (ed. Cook HE), pp. 26–42. USGS Open File Report 79-1218, U.S. Geological Survey, Menlo Park, CA. Myers DJ, Nabelek JL, Yeats RS (2003) Dislocation modeling of
  • 79. blind thrusts in the eastern Los Angeles basin, California. Journal of Geophysical Research-Solid Earth, 108, ESE 14, 1– 19. Nishikawa T, Siade AJ, Reichard EG, Ponti DJ, Canales AG, Johnson TA (2009) Stratigraphic controls on seawater intrusion and implications for groundwater management, Dominguez Gap area of Los Angeles, California, USA. Hydrogeology Journal, 17, 1699–725. Olson L (1978) Shallow Aquifers and Surface Casing Requirements for Wilmington and Belmont Offshore Oil Fields TR22. California Department of Conservation, Division of Mines and Geology, Sacramento, CA. Perez RJ, Boles JR (2007) Mineralization, fluid flow, and sealing properties associated with an active thrust fault: San Joaquin basin, California. American Association of Petroleum
  • 80. Geologists Bulletin, 88, 1295–314. Phillips SL, Ingbene A, Fair JA, Ozbek H, Tavana M (1981) A technical data book for geothermal energy utilization. DE81- 029868. Phillips SL, Ozbek H, Silvester LF (1983) Density of sodium chloride solutions at high temperatures and pressures. DE84- 004883. Reichenberger V, Jakobs H, Bastian P, Helmig R (2006) A mixed-dimensional finite volume method for two-phase flow in fractured porous media. Advances in Water Resources, 29, 1020–36. Reynolds SJ, Lister GS (1987) Structural aspects of fluid-rock interactions in detachment zones. Geology, 15, 362–6. Roberts SJ, Nunn JA (1995) Episodic fluid expulsion from
  • 81. geopressured sediments. Marine and Petroleum Geology, 12, 195–204. Sabins EF (1997) Remote Sensing: Principles and Interpretation. Wiley & Sons, New York. 494 pp. Scholz CH (2002) The Mechanics of Earthquakes and Faulting. 2nd edn. Cambridge University Press, Cambridge. pp. 471. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 Effects of episodic fluid flow 249 Shaw JH, Suppe J (1996) Earthquake hazards of active blind- thrust faults under the central Los Angeles basin, California. Journal of Geophysical Research, 101, 8623–42. Sibson RH (1994) Crustal stress, faulting and fluid flow. In: Geofluids: Origin, Migration, and Evolution of Fluids in Sedimentary Basins (ed. Parnell J) 78, pp. 69–84. Geological
  • 82. Society Special Publication, London. Sorkhabi R, Tsuji Y (2005) The place of faults in petroleum traps. In: Faults, Fluid Flow, and Petroleum Traps (eds Sorkahabi R, Tsuji Y), AAPG Memoir 52, Tulsa, OK. Sperrevik S, Gillespie PA, Fisher QJ, Halvorsen T, Knipe RJ (2002) Empirical estimation of fault rock properties. In: Hydrocarbon Seal Quantificaiton (eds Koestler AG, Hunsdale R) 11, pp. 109– 25. Norwegian Petroleum Society Special Publication, Stavanger, Norway. Walder J, Nur A (1984) Porosity reduction and crystal pore pressure development. Journal of Geophysical Research, 89, 11539–48. Wang K, He J, Hu Y (2006) A note on pore fluid pressure ratios in the Coulomb wedge theory. Geophysical Research Letters, 33,
  • 83. L19310. doi:10.1029/2006GL027233. Watson JT, Basu RS, Sangers JV (1980) An improved representative equation for the dynamic viscosity of water substance. Journal of Physical and Chemical Reference Data, 9, 1255–90. Wendebourg J (1994) Simulating hydrocarbon migration and stratigraphic traps, PhD thesis. Standford University. Wright TL (1991) Structural geology and tectonic evolution of the Los Angeles Basin, California. In: Active Margin Basins (ed. Biddle KT), pp. 35–134. AAPG Memoir 52, Tulsa, OK. Xue L, Li HB, Brodsky EE, Xu ZQ, Kano Y, Wang H, Mori JJ, Si JL, Pei JL, Zhang W, Yang G, Sun ZM, Huang Y (2013) Continuous permeability measurements record healing inside the Wenchuan earthquake fault zone. Science, 304, 1555–9.
  • 84. Yamaguchi A, Cox SF, Kimura G, Okamoto S (2011) Dynamic changes in fluid redox state associated with episodic fault rupture along a megasplay fault in a subduction zone. Earth and Planetary Science Letters, 302, 369–77. Yerkes RF (1972) Geology and oil resources of the western Puente Hills area, southern California. USGS Professional Paper, 420-C, 63. Zachos JC, Pagani M, Lisa S, Thomas E, Billups K (2001) Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686–93. Zhang Y, Gable CW, Zyvoloski GA, Walter LM (2009) Hydrogeochemistry and gas compositions of the Uinta Basin: a regional-scale overview. American Association of Petroleum Geologists Bulletin, 93, 1087–118. © 2013 John Wiley & Sons Ltd, Geofluids, 14, 234–250 250 B. JUNG et al.
  • 85. Volume 14, Number 2, May 2014 ISSN 1468-8115 Geofluids This journal is available online at Wiley Online Library. Visit onlinelibrary.wiley.com to search the articles and register for table of contents and e-mail alerts. Geofluids is abstracted/indexed in Chemical Abstracts CONTENTS 127 Review of ‘Too Hot To Touch: The Problem of High-Level Nuclear Waste’ by William M. Alley and Rosemarie Alley E.M. Kwicklis 128 Diffusion and kinetic control of weathering layer development D. Reeves and D.H. Rothman 143 Carbon dioxide controlled earthquake distribution pattern in
  • 86. the NW Bohemian swarm earthquake region, western Eger Rift, Czech Republic – gas migration in the crystalline basement F.H. Weinlich 160 Fractal analysis of veins in Permian carbonate rocks in the Lingtanchang anticline, western China B. Deng, S. Liu, L. Jansa, S. Yong and Z. Zhang 174 Fluid effect on hydraulic fracture propagation behavior: a comparison between water and supercritical CO2-like fluid X. Zhou and T.J. Burbey 189 Cementation and the hydromechanical behavior of siliciclastic aquifers and reservoirs D.F. Boutt, K.E. Plourde, J. Cook and L.B. Goodwin 200 Impacts of Pleistocene glacial loading on abnormal pore- water pressure in the eastern Michigan Basin O. Khader and K. Novakowski
  • 87. 221 Numerical simulation of mylonitization and structural controls on fluid flow and mineralization of the Hetai gold deposit, west Guangdong, China J. Zhu, Z. Li, G. Lin, Q. Zeng, Y. Zhou, J. Yi, G. Gong and G. Chen 234 Effects of episodic fluid flow on hydrocarbon migration in the Newport-Inglewood Fault Zone, Southern California B. Jung, G. Garven and J.R. Boles gfl_14_2_Issue toc_OC 4/11/2014 12:51 PM Page 1 Copyright of Geofluids is the property of Wiley-Blackwell and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission. However, users may print, download, or email articles for individual use.
  • 88. 747 Bulletin of the Seismological Society of America, Vol. 94, No. 2, pp. 747–752, April 2004 Activity of the Offshore Newport–Inglewood Rose Canyon Fault Zone, Coastal Southern California, from Relocated Microseismicity by Lisa B. Grant and Peter M. Shearer Abstract An offshore zone of faulting approximately 10 km from the southern California coast connects the seismically active strike-slip Newport–Inglewood fault zone in the Los Angeles metropolitan region with the active Rose Canyon fault zone in the San Diego area. Relatively little seismicity has been recorded along the off- shore Newport–Inglewood Rose Canyon fault zone, although it has long been sus- pected of being seismogenic. Active low-angle thrust faults and Quaternary folds
  • 89. have been imaged by seismic reflection profiling along the offshore fault zone, raising the question of whether a through-going, active strike-slip fault zone exists. We applied a waveform cross-correlation algorithm to identify clusters of microseis- micity consisting of similar events. Analysis of two clusters along the offshore fault zone shows that they are associated with nearly vertical, north- northwest-striking faults, consistent with an offshore extension of the Newport– Inglewood and Rose Canyon strike-slip fault zones. P-wave polarities from a 1981 event cluster are con- sistent with a right-lateral strike-slip focal mechanism solution. Introduction The Newport–Inglewood fault zone (NIFZ) was first identified as a significant threat to southern California resi- dents in 1933 when it generated the M 6.3 Long Beach earth- quake, killing 115 people and providing motivation for pas- sage of the first seismic safety legislation in the United States (Barrows, 1974; Hauksson and Gross, 1991; Yeats, 2001). The Rose Canyon fault (RCF) zone in the San Diego area
  • 90. has ruptured several times during the Holocene (Lindvall and Rockwell, 1995) and poses significant seismic hazard to San Diego area residents (Anderson et al., 1989). An off- shore zone of faulting approximately 10 km from the south- ern California coast connects the strike-slip NIFZ in the Los Angeles metropolitan region with the strike-slip RCF zone in the San Diego area. The activity and seismic potential of the intervening offshore fault zone, herein referred to as the offshore Newport–Inglewood Rose Canyon (ONI-RC) fault zone, has been the subject of debate for decades (e.g., Hill, 1971; Barrows, 1974; Fischer and Mills, 1991; Rivero et al., 2000; Grant and Rockwell, 2002). Recent attention has fo- cused on blind thrusts that may intersect the ONI-RC fault zone and accommodate some of the regional deformation (Grant et al., 1999; Rivero et al., 2000). Interaction with the thrust system could limit the magnitude of earthquakes on strike-slip faults in the ONI-RC fault zone, if they are active. Sparse offshore microseismicity has been difficult to locate accurately (Astiz and Shearer, 2000) but has been interpreted to be associated with an active ONI-RC fault zone (Fischer and Mills, 1991). We identified, relocated, and analyzed two clusters of microearthquakes within the northern and central ONI-RC fault zone to examine the fault structure, minimum depth of seismic activity, and source fault mechanism. The
  • 91. results suggest that the ONI-RC fault zone is a steeply dip- ping, active strike-slip fault to seismogenic depths. Tectonic and Geologic Setting In southern California, tectonic deformation between the Pacific and North American plates is accommodated pri- marily by a zone of strike-slip faults (Walls et al., 1998) with maximum deformation along the San Andreas system, decreasing westward into the offshore inner Continental Borderland region (Fig. 1). The inner Continental Border- land has complex structure including low-angle detachment faults, thrust faults, and high-angle strike-slip faults resulting from late Cenozoic rifting prior to the current transpressional tectonic regime (Crouch and Suppe, 1993; Bohannon and Geist, 1998). The geometry and slip rate of faults in the inner Continental Borderland are poorly constrained relative to onshore faults (Astiz and Shearer, 2000), yet they may pose significant seismic risk because they are close to populated areas, and several offshore faults appear to displace seafloor sediments (Legg, 1991). Grant and Rockwell (2002) pro- posed that an active �300-km-long Coastal Fault zone (Fig. 2) extends between the Los Angeles basin (USA) and coastal Baja California (Mexico), including the NIFZ and RCF zone in southern California, the Agua Blanca fault in
  • 92. 748 Short Notes Figure 1. Regional fault map (modified from Grant and Rockwell [2002]; after Walls et al. [1998]) summarizing regional tectonic deformation according to slip rate on major faults, including the San Andreas fault (SAF), San Jacinto fault zone (SJFZ), Elsinore fault zone (EFZ), Whittier fault (WF), Palos Verdes fault (PVF), NIFZ near Los Angeles (LA), RCF zone (RCFZ) near San Diego (SD), Agua Blanca fault zone (ABFZ), San Miguel fault zone (SMFZ), Imperial fault (IF), Cerro Prieto fault (CPF), and Laguna Salada fault (LSF). Offshore faults of the inner Continental Borderland include the Coronado Bank fault zone (CBF), San Diego Trough fault (SDTF), San Clemente fault zone (SCFZ), Santa Cruz Island fault (SCIF), and Santa Rosa Island fault (SRIF). The San Gabriel fault (SGF), San Cayetano fault (SCF), Oak Ridge fault (ORF), and Santa
  • 93. Ynez fault (SYF) are located in the Transverse Ranges. Baja California, and contiguous offshore fault zones. The structural style and slip rates of the NIFZ and RCF zone sug- gest that the intervening ONI-RC fault zone is an active, strike-slip fault zone with complex structure and a slip rate of 0.5–2.0 mm/yr. The NIFZ has been studied extensively in the Los An- geles basin by petroleum geologists. Hill (1971) proposed that the NIFZ overlies a major tectonic boundary separating eastern continental basement facies of granitic and associ- ated metamorphic rocks from oceanic Catalina schist facies of the inner Continental Borderland province to the west. The current strike-slip faulting regime apparently reactivated a structural weakness along the Mesozoic subduction zone (Hill, 1971) and initiated right-lateral motion in the mid-to- late Pliocene (Wright, 1991; Freeman et al., 1992). The NIFZ is a structurally complex series of discontinuous strike-slip faults with associated folds and shorter normal and reverse faults (Yeats, 1973), described by Wilcox et al. (1973) and Harding (1973) in a set of classic papers on wrench tecton- ics. Dextral displacement of late Miocene and younger sed- iments reveals a long-term slip rate of 0.5 mm/yr (Freeman
  • 94. et al., 1992). Multiple surface ruptures since the early Ho- locene indicate that the minimum slip rate is at least 0.3–0.6 mm/yr and may be substantially higher (Grant et al., 1997). A Holocene slip rate of �1 mm/yr has been assumed for seismic hazard assessment (SCECWG, 1995). Seismically active strands of the NIFZ have been mapped along the west- ern margin of the Los Angeles basin between the cities of Beverly Hills and Newport Beach (Bryant, 1988; Grant et al., 1997) where the fault steps over and continues offshore parallel to the coast (Morton and Miller, 1981). The subsurface structure and tectonic history of the RCF zone have not been studied as thoroughly as the NIFZ be- cause it lacks petroleum reserves. The Holocene sense of movement and slip rate of the RCF have been investigated for seismic hazard studies. Lindvall and Rockwell (1995) Short Notes 749 Figure 2. Map shows approximate location of major strands in the Coastal fault zone and dates of most recent rupture. The Coastal fault zone includes the northern
  • 95. NIFZ near Beverly Hills (BH) (dashed), southern NIFZ (bold) near Newport Beach (NB), the ONI-RC fault zone (gray) offshore of the San Joaquin Hills (SJH) and city of Ocean- side (O), the RCF zone (bold line) between La Jolla (LJ) and San Diego, the Coronado Bank fault zone (CBF), and the Agua Blanca fault zone (ABFZ, bold). Modified from Grant and Rockwell (2002); after Grant et al. (1997), Lindvall and Rockwell (1995), and Rockwell and Murbach (1999). and Rockwell and Murbach (1999) reported that it is pri- marily a strike-slip fault zone with a slip rate of 1–2 mm/yr, although individual strands display varying amounts of dip- slip motion in combination with strike slip. Location and Structural Models of the ONI-RC Fault Zone The location of the ONI-RC fault zone has been mapped by seismic reflection profiling in shallow water along the inner continental margin between Newport Beach and La Jolla, north of San Diego. Fischer and Mills (1991) sum- marized the results of eight investigations between 1972 and
  • 96. 1988, including unpublished mapping for the nuclear gen- erating station at San Onofre. Interpretations of Fischer and Mills (1991) and prior studies disagree about the location and number of traces, but all reveal a fairly continuous zone of faulting within a few kilometers of the coastline. Fischer and Mills (1991) presented several interpretive cross sec- tions with a steeply dipping fault and flower structure sug- gestive of strike-slip motion. They described an inner and outer “thrust fault-fold complex” consisting of multiple thrust faults and thrust-generated folds associated with a positive flower structure, and they described an offshore ex- tension of the San Joaquin Hills as being dextrally offset by the ONI-RC fault zone. Grant and others (1999, 2000, 2002) reported late Quaternary and Holocene uplift of the San Joa- quin Hills inboard of the ONI-RC fault zone and suggested that seismic activity of the San Joaquin Hills blind thrust and ONI-RC fault zone are linked. Recent offshore investigations have focused on the structure and potential activity of a low-angle fault, the Oceanside thrust, and its relationship to the ONI-RC fault zone (Bohannon and Geist, 1998; Rivero et al., 2000). Riv- ero et al. (2000) proposed that the San Joaquin Hills blind thrust is a backthrust soling into the larger Oceanside thrust.
  • 97. In this model, Quaternary uplift of the San Joaquin Hills and coastline south to San Diego is associated with movement of the Oceanside thrust. They presented four potential con- figurations for thrust interaction with strike-slip faults in the ONI-RC fault zone, each leading to a different maximum magnitude estimate, based on the activity and depth of the strike-slip system. Our analysis of microseismicity helps constrain the most viable model for interaction between the Oceanside thrust and ONI-RC fault zone. Seismicity and Event Location Method No significant historic earthquakes have occurred on the ONI-RC fault zone. Scattered seismicity has been recorded along the ONI-RC fault zone (Fig. 3), although events are difficult to locate accurately due to poor station coverage (Fischer and Mills, 1991; Astiz and Shearer, 2000). With the exception of the energetic 1986 ML 5.3 Oceanside earth- quake sequence, rates of microseismicity have been low since digital waveform data became available in 1981 (Astiz and Shearer, 2000). Fischer and Mills (1991) reported that the rate of seismicity along the ONI-RC fault zone is ap- proximately 10 times lower than along the onshore NIFZ. Rivero et al. (2000) interpreted the 1986 Oceanside earth- quake sequence as generated by the Thirtymile Bank thrust
  • 98. fault, a reactivated low-angle detachment fault. Figure 3 shows microseismicity of the inner Continental Borderland between 1981 and February 2003. The earth- quake locations were computed from the archived P and S picks using the source-specific station term method of Richards-Dinger and Shearer (2000). This method improves the relative location of events within compact clusters by solving for custom station terms for subsets of the events, thus removing the biasing effects of unmodeled three- 750 Short Notes -118.0 -117.5 33.2 33.4 33.6 33.8
  • 99. Oceanside Newport Beach A B Elsinore San Clemente Newport-Inglewood P alos Verdes 0 10 km Figure 3. Earthquakes (black dots) near the coastal cities (circles) of Newport Beach, San Cle- mente, and Oceanside, California, recorded from 1981 to February 2003. Black boxes indicate clusters of similar events near the ONI-RC fault zone. A 1981 cluster near Oceanside (box A) is shown in Figure 4,
  • 100. and a 2000 cluster near Newport Beach (box B) is shown in Figure 5. The approximate locations of the Elsinore, Newport–Inglewood, and Palos Verdes faults are also shown. -117.52 -117.51 33.26 33.27 A A' 0.0 0.4 0.8 12.0 12.5 13.0 13.5
  • 101. Distance (km) D e p th ( km ) A A' 0.5 km0 Figure 4. The 1981 Oceanside cluster (from box A on Fig. 3) is plotted at higher resolution in map view and cross section, after relocation using differential times computed using waveform cross- correlation. dimensional velocity and the varying station coverage among the events. In general, this results in sharper and better-defined seismicity alignments. The method does not,
  • 102. however, necessarily yield improvements in the absolute lo- cations of the clusters. We obtained waveforms for these events and performed waveform cross-correlation between each event and 100 neighboring events (Shearer, 1997; Astiz et al., 2000). We then identified clusters of similar events and relocated the microearthquakes within similar event clusters using the method of Shearer (1998, 2002). On Fig- ure 3, each cluster is so compact that these appear mostly as single dots. Two of these clusters, in boxes labeled A and B in Figure 3, are associated with the ONI-RC fault zone and are the focus of this article. Results The Oceanside Cluster The most interesting cluster is from a 1981 swarm of 19 M �3.0 earthquakes approximately 10 km northwest of Oceanside. This cluster should not be confused with the more energetic 1986 ML 5.3 Oceanside earthquake sequence located much further offshore. As shown in Figure 4, the events align along a north-northwest trend about 0.5 km long. In cross section, the events define a nearly vertical plane between 12.5 and 13.0 km depth. The strike, dip, and
  • 103. location of a plane fit by these events are consistent with active strike-slip faulting on the ONI-RC fault zone. Fischer and Mills (1991) reported that the 1981 cluster represents “a direct correlation of seismicity” (p. 30) with the ONI-RC fault zone because they located the earthquakes within 0.5 km of where they mapped the Oceanside segment of the fault zone. Our relocation results support this inter- pretation by showing that the events within the cluster form a linear trend that is approximately parallel to the fault zone. Fischer and Mills (1991) also calculated a strike-slip single event first motion focal mechanism solution for an earth- quake in the 1981 cluster. We are doubtful that unique focal mechanisms can be computed for the 1981 events, given the sparse distribution of seismic stations. However, we did ex- amine composite waveform polarities from the 1981 cluster using the method of Shearer et al. (2003) and found that the polarities are consistent with a right-lateral strike-slip focal mechanism solution, aligned with the trend of the ONI-RC Short Notes 751 -117.870 -117.860
  • 104. 33.545 33.550 33.555 B B' 0.0 0.5 1.0 1.5 6 7 Distance (km) D e p th
  • 105. ( km ) B B' 0.5 km0 Figure 5. The 2000 Newport Beach cluster (from box B on Fig. 3) is plotted at higher resolution in map view and cross section, after relocation using differential times computed using waveform cross- correlation. fault zone. We cannot, however, entirely eliminate other possible focal mechanisms. The Newport Beach Cluster A cluster of seven similar microearthquakes occurred offshore of Newport Beach in 2000 at a depth of approxi- mately 6.5–7.0 km. The cluster is near the offshore NIFZ as mapped by Morton and Miller (1981) and compiled by Fischer and Mills (1991). In map view and cross section
  • 106. (Fig. 5), five of the seven events are aligned in a pattern consistent with a shallow (7 km), north-northwest-striking, vertical or steeply dipping active fault. Observed polarities for this cluster are too sparse to provide any meaningful constraint on the focal mechanisms. The location of the 2000 cluster is southwest of an area of active faulting reported by Fischer and Mills (1991) from mapping offshore of Newport Beach and the San Joaquin Hills. Fischer and Mills (1991) analyzed microseismicity from 1982 to 1990 along this zone and reported strike-slip and reverse-fault first motion solutions from a cluster of epi- centers between 2 and 2.5 km on either side of the mapped fault zone. Discussion The mean locations of each cluster are determined using a standard 1D velocity model for all of southern California and are probably not very accurate, particularly in depth. We estimate uncertainties on the absolute cluster locations as roughly �2 km in horizontal position and �5 km in depth. However, the relative location accuracy among events within each cluster is much better constrained, so the trends defined by the seismicity alignments should be reliable. The
  • 107. estimated relative location accuracy is generally less than 50 m for the events in the Oceanside cluster and less than 150 m for the aligned events in the Newport Beach cluster. The two outlying events in the Newport Beach cluster have standard errors of about 500 m; thus it is entirely possible that their true locations lie within the linear trend of the other events. The location, alignment, and apparent dip of the Ocean- side and Newport Beach clusters of microearthquakes are consistent with the presence of active, steeply dipping faults in the ONI-RC fault zone. Waveform polarities are also con- sistent with strike-slip motion, although other mechanisms cannot be ruled out. For hazard estimation of offshore faults, it is not as im- portant to precisely locate active traces as it is for onshore faults in populated areas. A more important consideration is the potential for a through-going rupture and an estimate of the maximum magnitude earthquake. Our relocated micro- seismicity is too sparse to reveal whether or not there is a through-going strike-slip fault zone. The structure of the ONI-RC fault zone may be similar to that of the onshore NIFZ, which contains multiple strike-slip traces. Others have mapped a fairly continuous structurally complex zone of
  • 108. faulting �110 km long, subparallel to and within 10 km of the coast (Fischer and Mills, 1991). The maximum magni- tude of an ONI-RC fault rupture can be estimated from the length of the fault zone. Assuming a 110-km surface rupture length of a strike-slip fault zone yields an estimated M 7.4 earthquake (Wells and Coppersmith, 1994). If strike-slip faults do not terminate the Oceanside thrust, Rivero et al. (2000) estimate an Mw 7.5 maximum magnitude earthquake could result from rupture of the entire thrust fault. The �6.5-km depth of the Newport Beach seis- micity cluster does not provide information on the geometry or interaction between the strike-slip ONI-RC fault zone and the Oceanside thrust. However, the location and �13 km depth of the Oceanside cluster suggests that the Oceanside thrust is terminated by active strike-slip faults. According to Rivero et al. (2000), this geometry would lead to an Mw 7.3 maximum magnitude earthquake on the Oceanside thrust. The maximum magnitude estimate is at the upper range of magnitude estimated for an earthquake that uplifted the San Joaquin Hills circa A.D. 1635–1769 (Grant et al., 2002).
  • 109. 752 Short Notes Acknowledgments This research was supported by the Southern California Earthquake Center. SCEC is funded by NSF Cooperative Agreement EAR- 0106924 and USGS Cooperative Agreement 02HQAG0008. The SCEC Contribution Number for this article is 746. References Anderson, J. G., T. K. Rockwell, and D. C. Agnew (1989). Past and possible future earthquakes of significance to the San Diego region, Earth- quake Spectra 5, 299–335. Astiz, L., and P. M. Shearer (2000). Earthquake locations in the inner Con- tinental Borderland, offshore southern California, Bull. Seism. Soc. Am. 90, 425–449.
  • 110. Astiz, L., P. M. Shearer, and D. C. Agnew (2000). Precise relocations and stress-change calculations for the Upland earthquake sequence in southern California, J. Geophys. Res. 105, 2937–2953. Barrows, A. G. (1974). A review of the geology and earthquake history of the Newport–Inglewood structural zone, Calif. Div. Mines Geol. Spe- cial Report 114, 115 pp. Bohannon, R. G., and E. Geist (1998). Upper crustal structure and Neogene tectonic development of the California continental borderland, Geol. Soc. Am. Bull. 110, no. 6, 779–800. Bryant, W. A. (1988). Recently active traces of the Newport– Inglewood fault zone, Los Angeles and Orange Counties, California, California Division of Mines and Geology, DMG OFR 88-14.
  • 111. Crouch, J. K., and J. Suppe (1993). Late Cenozoic tectonic evolution of the Los Angeles basin and inner California borderland: a model for core complex-like crustal extension, Geol. Soc. Am. Bull. 105, 1415–1434. Fischer, P. J., and Mills, G. I. (1991). The offshore Newport– Inglewood– Rose Canyon fault zone, California: structure, segmentation and tec- tonics, in Environmental Perils San Diego Region, P. L. Abbott and W. J. Elliot (Editors), San Diego Association of Geologists, San Di- ego, Geological Society of America Annual Meeting, 17–36. Freeman, S. T., E. G. Heath, P. D. Guptil, and J. T. Waggoner (1992). Seismic hazard assessment, Newport–Inglewood fault zone, in En- gineering Geology Practice in Southern California, B. W. Pipkin and R. J. Proctor (Editors), Star, Belmont, California, 211–231.
  • 112. Grant, L. B., and T. K. Rockwell (2002). A northward- propagating earth- quake sequence in coastal southern California? Seism. Res. Lett. 73, no. 4, 461–469. Grant, L. B., J. T. Waggoner, C. von Stein, and T. K. Rockwell (1997). Paleoseismicity of the north branch of the Newport–Inglewood fault zone in Huntington Beach, California, from cone penetrometer test data, Bull. Seism. Soc. Am. 87, 277–293. Grant, L. B., K. J. Mueller, E. M. Gath, H. Cheng, R. L. Edwards, R. Munro, and G. L. Kennedy (1999). Late Quaternary uplift and earthquake potential of the San Joaquin Hills, southern Los Angeles basin, Cali- fornia, Geology 27, 1031–1034. Grant, L. B., K. J. Mueller, E. M. Gath, and R. Munro (2000). Late Qua- ternary uplift and earthquake potential of the San Joaquin Hills,
  • 113. south- ern Los Angeles basin, California—Reply, Geology 28, 384. Grant, L. B., L. J. Ballenger, and E. E. Runnerstrom (2002). Coastal uplift of the San Joaquin Hills, southern Los Angeles Basin, California, by a large earthquake since 1635 A.D., Bull. Seism. Soc. Am. 92, no. 2, 590–599. Harding, T. P. (1973). Newport–Inglewood Trend, California: an example of wrenching style of deformation, Am. Assoc. Petrol. Geol. Bull. 57, no. 1, 97–116. Hauksson, E., and S. Gross (1991). Source parameters of the 1933 Long Beach earthquake, Bull. Seism. Soc. Am. 81, 81–98. Hill, M. L. (1971). Newport–Inglewood zone and Mesozoic subduction, California, Geol. Soc. Am. Bull. 82, 2957–2962.
  • 114. Legg, M. R. (1991). Sea Beam evidence of recent tectonic activity in the California Continental Borderland, in The Gulf and Peninsular Prov- ince of the Californias, P. Dauphin and G. Ness (Editors), American Association of Petroleum Geologists Memoir 47, 179–196. Lindvall, S. C., and T. K. Rockwell (1995). Holocene activity of the Rose Canyon fault zone in San Diego, California, J. Geophys. Res. 100, no. B12, 24,121–24,132. Morton, P. K., and R. V. Miller (1981). Geologic map of Orange County California, showing mines and mineral deposits, Calif. Div. Mines Geol. Bull. 204, plate 1, 1:48,000 scale. Richards-Dinger, K. B., and P. M. Shearer (2000). Earthquake locations in southern California obtained using source-specific station terms,
  • 115. J. Geophys Res. 105, 10,939–10,960. Rivero, C., J. H. Shaw, and K. Mueller (2000). Oceanside and Thirty-Mile Bank blind thrusts: implications for earthquake hazards in coastal southern California, Geology 28, 891–894. Rockwell, T., and M. Murbach (1999). Holocene earthquake history of the Rose Canyon fault zone: final technical report submitted for USGS Grant No. 1434-95-G-2613, 37 pp. Shearer, P. M. (1997). Improving local earthquake locations using the L1- norm and waveform cross-correlation: application to the Whittier Nar- rows, California, aftershock sequence, J. Geophys. Res. 102, 8269– 8283. Shearer, P. M. (1998). Evidence from a cluster of small earthquakes for a fault at 18 km depth beneath Oak Ridge, southern California,
  • 116. Bull. Seism. Soc. Am. 88, 1327–1336. Shearer, P. M. (2002). Parallel faults strands at 9-km depth resolved on the Imperial Fault, southern California, Geophys. Res. Lett. 29, no. 14, 1674. Shearer, P. M., J. L. Hardebeck, L. Astiz, and K. B. Richards- Dinger (2003). Analysis of similar event clusters in aftershocks of the 1994 Northridge, California, earthquake, J. Geophys Res. 108, no. B1, 2035, doi 10.1029/2001JB000685. Southern California Earthquake Center Working Group on the Probabilities of Future Large Earthquakes in Southern California (SCECWG) (1995). Seismic hazards in southern California: probable earthquakes, 1994–2024, Bull. Seism. Soc. Am. 85, 379–439. Walls, C., T. Rockwell, K. Mueller, Y. Bock, S. Williams, J.
  • 117. Pfanner, J. Dolan, and P. Feng (1998). Escape tectonics in the Los Angeles met- ropolitan region and implications for seismic risk, Nature 394, 356– 360. Wells, D. L., and K. J. Coppersmith (1994). New empirical relationships among magnitude, rupture length, rupture area, and surface displace- ment, Bull. Seism. Soc. Am. 84, 974–1002. Wilcox, R. E., T. P. Harding, and D. R. Seely (1973). Basic wrench tec- tonics, Am. Assoc. Petrol. Geol. Bull. 57, no. 1, 74–96. Wright, T. L. (1991). Structural geology and tectonic evolution of the Los Angeles basin, In Active Margin Basins, K. T. Biddle (Editor), Amer- ican Association of Petroleum Geologists Memoir 52, 35–134. Yeats, R. S. (1973). Newport–Inglewood fault zone, Los Angeles basin,
  • 118. California, Am. Assoc. Petrol. Geol. Bull. 57, no. 1, 117–135. Yeats, R. S. (2001). Living with Earthquakes in California, Oregon State University Press, Corvallis, 406 pp. Department of Environmental Health, Science, and Policy University of California Irvine, California 92697-7070 [email protected] (L.B.G.) Institute of Geophysics and Planetary Physics Scripps Institution of Oceanography University of California, San Diego La Jolla, California 92093-0225 [email protected] (P.M.S.) Manuscript received 23 July 2003. Paper Outline Due Thursday April 23 Posted on GWC Blackboard 3-5-20 Announced in class and illustrated on Thursday March 5, 2020
  • 119. Required: 1. Two (2) reviewed articles with a person or persons as the author. The U.S.G.S or the EPA is not a person. The last pages of your article must have a reference page. No reference page or named person or multiple authors are not acceptable. 2. Two but not more than five pages of text double or 1.5 spaced with normal margins like those in the sample outline. 3. Sources that are good various reviewed journals Science, Nature, Scientific American. These publications will have sources at the end with references within the text of the paper. 4. Sources to stay away from Science Direct, Geology.com, National Geographic, Newspapers, Magazines. If the article you chose has advertisements then stay away
  • 120. from these sources, or any with advertising ads. What to hand in. 1. Your outline as a hard copy. 2. Complete printed out copies of your sources all clipped together. 3. Must be clipped together with the outline. No folded corners on the papers as an attempt to keep them together. Use a clip. Failure to provide a copy of your sources along with the outline will result in a failing grade for this outline. No late assignments will be accepted. No electronic files accepted.