You are on page 1of 30

Back

13
Hein, Frances J., 2013, Overview of heavy oil, seeps, and oil (tar) sands, California,
in F. J. Hein, D. Leckie, S. Larter, and J. R. Suter, eds., Heavy-oil and oil-sand
petroleum systems in Alberta and beyond: AAPG Studies in Geology 64,
p. 407 – 435.

Overview of Heavy Oil, Seeps, and Oil (Tar)


Sands, California
Frances J. Hein
Energy Resources Conservation Board, Suite 1000, 250-5th St. SW, Calgary, Alberta, T2P 0R4, Canada
(e-mail: fran.hein@ercb.ca)

ABSTRACT
California has one of the largest reserves for heavy oil in the world, second only to Venezuela.
Recent declines in conventional resources and reserves during the last decade have prompted
other jurisdictions to examine their prospective unconventional resources, such as heavy oil
and oil sands, in a more favorable technological and economic setting. However, this has not
been done universally in the United States, where thermal enhanced oil-recovery technologies
(mostly used to produce heavy oil) have experienced a decline in production, concomitant with
the downturn in conventional production.
In California, the seep and oil-sand deposits are mostly unconsolidated sands bound to-
gether by biodegraded bitumen. Source rocks for both modern seeps and oil sands and ancient
heavy-oil deposits are mainly the Miocene Monterey diatomites and equivalent diatomaceous
mudstones and organic shales. In California, most of the seeps and oil sands overlie or are
updip from underlying heavy-oil reservoirs. The seep and oil-sand deposits occur in areas
where cap-rock integrity was compromised for the underlying heavy-oil reservoirs, breeched
mainly by faults or fractures. Hydrocarbons migrated updip into basin-marginal settings or in
structural areas of compromised cap rock and then pooled to form the seeps, later hardening
into oil-sand deposits. Hydrocarbons accumulated in a wide variety of depositional environ-
ments from deep-sea fans, lobes, and submarine channels to fluvial-lacustrine deltas and in-
cised valleys and every other sedimentary environment in between. This makes it difficult to
identify type examples for the California accumulations, although case examples are given.
In the past, steam-flood, condensed water drive, cyclic steam stimulation (CSS), and fire-
flood were used to produce the California heavy-oil reservoirs. Currently, significant California
CSS projects underway include Belridge, Cymric, and northern Midway Sunset fields to stim-
ulate intermediate-gravity hydrocarbons in the Monterey, Reef Ridge, and Etchegoin diatomite
lithologies. Elsewhere, for example, in Canada, in-situ bitumen and extra-heavy-oil sands are
commonly developed using CSS or steam-assisted gravity drainage (SAGD). Combined ap-
plication of CSS along with SAGD from horizontal wells may recover bypassed pay in heavy-
oil reservoirs and may be used to recover bitumen from associated oil sands, and multistage
multifracing technologies may recover oil from the deeper unconventional (Monterey) source
rocks.

Copyright n2013 by The American Association of Petroleum Geologists.


DOI:10.1306/13371587St643550

407
408 Hein

These technological developments, along with improved computing techniques (i.e., three-
dimensional [3-D] geologic modeling/visualization), allow for real-time exploration and de-
velopment of unconventional reservoirs. A significant effort exists in California to improve
recovery from Pleistocene, Pliocene, and Miocene heavy-oil deposits; for example, at present,
70 to 80% recovery from heavy-oil steam drives is seen in Pleistocene Tulare Formation fluvial
and alluvial sands. Full 3-D models anchored by extensive coring and logging programs have
reaped benefits in many older oil fields (e.g., South Belridge, Midway Sunset, Cymric, Lost
Hills, Kern River). Bypassed pay and new production from associated shallow oil sands and
deeper source rocks may ultimately be a key to attainment of increases in secure unconven-
tional energy reserves in North America. In the future, full integration of new technologies,
along with technology sequencing, may be applied to old California oil fields for production
of bypassed pay in heavy-oil fields.

INTRODUCTION bbl of oil remaining (Jaramenko, 2011). For more than


60 yr in California, steam injection, along with steam-
In its most recent assessment, the International Energy flood and condensed-water drive, has been used to
Agency (2010) predicted that unconventional oil pro- produce heavy oil (in-situ viscosity <10,000 cp). Pres-
duction will rise to more than 9 million bbl/day by sure gradients are maintained by continuous injection
2035 compared with a production of about 2 million that keeps the heated oil and condensed water flowing
bbl/day in 2009. to production wells (Dusseault, 2013).
Currently, conventional oil production and thermal In the past, steam-flood, condensed-water drive, cyclic
enhanced oil recovery (EOR) in the United States and steam stimulation (CSS), and fire-flood were used to
heavy-oil production in California are in a decline (U. S. produce the California heavy-oil reservoirs. Currently
Department of Energy, 2010). As production from these significant California CSS projects underway include
commodities becomes more limited, other unconven- Belridge, Cymric, and northern Midway Sunset fields
tional resources such as tight oil, shale gas, oil sands to stimulate intermediate-gravity hydrocarbons in the
(also called tar sands), oil shale, and shale oil are re- Monterey, Reef Ridge, and Etchegoin diatomite lithol-
ceiving increased interest on a world basis. ogies. These technological developments, along with im-
Almost all of California’s main oil fields (Figure 1) proved computing techniques (i.e., three-dimensional
were discovered between 1890 and 1925, by which time [3-D] geologic modeling/visualization), allow for real-
California was a major exporter of oil, accounting for time exploration and development of unconventional
more than 22% of the world oil production (Tennyson, reservoirs. A significant effort exists in California to
1998, 2005; Croft and Patzek, 2009). In 1985, California improve recovery from Pleistocene, Pliocene, and Mio-
oil production peaked at 1.1 million bbl of oil/day (BOPD) cene heavy-oil deposits; for example, at present, 70 to
(California Department of Conservation, Division of Oil, 80% recovery from heavy-oil steam drives is seen in
Gas and Geothermal Resources, 2006). As conventional Pleistocene Tulare Formation fluvial and alluvial sands.
hydrocarbon production decreased, the development of Full 3-D models anchored by extensive coring and
thermal production in the 1960s allowed continued oil logging programs have reaped benefits in many older
production of California’s heavy-oil deposits, which sur- oil fields (e.g., South Belridge, Midway Sunset, Cymric,
passed conventional oil production in the mid-1980s Lost Hills, Kern River). Bypassed pay and new produc-
(Olsen and Ramzel, 1995). By 2006, California’s oil pro- tion from associated shallow oil sands and deeper source
duction declined to 685,000 BOPD, of which 409,000 rocks may ultimately be a key to the attainment of in-
BOPD was heavy oil and 113,000 BOPD was offshore creases in secure unconventional energy reserves in
production (California Department of Conservation, North America. In the future, full integration of new tech-
Division of Oil, Gas and Geothermal Resources, 2006). nologies, along with technology sequencing (Dusseault,
Much of the world’s heavy-oil technologies were orig- 2013), may be applied to old California oil fields for
inally developed in California. One example is the Kern production of bypassed pay in heavy-oil fields.
River field of the San Joaquin Basin (Figure 1), the 10th Most of the California oil seeps, heavy-oil, and oil-
largest heavy-oil field in the world (10–148 API). It started sand occurrences are located along the margins of the
producing in 1902 and 107 yr later (in 2009), contin- San Joaquin Basin and Central (or Great) Valley or in
ued to produce about 29 million bbl of oil (Beeson and the Los Angeles and Ventura basins, generally along
Singer, 2011). As of 2011 it has an estimated 725 million the main San Andreas Fault system (Colvin, 1968; U.S.
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 409

Figure 1. General map of Califor-


nia showing the location of the
oil and gas fields and of the major
sedimentary basins that contain
discovered or prospective hydro-
carbon resources. Generalized
sediment thicknesses are shown:
yellow = 6096 m (0 – 20,000 ft);
orange = 6096 to 12,192 m
(20,000 – 40,000 ft); brown =
12,192 – 18,288 m (40,000 –
60,000 ft); dashed basin out-
lines indicating areas where no
thickness data were available
(from Zieglar and Cassell, 1978;
Alexander, 1986). 100 mi
(161 km).

Geological Survey, 2005) (Figures 1, 2; Table 1). Seeps, old, long-producing oil and heavy-oil fields of California
heavy oil, and oil sands in California are associated have significant parts of bypassed pay that may be re-
with underlying major hydrocarbon fields, where the covered with these improving technologies (Olsen and
subsurface oil fields are in communication with sur- Ramzel, 1995; Dusseault, 2013).
face or shallow groundwaters. The oil sands of Cali-
fornia are relatively small (5.34 billion bbl); however, Previous Work
the oil sands are associated with heavy-oil fields, for
which California has the largest heavy-oil reserves One of the first compilations concerning the reserves,
(62.85 billion bbl) in North America (Hein, 2006). Sev- geology, and technological developments associated
eral significant heavy-oil fields on the east side of the with heavy oil and bitumen in California was done as
San Joaquin Basin (e.g., Kern River, Kern Front, Poso an AAPG Research Conference held in 1984 in Santa
Creek, Mt. Poso) and the Wilmington field of the Los Maria, California (Meyer et al., 1984a), published later
Angeles Basin are extensively developed with thermal by Meyer (1987). Other reviews of heavy-oil and bitu-
methods, using state-of-the-art techniques, such as men occurrences include those of Jennings (1957),
horizontal wells, full 3-D geologic and reservoir mod- Hills (1974), Walters (1974), Demaison (1977), Hallmark
els, use of aquifer drawdown to expand steam fronts, (1980), Interstate Oil Compact Commission (1984),
and so on. As estimated by several workers, many of the Meyer et al. (1984b, 2007), Herron (2003), Meyer and
410 Hein

Figure 2. Index map of California


showing the main areas of oil
and gas fields and active gas and
oil seeps with the location of the
San Andreas Fault system and
other major faults (modified from
U.S. Geological Survey, 2005).

Attanasi (2003), Hein (2006), and Hein and Higley Oil Seeps, Heavy Oil, Bitumen, Asphaltum, and Oil
(2010). Further references are given in Table 1. (Tar) Sand
The history of oil and gas production in California
has been well documented by several authors, most no- Studies of oil and gas seeps have led to the discoveries
tably by Orcutt (1924), Heizer (1940), California Depart- of many hydrocarbon deposits throughout the world
ment of Conservation, Division of Oil and Gas (1981), (U.S. Geological Survey, 2000b). In California, the oc-
and the San Joaquin Geological Society (2002a, b; 2005). currence of seeps have been known for thousands of
An overview of heavy-oil technologies in California was years, with aboriginals using hardened oil from seeps
published by Guerard (1989), with a recent review of the for various waterproofing and for making implements
future of California’s oil supply by Croft and Patzek (2009). (Heizer, 1940; Mulqueen, 2007). The most famous of
Ongoing environmental concerns associated with both the southern California oil seeps is the La Brea Tar Pits
natural and anthropogenic oil seeps have led to a cat- at Rancho La Brea in the Los Angeles Basin (Figures 1,
alog of several seep occurrences in both the offshore and 3). During the Pleistocene, these tar pits were a trap for
onshore areas of California (Wilkinson, 1972; Hodgson, large mammals and birds, forming one of the world’s
1980; U.S. Geological Survey, 1999a, b; 2000a, b; Kunitomi most prolific fossil sites (Natural History Museum of
et al., 2007). Los Angeles County, 2005). Other similar tar pits near
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 411

Carpinteria in the Ventura Basin (Figures 1, 3) have and about 2% of the bitumen resources in the United
yielded a similar prolific collection of fossils (California States and Canada (Hein, 2006). As with many other
Department of Conservation, Division of Oil and Gas, heavy-oil and bitumen deposits, the California occur-
1993). rences are at relatively shallow depth; associated with
For seeps and the associated heavy oil, bitumen, and oil seeps, springs, and/or mud volcanoes; occur near
asphaltum to occur, subsurface reservoirs need to have faults or unconformities; and are commonly in com-
communication to surface. Oil, water, and gas follow bined structural-stratigraphic traps near the updip end
open pathways (most commonly faults and fractures of basin margins. The largest seeps, heavy oil, and oil
or open porosity networks in mainly unconsolidated sands in California are associated with faulted margins,
sediments) to the ground surface where the oil pooled, anticlines, and faulted anticlines, most as structural or
forming the seeps. Petroleum seeps have a variety of combined structural and stratigraphic traps (Table 1)
forms and compositions (Figure 4), the more common (de Chadenedes, 1984, 1987; Hein, 2006).
being (Mulqueen, 2007) The California occurrences are relatively small- and
medium-size deposits (compared with the Canadian
 Oil emanation: surface discharge of crude light or examples) that are related to the loss of cap-rock integ-
heavy oil, some with gas and/or water (Figure 4B); rity of underlying hydrocarbon reservoirs. These oc-
 Light crude: oil with more than 208 API and vis- currences are similar to other small- and medium-size
cosities of less than 100 cp; deposits of the Basin and Range Province (Schamel,
 Heavy oil: oil with 10 to 208 API and viscosity from 2013) or in the midcontinent areas of the United States
100 to 10,000 cp, a product of biodegradation by (May, 2013). In all of these cases, the loss of integrity of
mostly anaerobic bacteria of light oil in the sub- the cap-rock seals (mostly by faulting) put the subsur-
surface, with some volatilization (Figure 4B); face hydrocarbons in contact with groundwater, atmo-
 Bitumen: extra-heavy oil and oil (or tar) sands with sphere, or seawater (for the coastal and offshsore
less than 108 API and viscosity of more than 10,000 cp. California occurrences). This leads to contamination
Surface accumulations of bitumen may extrude of the shallow subsurface reservoirs by anaerobic
from fractures or along faults (Figure 4C), and some bacteria, which then biodegrade the light-oil hy-
may become resedimented into tar balls, ropes, or drocarbons into heavy oil or bitumen (Requejo and
petroleum-saturated sand in beach environments; Halpern, 1989; Head et al., 2003). Usually, the heavier
 Asphaltum: highly viscous crude oil consisting of oils and oil sands occur close to or at the surface. In some
bitumen, asphaltene, water, soil, rock, and organic cases, such as the Ventura Avenue Anticline, reverse
matter. Asphaltum can occur as mounds or flows, gradients in oil viscosity occur, with the heavier oils
and some asphaltum is liquid hydrocarbon that lying below lighter oil, mostly a consequence of ground-
has weathered into a semisolid after being exuded water washing through fracture networks at depth
on the surface. Many asphaltum mounds continue (Roberts, 1987; Hein, 2006).
to flow for years after active discharge of the seep The oil-producing basins of California are mainly
has stopped, depending on local conditions such strike-slip and sag basins that developed in response to
as temperature and ground slope (Figure 4A) transpressive forces along the strike-slip San Andreas
(Mulqueen, 2007); Fault and its subsidiary faults (Figure 2). Within the
 Oil (tar) sand: generally unconsolidated sand that California basins, 3.6 to 14 km (2.2–8.7 mi) of sed-
is held together by degraded hydrocarbon, com- imentary fill accumulated. Basin trends are generally
monly bitumen; has less than 108 API, with vis- parallel to the strike of the San Andreas Fault (Figures 1,
cosity of more than 10,000 cp (Figure 4A, C). 2). In some cases, basins such as the Miocene Ridge
Basin are off-trend. Here, increased subsidence along
the subsidiary Garlock Fault caused a thick, local sag
GENERAL GEOLOGY OF HEAVY OIL, SEEPS, AND OIL basin to form, which accumulated up to 14 km (8.7 mi)
SANDS IN CALIFORNIA of sedimentary fill (Crowell, 1974a, b, 1987). Other tec-
tonic effects on sedimentation occur east of the San
Oil exploration in California began in 1865, originally on Andreas Fault in the San Joaquin Basin (Figures 1, 3),
surface outcrops or shallow oil sands and seeps, many where subsidence and uplift relate to differential tec-
of which are updip or lie above the largest oil and gas tonic movements along the transform boundary. Here,
fields in the area (Figures 1–3) (de Chadenedes, 1984, local transpressional folding occurred along the east
1987). California is estimated to have about 63 billion side of the San Andreas Fault, creating en echelon anti-
bbl of heavy oil and just more than 5 billion bbl of bi- clinal structures that form many of the oil fields in the
tumen in place, representing about 78% of the heavy oil area (Harding, 1976). In general, for California, the main
412
Table 1. Summary of characteristics of the main oil (tar) sand and heavy oil deposits in California.*

Formation and Reserves

Hein
Name Type Lithology (billion bbl) Age Basin or Uplift Structural Setting References
Careaga and Foxen Oil (tar) sands Pliocene Santa Maria Closed anticlines 1, 2**
Canyon
Casmalia Heavy oil; Sisquoc dt/ 0.9 – 1.5 Miocene/Pliocene Santa Maria Faulted anticline 3, 9**
diatomite Careaga ssy
Cat Canyon Oil (tar) sands 0.83 Miocene Santa Maria Faulted margin 9**
Chino Oil (tar) sands Pliocene ssy >0.05 Pliocene Unknown 3**
Cholame Creek Oil (tar) sands Poncho Rico ssy 0.05 – 0.10 Miocene/Pliocene Unknown 3**
Coalinga Heavy oil >0.7 Miocene Coalinga Faulted anticline 4**
Edna (Arroyo Grande) Oil (tar) sands; Pismo ssy 0.2 – 0.31 Miocene Santa Maria Flank of anticline/ 1, 2, 3, 5, 9**
heavy oil regional sand wedge
Foxen (Basal Foxen) Oil (tar) sands 1.9 Pliocene Santa Maria Closed anticlines 9**
Gaviota Oil (tar) sands Pleistocene ss,y 0.1 Miocene/Pleistocene Santa Barbara Faulted margin 3**
Vaqueros ssy
Indian Valley Oil (tar) sands Paso Robles ssy >0.05 Miocene Santa Barbara Faulted margin/ 3**
sandstone wedge
La Brea Oil (tar) sands Miocene ssy 0.05 – 0.15 Miocene/Pleistocene Los Angeles Anticline 3**
Loma Verde Oil (tar) sands 0.9 Miocene/Pliocene Los Angeles Anticline 9**
McKittrick Heavy oil; Tulare ss, dty 0.83 Miocene San Joaquin Faulted anticline 1, 2, 3, 6**
diatomite
Mylar Oil (tar) sands Poncho Rico ssy 0.05 – 0.1 Miocene/Pliocene San Joaquin Faulted anticline/ 3**
sandstone wedge
Orcutt Heavy oil, Sisquoc dty 0.05 – 0.1 Miocene Santa Maria Anticline/flank 3**
diatomite of anticline
Oxnard Oil (tar) sands 0.5 Miocene/Pliocene Ventura Syncline 3, 9**
Paris Valley Oil (tar) sands 0.7 Miocene/Pliocene Paris 9**
Point Arena Oil (tar) sands Vaqueros ssy 0.1 Miocene/Pliocene Santa Maria Anticline 3**
Red Rock Mountain Oil (tar) sands, Sisquoc dt/ 0.15 Pliocene Santa Maria Anticline 3, 6**
diatomite Careaga ssy
Reef Ridge Oil (tar) sands Miocene ssy >0.05 Miocene San Joaquin Faulted margin/ 3**
sandstone wedge
Richfield Oil (tar) sands 0.4 Miocene/Pliocene Los Angeles Faulted anticline 9**
Santa Cruz Oil (tar) sands Santa Margarita ssy 0.1 – 0.15 Miocene/Pliocene Santa Cruz Faulted margin/ 3, 5, 9**
sandstone wedge
Sargeant Ranch Oil (tar) sands, Purisma ss, 0.4 Miocene/Pliocene Purisma Anticline 3, 5**
diatomite Monterey ss, dty
Sisquoc (Zaca_Sisquoc) Oil (tar) sands Careaga ss, 0.14 – 0.5 Pliocene Santa Maria Faulted anticline 3, 5, 6, 9**
Monterey dty
South Belridge Oil (tar) sands 1 Pleistocene San Joaquin Unconformity and 7**
anticline
South Belridge and Heavy oil; 2 Miocene San Joaquin Unconformity and 6, 7, 8**
Belridge diatomite anticline
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 413

1, 2, 6** extensional phase in middle–late Miocene was followed


3, 6**

6, 7**
by two contractional phases between the Pliocene and

3**
3**

3**
3**

3**

3**

**(1) Dibblee (1984); (2) Dibblee et al. (1987); (3) de Chandenedes (1987); (4) Bate and Graham (1987); (5) Walters (1974); (6) Ingle (1981); (7) Beyer (1987); (8) Roadifer (1987);
the Holocene (Karner and Dewey, 1986).
Initial fill of the basins, near adjacent uplands, was
mostly fluvial to shallow marine, evolving downdip
into more marine and deep-marine embayments. At
Synclinal valley fill

Complex anticline
Anticline/faulted

sandstone wedge

sandstone wedge

sandstone wedge
different times during the Miocene, regional upwelling
Faulted margin/

Faulted margin/

Faulted margin/
Faulted anticline
Anticline/flanks

Faulted margin
along the continental borderland resulted in a very
of anticline

high production of organics, along with biogenic silica,


mainly associated with diatom blooms. The resulting
margin

organic-rich shales and diatomites, mostly belonging to


the siliceous Miocene Monterey Formation (Figures 5,
6), provided much of the source rocks for the heavy-oil
reservoirs and oil-sand deposits along the margins of
Santa Barbara

Santa Maria
San Joaquin

San Joaquin

San Joaquin
Santa Cruz

Santa Cruz

the basins (Ingle, 1981; Isaacs, 1981a, b, 1987; Reid and


Ventura

McIntyre, 2001). True diatomites within the Monterey


Formation tend to have high contents of biogenic silica
and generally lack organic matter. By contrast, the as-
sociated phosphatic shales and mudstones of the Mon-
terey Formation tend to be the best and most prolific
Miocene/Pliocene

Miocene/Pliocene
Miocene/Pliocene

source rocks. Hydrocarbon migration was updip to the


margins of the individual basins and occurred during
the latest Tertiary (de Chadenedes, 1984, 1987; Dibblee,
1984; Dibblee et al., 1987; Yeates and Beall, 1991; Magoon
Miocene

Miocene

Miocene
Miocene

Miocene

Miocene

et al., 2004). Current resource plays within the Monte-


rey Formation are mainly light, tight oil that have high
enough flow rates through induced fracs for econom-
ic recovery. Typically, on their own, the more heavy-oil
0.05 – 0.15

0.05 – 0.15

0.1 – 0.2

zones within the Monterey Formation are not econom-


>0.05

>0.5

>0.5

0.13
0.1

ically produced at high enough rates through induced


fracs. For economic production, the heavy-oil zones may
be produced, along with lighter oils, that are within and
adjacent to areas of very mature Monterey source rocks.
Santa Margarita ssy

The Monterey Formation is highly fractured (Figure 6B)


and locally forms structural traps of fractured reser-
Monterey ssy

Fernando ssy
Vaqueros ssy

Temblor dty

Miocene ssy

voirs in the Santa Maria Basin and offshore Santa Barbara


Cantua ss,

Channel (Figure 1). The Monterey succession is the main


source for much of the hydrocarbons in the coastal and
offshore seeps, in the heavy-oil fields, and the oil (tar)
sand outcrops (Wilkinson, 1972; Hostettler et al., 2004;
*Modified from Hein (2006) and Dibblee (1984).
Oil (tar) sands,
Oil (tar) sands

Oil (tar) sands

Oil (tar) sands

Oil (tar) sands

Oil (tar) sands

Oil (tar) sands

U. S. Geological Survey, 2005) (Figures 1–3). Elsewhere,


other source rocks contribute to the multiple petro-
Heavy oil

Heavy oil
diatomite

leum systems operating within the different subbasins


of California. The stable carbon isotope, biomarker,
and Rock-Eval pyrolysis data indicate that, in addition
ss = sandstone; dt = diatomite.

to the Miocene Monterey Formation, some of the other


South Sulfur Mountain

(9) Kuuskraa et al. (1987).

source rocks include the Eocene Kreyenhagen shale


West Midway/Sunset

(Reef Ridge – Kettleman Hills and Coalinga oil field),


West Cat Canyon

the Miocene Point Sal shale (Santa Maria Basin), and the
West King City

West San Ardo


Tapo Canyon

upper Miocene Belridge diatomite (Reef Ridge Mem-


Upper Ojai

ber) and associated underlying shales of the McClure


Vallecitos

Formation (Reef Ridge–Kettleman Hills, South Belridge,


and McKittrick oil fields) (Figure 5) (Lillis and Stanley,
Taft

1999; Magoon et al., 2004).


y
414 Hein

Figure 3. Google map of California with highlighted part showing locations of the main oil seep and tar pits discussed in this
chapter; inset index map of part of California illustrating the coastal area, the Los Angeles and San Joaquin basins, with major oil
and heavy oil field outlines indicated in black. Modified from Tennyson (2005); U.S. Geological Survey (2005). 100 km (62 mi).

A significant amount of structural variation devel- series of transgressive–regressive cycles superimposed


oped as this part of the western North American plate wedges of sands and associated diatomites on the larger
boundary reorganized from a convergent plate boundary evolving tectonic features to produce combination
(before the Oligocene) to a fully developed right-lateral stratigraphic-structural traps for many of the California
San Andreas transform boundary by the Miocene. A oil fields (de Chadenedes, 1984, 1987). Similar regional
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 415

Figure 4. Examples of natu-


ral oil seeps, coastal area
(Figure 3), between Point
Conception and Los Angeles,
California: (A) Koeningstein
Road oil seep, north of Santa
Paula, east of Ojai, consisting
of bitumen and asphaltum;
(B) Surface emanation of
heavy oil and gas, north Sulfur
Mountain area, Ojai oil field,
Ventura County, California;
(C) Hamp oil seep (arrow),
consisting of heavy oil, bitu-
men, and bitumen-saturated
oil sand, Silverthread area,
Ojai oil field. Note fault trace
shown by the line.

interactions of sedimentation, eustasy, and tectonics re- tridynamite; opal-CT), and then eventually to form the
sulted in structural and stratigraphic complexity, inter- more stable quartz phases. These diagenetic changes
rupted by major unconformities (Figure 5A). On a re- resulted in the development of systems of fractures,
gional basis, the diatomaceous mudstones, diatomites, which augmented the structural fractures that were
and black shales are the main source rocks (Figure 5A). forming in response to tectonic forces on the brittle,
Associated heavy-oil reservoirs were emplaced within hard, siliceous rocks. The diagenesis resulted in the ex-
a variety of sandstones, including basal lowstand de- pulsion of significant volumes of water from the highly
posits and transgressive and regressive sandstones. porous matrix of the diatomite and siliceous muds.
Surficial oil sands are hosted within mainly unconsol- The micro-overpressure associated with the dewater-
idated Pliocene–Pleistocene and Pleistocene sands. ing and associated column loss (opal-A porosity up to
70%, opal-CT up to 40%, and quartz up to 20%) generated
a microfracture fabric in the less terrigenous facies of
HEAVY OIL the Monterey Formation. Because of the original low ma-
trix permeability of the Monterey diatomaceous shales,
For decades, much of the heavy-oil production has been the petroleum system flow paths were dependent on
associated with the Miocene Monterey Formation. This the fracture flow paths created during diagenesis and
is a thin-bedded, mixed-lithology succession of fine- tectonism.
grained mudstones (Figure 6)—one of the more typical Fracture and fault system patterns that developed
unconventional mudstone and mudrock reservoirs pres- were a function of the host rock type, the bed thickness,
ently being exploited in North America. The Monterey is and regional stress regimes. In the Monterey succession,
an organic-rich hemipelagic deposit, locally phosphatic, a wide variety of structural features occurs at several
calcareous, or dolomite rich. It represents hemipelagic scales (Gutierrez-Alonso and Gross, 1997). Large-scale
deposition of drowned basins in a variety of environ- structural features include normal faults, bedding-plane
ments from shelves to banktops, deep sea slopes, and detachments, low-angle thrust faults, triangle zones,
base-of-slope settings (Figure 7). thrust duplex structures, and fault blocks. Mesoscale
Later diagenesis and structural deformation of the features include: ptygmatically folded veins, folded
Monterey Formation was complex, mostly a function of beds of chert, inverted normal faults, fault-propaga-
mudrock composition, petrography, and burial history. tion folds, and breccia zones (Jeffrey et al., 1991).
Originally, the hemipelagic diatomaceous silica was Aside from the Monterey Formation, it is difficult to
deposited as a metastable opal (amorphous opal; opal-A), present representative or type deposits of heavy-oil res-
which underwent two stages of dissolution and repre- ervoirs in California because such significant variation
cipitation to form the metastable opal (cristoballite and exists between the different subbasins (Figure 7). This is
416 Hein

Figure 5. (A) Stratigraphic nomenclature of southern California for the offshore and onshore areas and oil fields discussed
in this chapter. Yellow indicates source rocks, black bars are main hydrocarbon fields, red bars show host rocks for surficial
oil-sand deposits exposed in either basin-marginal settings and/or surface areas with structural deformation, exposing the
deeper successions at surface (sources: Oakeshott et al., 1954; de Chandenedes, 1984, 1987; Dibblee, 1984; Dibblee et al.,
1987; Norton and Otott, 1996; Bridges and Castle, 2003; Imhof and Castle, 2004). (B) Approximate line of section for the
stratigraphic nomenclature discussed from the Wilmington oil field southeast to the Coalinga oil field in the northwest.
100 km (62 mi).

caused by differential rates of subsidence and uplift, shore, nonmarine fluviolacustrine and alluvial-fan en-
sediment supply, paleooceanographic histories of re- vironments. On a field basis, facies models from most
gional circulation and upwelling, and sill depths be- environments are applicable (dependent on the sedi-
tween the different subbasins (Bridges and Castle, 2003; mentary and tectonic history within an individual ba-
Bowersox, 2004; see Schwartz, 1988, for a good sum- sin). Because of this inherent complexity, in this section,
mary). All of these paleogeomorphic factors were coupled some of the major oil fields are discussed that show the
with several transgressive-regressive pulses associated main reservoir types and their dominant controls.
with changes in relative base level from the post-
Cretaceous to the Holocene. This results in stratigraphy Fluvial and Lagoonal and Nearshore Reservoirs:
and petroleum flow system(s) that are nonuniform and Coalinga Oil Field
complex (Figure 5). The overall result is a patchwork of
different sedimentation and tectonic histories and reser- The Coalinga oil field, central California (Figures 3B,
voirs that range the spectrum from deep-marine to near- 8), originally discovered in the late 1800s, started
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 417

Figure 5. (cont.).

production in 1890. Cumulative production was in Eocene Kreyenhagen Formation were the source rocks,
excess of 912 million bbl, with estimated remaining and the accumulated oil was reservoired in the overlying
reserves of 58 million bbl, and more than 1600 active Temblor Formation sandstones (Figures 5, 8). Traps are
wells (end of 2006). It is now the eighth largest pro- mostly stratigraphic in nature, with some structural
ducing oil field in California. The main operators are complications. Along the eastern side of the San Andreas
Aera Energy LLC and Chevron Corporation (California Fault reservoir rocks outcrop, several historical oil seeps
Department of Conservation, Division of Oil, Gas and (Figure 8A, B) also exist. Here, local solidified tar mats
Geothermal Resources, 2006). and tight zones within the reservoir rocks provide cap
The Coalinga oil field is one of the first discovered rocks to the subsurface reservoirs (Clark et al., 2001;
and described oil fields in California and became one of Imhof and Castle, 2004). Structure at the Coalinga oil
the more prolific producers of heavy oil in California field is a southeastern-plunging anticline (Figure 8A)
(Arnold and Anderson, 1910). The oil on the west side of (Clark et al., 2001). The Coalinga anticline is one of a
the field is heavy (11–198 API); on the east side, both series of enechelon anticlines that occur on the eastern
heavy and medium-grade oil (12–308 API). Most of the side of the San Andreas Fault and define the western fold
oil and heavy oil are produced from multiple discrete belt of the San Joaquin Basin (Bridges and Castle, 2003).
intervals within the lower part of the Temblor Formation The Temblor Formation unconformably overlies the
(Figure 5) (Bate, 1984; Bate and Graham, 1987; Clark et al., Eocene Kreyenhagen Shale and, in turn, is unconform-
2001). The Coalinga field is divided into the East and ably overlain by the Etchegoin Formation shales, the
West Coalinga, separated by the northwest–southeast- McLure Shale, the Reef Ridge diatomaceous shale, or
trending Coalinga anticline (Figure 8) (Clark et al., 2001). by regressive sands of the Santa Margarita Formation
The Coalinga oil field is part of a petroleum system, in (Figures 5; 8B, D). The Coalinga oil field reservoir con-
which the organic-rich rocks of the underlying middle sists of the Miocene Temblor Formation braided stream
418 Hein

Figure 6. Miocene Monterey Formation diatomite, southern California. (A) Representative stratigraphic column showing textural
variation on the left, mineralogy and total organic carbon plotted on the right (from Isaacs, 1981b). (B) Outcrop photograph
of wispy-laminated dark organic shale, light-gray silty mudstone, and white very fine sandstone/siltstone interlaminations,
Monterey Formation (from Isaacs, 1981b). (C) Outcrop photograph of fractured, light-tan, silica-rich diatomite, with bitumen
and heavy-oil fracture fills, overlying unfractured, parallel-laminated, fissile organic shale at the base, coastal outcrops, Point
Conception, California. (D) Photograph of hand specimen of fractured and folded Miocene Formation diatomite, Arroyo Burro,
Santa Barbara, California.

Figure 7. Submerged topography of the


California borderland during highstands
that emplaced the regional, hemipelagic,
diatomaceous shales and mudstones. These
organic mudstones, mainly from the Mio-
cene Monterey Formation, were source
rocks for later hydrocarbon accumulations,
including heavy oil, seeps, and oil sands in
California (from Gorsline and Emery, 1959,
published in Pickering et al., 1989).
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 419

Figure 8. Coalinga oil field, San Joaquin Basin. (A) Structure contour map showing the outline of the field, the southeast-
plunging Coalinga anticline, and line of section BB0. Division of the Coalinga oil field is into the West Coalinga and East Coalinga
areas. Map also shows lease locations, cored wells, sonic logs, seismic data, and outcrops. The Coalinga Nose anticline represents
the East Coalinga part of the field, and the monclinal flank is the West Coalinga part of the field. (B) Cross section BB0 of the East
Coalinga field showing the absence of an updip seal at Coalinga, where the Temblor reservoir is breached at the surface. Oil
accumulates behind near-surface tight cemented rocks and tar seals to create the underlying giant oil field. (C) Core photograph
(white light) of the basal Temblor unconformity (BT) showing onlap of Temblor sands onto the diatomaceous shales of the
Kreyenhagen Formation. (D) Core photograph (white light) of the top Temblor unconformity (TT) showing onlap of Etchegoin
shales onto the cemented, bioturbated, and oil-stained Temblor sands. Scale bar in tenths of feet. Modified from Clark et al. (2001).

sands that are overlain by transgressive lagoonal and Graham, 1987). Internally, the lower Temblor reservoir
shallow-marine sediments. Internally within the Tem- was highly compartmentalized with a complete lack of
blor Formation are five different systems tracts, including vertical transmissibility between the different braid-
incised-valley, estuarine, tide- to wave-dominated shore- channel sands (Lennon, 1976).
line, diatomite, and subtidal deposits (Bridges and Castle, Detailed mapping of both vertical and lateral res-
2003). The basal reservoir cobble conglomerates and cob- ervoir anisotropy is needed for a 3-D characterization
ble sands were emplaced as broad (tens of meters wide), and visualization of the geologic framework. This al-
shallow (several meters deep), braided channels that lows for proper placement of wells to ensure optimum
were cut into and deposited over the siliceous source production from a very heterogeneous heavy-oil res-
rock shales of the Kreyenhagen Formation (Figures 5, ervoir in a basin-marginal setting (Lennon, 1976; Bate
8C). The overlying coarse clastics are a sequence of sand- and Graham, 1987). Recently, seismic facies mapping
stone and mudstones deposited in a variety of nearshore with 3-D reservoir visualization and modeling tools
marine environments that draped the Coalinga anticline are being used routinely in the Coalinga field. These
(Clark et al., 2001). The trap is a tar seal that formed at the are used to build geologic models for reservoir plan-
air-oil contact in the eastern part of the oil field (Bate and ning, simulation, and monitoring and to construct the
420 Hein

Figure 9. (A) Cross section of the


McKittrick oil field along N408E,
subsurface depths in feet. Modi-
fied from Allan et al. (2006).
(B) Cross section of the McKittrick
oil field along N308E, subsurface
depths in feet. Modified from
Dibblee (1984). (C) Three-
dimensional block model of the
faulted anticline of the South
Belridge oil field, with zone color
key indicating the main strati-
graphic units (from Allan et al.,
2006). 1000 ft (305 m).

pathways for horizontal wells, aided with logging-while- productive in the United States. Cumulative produc-
drilling (LWD) systems (Sanford and Wildman, 1999; tion is close to 2 billion bbl, with estimated remain-
Clark et al., 2001; Imhof and Castle, 2004). ing reserves of 520 million bbl and about 6000 active
wells (end of 2006). The main operator is Aera Energy
Ltd. that has stated a production of 140,000 bbl of oil-
Fluvial, Lacustrine, and Deep-Sea Slope Reservoirs: equivalent per day (Miller and McPherson, 1992; Cali-
South Belridge Oil Field fornia Department of Conservation, Division of Oil, Gas
and Geothermal Resources, 2006).
The South Belridge oil field, southern San Joaquin Ba- This oil field has mostly produced heavy oil (10–148
sin (Figures 3, 9A – C), originally discovered in 1911, is API), mainly from the Pliocene–Pleistocene Tulare For-
the fourth largest oil field in California and sixth most mation (Beyer, 1987; Eagan et al., 1999) (Figures 5, 9B).
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 421

Figure 9. (cont.).

Different reservoir types within the field have been transition from opal-A to opal-CT occurs between 457
produced by either thermal stimulation or through hy- and 549 m (1500 and 1800 ft) depth in the southern part
draulic fracturing. The structure in the South Belridge of the field. The anticlinal structure of the diatomite
oil field consists of a northwest-trending elongate anti- rocks that contains medium-gravity oil (Figure 9C) un-
cline that is plunging to the southeast and oriented sub- derlies the Tulare oil sands (Figure 9A, B). At present,
parallel to the San Andreas Fault system (Figure 9A–C) the diatomite of the South Belridge field is being steamed
(Miller and McPherson, 1992). The shallow successions continuously on the crest of the anticline and cyclically
dip gently on the flanks of the fold. Several unconfor- on the east flank of the anticline (Miller and McPherson,
mities occur in the field, with the most prominent at the 1992). The trap at South Belridge consists of a combi-
base of the Pleistocene Tulare Formation. The produc- nation of structural and stratigraphic traps and local tar
ing intervals within the Tulare Formation are lenticular, seals in both the Tulare sands and the underlying di-
nonmarine, fluvial-lacustrine deltaic sandstone reser- atomite reservoirs. Compared with other fields in the
voirs, which attain aggregate thicknesses of about 70 m San Joaquin Basin, the petroleum accumulation in the
(230 ft) on structural closures and up to about 180 m South Belridge oil field is relatively young, and the field
(590 ft) on the flanks of the fold. These lenticular sands is not in equilibrium. Hydrocarbon migration was pre-
are interbedded with siltstone and claystone, and the depositional or syndepositional with the Tulare Forma-
main reservoir is a stratigraphic trap associated with tion, as indicated by tar mats in the Pliocene Ethegoin
a westward lensing out of the Tulare sands. Thermal Formation and earliest Tulare sediments along the bas-
stimulation is used to produce heavy oil (oil gravity al unconformity. At present, the deeper Monterey diat-
ranges from 12 to 158 API) from unconsolidated to omite is generating hydrocarbons at depth. The Mon-
weakly indurated, poorly sorted Tulare sands. Histor- terey diatomite has a subcropping relationship with
ically, the Tulare production accounted for about 80% the overlying Tulare succession along the axis of the
of the production and reserves of the South Belfied oil South Belridge anticline, thus providing the pathway
field (Beyer, 1987). for hydrocarbon charging in the oil field (Figure 9A, B)
A cross section and seismic survey of the South (Miller and McPherson, 1992). Several bitumen accu-
Belridge oil field (Figure 9A, B) shows the various struc- mulations are present at the crest of the structure that
tural and stratigraphic relationships of the Tulare reser- are thought to be paleo-tar pits, formed when the
voir sands that overlie the Monterey diatomite (opal-A) diatomite was exposed at the surface in the Pliocene –
reservoirs. In the South Belridge field, the diagenetic Pleistocene. These were subsequently overtopped by
422 Hein

late Pliocene and Pleistocene fluvial and alluvial de- the overlying Tulare sand reservoirs, where locally, they
posits. The bitumen contains very little sedimentary pinch out along the anticlinal structure, forming a mostly
material, indicating it was relatively solid at the time of stratigraphic trap. The emplacement of the hydrocarbons
Tulare deposition. into the Tulare is a complex story. What is amazing is that
The biggest challenge in production from the Tulare so much oil entered the sands in such a short period.
reservoirs is the lateral discontinuity of the reservoir A similar, but more complex, history of emplacement
sands, which typically die out in 3 to 30 m (10–100 ft) occurred at the nearby Taft-McKittrick oil field and tar
and are bounded by floodplain and lacustrine mud and pit (de Chadenedes, 1984, 1987) (Figure 9A, B). Plung-
siltstones that range from local to fieldwide in extent ing folds are faulted, deeper successions are more tight-
(Eagan et al., 1999). Normal faulting (at least five were ly folded and faulted than shallower successions, and
mapped) further complicates the stratigraphy and lo- internal unconformities and faults compartmentalize
cally may lead to compartmentalization of the heavy-oil the reservoir. Initially, a simple faulted anticline formed
reservoir. In the past, it has been difficult to place the the trap, with oil accumulating in the Miocene Reef Ridge
injector and producer wells in ideal locations. It was Member and overlying Pliocene–Pleistocene Tulare For-
also impossible to ascertain the actual geospatial rela- mation. Structural deformation during the Pleisto-
tionship of the injector and producer wells to one an- cene then thrust the Miocene on top of the Pliocene –
other and their orientation with respect to the thin Pleistocene Tulare sandstone reservoirs. Subsequent
lenticular reservoirs and internal cap rocks. In 1998, this fracturing and faulting of the Miocene overthrust suc-
led to a fieldwide development of a 3-D surveillance cession permitted further petroleum migration into
geologic framework using data from 650 wells to de- the Miocene overthrust and to the surface forming the
scribe the complex stratigraphic and structural relation- McKittrick tar pits (Figure 9A, B) (de Chadenedes,
ships of the reservoir and nonreservoir rock. The result- 1984, 1987).
ing computer model has 22 different reservoir zones,
each from 1.5 to 6 m (5–20 ft) thick, and specifically Deep-sea Submarine Slope and Base-of-slope Fan
models the influence of the five most prominent faults. Reservoirs: Wilmington Oil Field
Preliminary results of the 3-D geologic model and visu-
alization have led to optimizing the placement of the The Wilmington oil field, Los Angeles Basin (Figures 3
injector and producer wells and provide for real-time [inset], 10, 11), originally discovered in 1932, is the larg-
visualization and quantification of production and flow est oil field in California (Mayuga, 1970; Biddle, 1991)
units (Eagan et al., 1999). and the third largest in the United States (Clarke and
In the South Belridge oil field, below the Pleistocene Phillips, 2003, 2004). Original oil in place is estimated
unconformity is the Belridge diatomite (Reef Ridge to be about 9 billion bbl. Cumulative total production
Member, Monterey Formation) (Figure 5A). Oil from is more than 2.6 billion bbl, with remaining reserves of
this deeper diatomite is lighter oil (20 – 318 API). This about 300 million bbl (as of 2002), and more than 1200
deeper succession is more sharply folded than the over- active wells (California Department of Conservation,
lying Tulare Formation. The Belridge diatomite consists Division of Oil, Gas, and Geothermal Resources, 2006).
of silty or clayey diatomite with thin sandstone and The main operator is the Tidelands Oil Production
siltstone interbeds and lenses of porcelaneous clay-shale Company in the Old Wilmington, or western part, of
(Schwartz, 1988). Wells were completed by hydraulic the field, with THUMS’ (artificial islands named after
fracturing. The Belridge diatomite (Reef Ridge shale) the parent companies: Texaco, Humble Oil [now
accounted for about 20% of the production and reserves Exxon], Union Oil [now ConocoPhillips], Mobil Oil
for the field. Other very limited production occurred [now Exxon], and Shell Oil) Long Beach Unit the field
from underlying siliceous rocks of the Monterey For- contractor for the eastern part of the field (Figure 10A)
mation (below the Reef Ridge Member) because they (Clarke and Phillips, 2003, 2004).
are opal-CT and very tight. Some production occurred This oil field has highly variable oil gravities, with
from Pliocene sands and diatomite preserved on the shallower pools containing heavier oil (12–148 API) and
flanks of the anticline, where they were not removed by some deeper pools with light oil (25–328 API). Other
the Pleistocene erosion (Beyer, 1987). variations are between onshore and offshore pools
The history of the South Belridge oil field is similar (Berman and Clarke, 1987; California Department of
to other fields in the southern San Joaquin Basin. Oil Conservation, Division of Oil, Gas and Geothermal
initially accumulated on the main anticlinal structure Resources, 2006). Oil is produced mainly from the
at South Belridge in the fractured Belridge diatomite. Pliocene Repetto turbidite lobes and the underlying
Further faulting and fracturing of local cap rocks above Miocene deep-sea turbidite sands from slope and base-
the Belridge diatomite allowed the oil to migrate up into of-slope settings of the Puente Formation (Figure 5)
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 423

Figure 10. (A) Map of the Long Beach area showing the areas of the Wilmington oil field where Tidelands Oil Production Company
(western part) and THUMS Long Beach Company (Long Beach Unit) are the field contractors. The city of Long Beach, Department
of Oil Properties, operates both. The coastline, harbor areas, breakwater, and oil-drilling islands are shown for reference.
Pacific Energy Resources operates the polygonal area between the properties. (B) Aerial photograph of Long Beach – Los Angeles
Harbor showing the location of the three horizontal well projects (case histories 1, 2, 3) and fault blocks II through V (FB II– FB V).
Oil activities coexist with the busiest harbor in the country, with fault block V steam-flood operations (case history 2) beneath
Long Beach’s new $180 million aquarium and northeast of the Queen Mary dockyard. (C) Type logs for the wells 2AU 30B1 and
2AT58B 0 for the steam-flood project in FB V; original picked markers are shown in black, and newly picked markers are shown
in red, the difference of which was used to calculate subsidence related to 60 yr of oil production in the area. The inset shows
the T4 paleochannel fill from well 2AT58B. Location of wells given in Figure 11C (from Clarke and Phillips, 2003, 2004).

(Truex, 1972, 1974; Henderson, 1987; Slatt et al., 1993; 11B, C). The basic structure of the field is a large, sygmoidal,
Norton and Otott, 1996). asymmetric highly faulted anticline, with field produc-
The Wilmington oil field is part of a series of oil fields tion progressively from east to west in a series of 10
in the Los Angeles Basin that lie on a series of northwest– different fault blocks (Norton and Otott, 1996; Clarke,
southeast-trending en echelon anticlines and complex 1999). Internally, the main Wilmington anticline is cut
antiforms that are associated with the Newport- by several normal faults that compartmentalize the
Inglewood Fault (Pickering et al., 1989; Norton and reservoir into separate production units (Figures 10B,
Otott, 1996) (Figures 11A, 12). Production is from slope 11A). These faults also offset both the paleochannel
and base-of-slope turbidites and channel sands, with the reservoir sands and the onlapping turbidite reservoir
entire succession being folded and faulted (Figures 10C; sands (Figure 11B, C) (Clarke and Phillips, 2003, 2004).
424 Hein

Figure 11. (A) Structure map of the T marker in fault block IIA. Observation well locations shown by yellow dots, and trajectories
of the horizontal wells are shown by brown curving lines. Contour intervals are 15.2 m (50 ft) from 731.5 to 1036 m
(2400 to 3400 ft) below sea level (B) Three-dimensional geologic visualization of the D1F onlap onto the D2 shale in fault
block II, vertical exaggeration 2, and the units are displayed in feet; (C) Three-dimensional geologic visualization of the T2
horizon in fault block II, showing the locations of the two type wells in Figure 10. The T4 paleochannel cuts through several
horizons and is offset by the Wilmington and Ford faults (from Clarke and Phillips, 2003, 2004). 2000 ft (609.6 m).

One of the biggest challenges in secondary and ter- cally, some faults in the Repetto Formation are perme-
tiary recovery from this field is the extreme lateral and ability barriers, whereas other faults are only partially
vertical heterogeneity in both the reservoir and non- sealing (Clarke and Phillips, 2003, 2004).
reservoir rocks. Geologic heterogeneity at all scales has In some areas of the Wilmington oil field, steam-
made it difficult to properly characterize the reservoir flood expansion in the mid-1990s was done using hor-
heterogeneity. Individual turbidite beds are normally izontal wells that each replaced four or five vertical
graded and interbedded with siltstones and shales; wells (Figure 10B), with five observation wells used to
however, the turbidites occur in multistory, stacked, monitor the steam growth in the subsurface (Clarke
and faulted base-of-slope paleochannel fills or in over- and Phillips, 2003, 2004). This scheme had two steam
lapping and onlapping lenticular lobes, with internal injectors and two producers placed 122 m (400 ft)
channeling (Slatt et al., 1993). In addition to the strat- apart horizontally in a pseudo-steam assisted gravity
igraphic compilations, many small structural compli- drainage (SAGD) scheme. For the scheme to be
cations exist. The structural faulting is complex, with effective, a complete 3-D geologic model was built
vertical offsets ranging from 15 to 30 m (50–100 ft). Lo- by doing detailed wire-line-log interpretation tied to
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 425

Figure 12. Sedimentary basins


and main structural features of
the southern California Borderland
and immediately adjacent on land
areas, with approximate thick-
nesses of basin fills shown in pa-
rentheses in kilometers; as follows:
1. Ridge Basin (4.5 km; 2.8 mi);
2. Los Angeles Basin (8.0 km;
5 mi); 3. San Diego Trough (2.5 km;
1.6 mi); 4. Santa Monica Basin
(3.5 km; 2.2 mi); 5. Catalina Basin
(0.8 km; 0.5 mi); 6. Santa Cruz
Basin (1.8 km; 1.1 mi); 7. San Pedro
Basin (1.8 km; 1.1 mi); 8. San
Clemente Basin; 9. San Nicolas
Basin (1.4 km; 0.9 mi); 10. East
Cortes Basin; 11. Tanner Basin
(1.4 km; 0.9 mi); 12. Patton Basin
(1.7 km; 1.056 mi). SMI = San
Miguel Island; SRI = Santa Rosa
Island; SCrI = Santa Cruz Island;
SCI = Santa Catalina Island; SBI =
Santa Barbara Island; SNI = San
Nicolas Island; SCLI = San Cle-
mente Island. Northwest-trending
faults are dextral and synthetic to
the San Andreas Fault. Santa
Monica Fault is sinistral. Anticline
trend is mainly west-northwest
(from Pickering et al., 1989).

reservoir characterization and structural fault delinea- where water cuts are less, secondary and tertiary steam-
tion (Figure 11B, C). recovery techniques have proven to successfully pro-
A real-time 3-D working geologic model with 3-D duce the heavy oil. In some cases, steam was used to
visualization allowed for data inconsistencies to be re- fuse and indurate the unconsolidated sands to avoid
conciled and also identified areas of intraformational wellbore breakthrough and to allow selective perfora-
subsidence and compaction that related to oil produc- tions to optimize water recoveries. Three-dimensional
tion and heavy-oil withdrawal during the previous seismic studies have aided in reservoir characteriza-
60 yr, with maximum subsidence in this area of 8.8 m tion and delineation of the complete geologic frame-
(29 ft) (Figure 10C). In January 1999, a steam-flood pilot work; were used to geologically model, define, and
project was halted because of increased subsidence. exploit bypassed pay; and were used to extend the life
In October 1999, flank wells were converted to cold- of the field into more marginal areas, which previous-
water injection wells, and subsidence was halted by ly have not been developed (Clarke, 1999; Clarke and
September 2002. Elsewhere, experience from this Phillips, 2003, 2004).
steam-flood pilot project was applied successfully to
the development, definition, exploitation, and capture Seeps
of bypassed pay (Clarke and Phillips, 2003, 2004).
This recent work involving 3-D geologic modeling, Recent world estimates indicate that about 47% of crude
visualizing, and measuring-while-drilling technologies oil currently entering the marine setting is from natural
have clearly demonstrated that these new and evolving oil seepage, and natural oil seeps are the single most
technologies can greatly increase reserves and ulti- important source of oil in the ocean (Kvenvolden and
mate recovery from old oil fields in California, such as Cooper, 2003). Oil seeps are common in southern and
the Wilmington oil field (Blesener and Henerson, 1996; central California and are a natural part of the oil-
Clarke, 1999). Production from the western fault blocks producing area along the California coastline. Here,
are reaching economic limits; elsewhere, in the field, many winter storms wash up tar balls along the beaches
426 Hein

from oil that was mostly sourced in the Miocene Mon- largest oil-sand deposits occur in the Santa Maria Ba-
terey diatomite that crops out along the shoreline sin, in the coastal area of southern California (Figure 3;
(Hartman and Hammond, 1981; Kvenvolden et al., 1999, Table 1). These are mostly hosted within the Sisquoc
2002a, b; Kvenvolden and Cooper, 2003; Hostettler et al., Formation (Figure 5), which may have been a seal to
2004; Minerals Management Service, 2005). Since about migrating Miocene oil, hosted from the deeper source
1980, more than 1 million bbl of oil have seeped off along rocks of the Monterey Formation and Point Sal Mem-
the coast of central and southern California (Minerals ber (Kuuskraa et al., 1987). Other oil sands in the Santa
Management Service, 2005). These seeps form oil slicks Maria Basin are hosted in basal Foxen deposit and in
on the water surface, and some have had perennial or the overlying Careaga Formation, where underlying
continuous discharge since the late 1700s, as noted by source rocks of the Sisquoc Formation are relatively
early English and Spanish explorers (Wilkinson, 1972). thin (Figure 5) (Kuuskraa et al., 1987). In California, the
In submarine areas, a unique and diverse macrofaunal sixth largest oil-sand deposit is the Oxnard oil field of
seep community exists (Hornafius et al., 1999; Lorenson, the Ventura Basin, where bitumen is hosted within two
2001). Similarly, on land, ecological studies show areas sands of the Pico Formation (Figure 5), which uncon-
surrounding seeps to be more diverse and biologically formably overlie the source rocks of the Monterey For-
active than surrounding areas lacking seeps. This is in- mation (Kuuskraa et al., 1987).
terpreted to be caused by the bubbling of both water and Oil-sand deposits have been mined intermittently
oil from the seeps (Lorenson, 2001; Mulqueen, 2007). for hundreds of years (Earley, 1984), but to date, no
One particularly active area is in the Coal Oil and commercial in-situ thermal schemes have been oper-
Goleta points seep field, located in the offshore Santa ating. In the 1980s, Getty Oil Company ran an experi-
Barbara Channel, between Point Conception and Ventura mental heat distillation scheme on the giant McKittrick
(Figure 3) (Fischer, 1978; Hornafius et al., 1999; Broderick, oil field in the southwest San Joaquin Basin (Figures 3
2001; Washburn et al., 2005). Here, a series of seeps have [inset], 9C). Here, the basic structure is a series of
been studied since the 1980s by sampling and collection faulted (both thrust and normal) anticlines or anticline-
of natural hydrocarbon emanations in ocean-bottom syncline pairs, with most of the heavy oil and oil (tar)
steel pyramids (350-ton and 15-m [50-ft]-high structures) sands in combination structural-stratigraphic traps. Al-
called seep tents. The Coal Oil Point offshore seep area though called tar sands, the deposit is technically oil
has collectively discharged about 55,000 bbl of oil/yr sand and heavy oil, with oil viscosity increasing with
(equivalent to 50–70 bbl/day or about 1–2  104 L/day), depth from 5 to 208 API (de Chadenedes, 1984, 1987). In
with an additional 1.8 billion ft3 (510 million m3) of nat- the McKittrick oil district, oil is produced mainly from
ural gas (Allen et al., 1970; Broderick, 2001; Kvenvolden almost all of the Tertiary sands, with the largest produc-
et al., 2002b). This seep area also contributes about tion from the Miocene Temblor Formation (Figures 5;
22 tons of reactive organocarbons (ROGs precursors to 9A, B) (de Chadenedes, 1987). A diatomite project is
smog) in the Santa Barbara area each day (Clark et al., also in place to extract hydrocarbons from surface ex-
1999, 2005). The connection of submarine discharge in posures of the charged diatomite; at present, cat litter
the Coal Point seep field has been tied to subsurface is made from the non-hydrocarbon-bearing offsetting
reservoirs, where decreased natural-marine seepages diatomite strata.
correlate with increased offshore oil production in the
area (Quigley et al., 1999a, b). The collection of hydro-
carbons on the ocean floor in seep tents have cut em-
anations to the seawater and atmosphere, with notice- HISTORICAL AND TECHNOLOGICAL DEVELOPMENT OF
able reduction in the natural levels of ROGs, oil slicks HEAVY OIL, SEEPS, AND OIL SANDS IN CALIFORNIA
on the sea surface, and the number of tar balls along the
shoreline. In addition to the Coal Point seep field, a In 1850, the first distillation of seep oil into lamp oil was
1000-m (3280-ft)-thick tar mat is within and overlying by Andreas Pico, who collected and distilled seep oil in
the Monterey structures offshore Santa Maria. This tar Pico Canyon, to be used as lamp oil in the San Fernando
mat is the seal for fractured Monterey strata that was Mission (California Department of Conservation, Divi-
explored in the 1980s and 1990s. sion of Oil and Gas, 1981). By the mid-1860s, oil dug in
pits or underground tunnels was being mined in large
operations, including those at McKittrick tar pits in
Oil (Tar) Sands Kern County and the Adams Canyon area of the Santa
Paula oil field, in Ventura County. More than 130 yr ago,
Hundreds of asphaltum and oil (tar) sands locations are excavations of more than 25 of these tunnels were done
in California (de Chadenedes, 1984, 1987). Most of the by companies that eventually merged to form Union
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 427

Oil Company. Tunnels followed the natural oil seep- erators, mostly fueled by natural gas, is injected under
ages in the area, and at many of these sites, water and oil high pressure into the heavy-oil reservoir. The steam
are still flowing to the surface. In the 1860s, the tunnels can be injected in either a continuous mode or in a cyclic
in the Adams Canyon area of the Santa Paula oil field basis (Figure 13). In continuous steam-flooding, steam is
produced more oil than any other production method pumped through injector wells into the heavy-oil res-
in California. At the time, the long distances to markets, ervoir. As the steam rises, eventually, it reaches a cap-
the relatively high operating costs, and economic fac- rock barrier and then begins to spread laterally at the top
tors meant that such operations were sporadic and of the reservoir. The steam warms the heavy oil in the
seasonal. As the oil fields were by Union Oil Company reservoir and provides drive toward a nearby vertical
(or by the companies that merged later to form Union production well. In CSS, the steam is injected for a
Oil), the seep oil production declined, along with a few months, followed by a brief (few months) soak
replacement of seep oil production by conventional time (to allow the heat to penetrate the formation,
drilling operations (California Department of Conser- heat the heavy oil, and lower the viscosity), and then the
vation, Division of Oil and Gas, 1981). heated oil is produced up the same well (Guerard, 1989;
Most of the technologies used to produce heavy oil Dusseault, 2013).
were originally developed in California. Guerard (1989) Disadvantages of the CSS and steam-flood processes
gives a comprehensive review of the history of the tech- are the costs associated with using natural gas to heat
nologies developed in California for extraction and pro- the water for the steam; water requirements for steam
duction of heavy oil. In the past, although California’s generation (Veil and Quinn, 2008); and heat losses
heavy-oil potential was enormous, it was not economical through surface lines, wellbores, and latent heat losses
to produce them by conventional means and as such, in to adjacent nonproducing rocks in the formation. Con-
the past, these deposits were not economically attractive. ventional steam injection using vertical wells has a
With the evolution of secondary and tertiary schemes, practical limitation of 914 to 1219 m (3000 – 4000 ft)
mostly thermal steam-flood and fire-flood (in-situ com- (Guerard, 1989). Recently, horizontal drilling has been
bustion), it became possible to develop these vast re- used to reduce well spacing in areas with urbanization
sources. As noted by Schamel (1998), recoveries in- or environmental concerns (Figure 10A, B) (Clarke and
creased from presteam values of less than 10% in the Phillips, 2003, 2004) and to target some of the more
Midway-Sunset field to up to 40 to 70% after steam- marginal areas for heavy-oil production. In pilot areas
injection projects were beyond the pilot stage of of California, some of the horizontal schemes are start-
development. ing to use SAGD to access the heavy-oil reservoirs, as is
To produce the heavy oil, the viscosity has to be re- more commonly done for the Athabasca oil sands of
duced, which is done mainly by either steam injection Canada (Hein et al., 2013). The advantage of SAGD with
or in-situ combustion. Steam produced in steam gen- horizontal drilling is that it may allow more uniform

Figure 13. Steam-flooding scheme


used in early development of the
California heavy-oil fields. Heat
from steam is either injected con-
tinuously or cyclically into the
heavy-oil reservoir. The injected
steam heats the reservoir, lower-
ing the viscosity of the oil, making
it easier for the steam to push
the oil through the formation to
the production wells (from Stosur
and Slater, 1987).
428 Hein

heating of the reservoir, the steam and the overlying Ongoing challenges for thermal stimulation of these
nonproductive formations have less contact, and latent heavy-oil reservoirs relate to steam-front or burn-front
heat losses may be less than for schemes that use conformance issues, wellbore breakthrough, cap-rock
vertical wells with steam-flooding. The difficulty of integrity, the use of water, greenhouse gas emissions,
SAGD is that many of the heavy-oil reservoirs in Cali- and other environmental concerns. Recently, one com-
fornia are very complex, with channelization, faulting, pany, GlassPoint Solar, has built the first solar-powered
and significant vertical and lateral barriers. This is in steam-flood (called Solar Pilot 1) for an oil field in Kern
contrast to the Canadian reservoirs that are more lat- County, California (Terrell, 2011). The solar pilot is a
erally continuous and have fewer lateral baffles and hybrid system of both natural gas and solar energy. The
barriers. solar pilot array is too far from the injection wells to
The final thermal process that has been used for provide steam directly; instead, the water is heated to
heating of heavy-oil reservoirs in California is in-situ subboiling temperatures and then fed to gas-fired steam
combustion, more commonly known as fire-flooding generators. This pilot has reduced gas usage by 20%, and
(Garon et al., 1984; Guerard, 1989) (Figure 14). Here, the hybrid system is below the market prices of natural
the heavy-oil (or oil-sand) reservoir is heated under- gas (Terrell, 2011). Several clean-steam and green-steam
ground, with initial ignition from a downhole-heating projects are being piloted in the San Joaquin Basin.
device. Air is then pumped downhole in the injection
wells, and a burning front then moves slowly through
the reservoir toward the producing wells. Lighter
fractions of crude oil ahead of the front are vaporized, RESERVE GROWTH OF HEAVY OIL, SEEPS, AND OIL
leaving a coke behind that is the principal fuel for the SANDS IN CALIFORNIA
in-situ combustion process. The rate of migration of
the burning front is dependent on the air injection In his review, Tennyson (2005) looked at the long-term
rate and the type and amount of crude oil being burned growth history of the 52 giant oil fields in California to
(Guerard, 1989). In some cases, water is injected simul- determine whether growth patterns are a function of
taneously or alternatively with the air to help transfer geologic or other characteristics of the fields, including
some of the heat to the oil ahead of the burn because of advances in technology. In California, most of the oil
the low heat-carrying capacity of the gases ahead of the fields had early discoveries, between 1890 and 1925, gen-
burning front. In all cases, once the heavy oil reaches the erally growing to significant volumes (>500 million bbl)
producing well, artificial lifts, including conventional within the first 20 yr after discovery. Nine giant fields in
rods, progressive cavity pumps, or more recently, elec- the San Joaquin Basin had significant stepped growth
trical submersible pumps, are used to lift the oil to the late in their life—the Midway-Sunset, Kern River, Elk
surface. Hills, South Belridge, and Coalinga. Fields in other

Figure 14. In-situ combustion


scheme used in early develop-
ment of the California heavy-oil
fields. Heat is used to reduce the
viscosity of the oil and permit it
to flow more easily to the pro-
duction wells. In a fire-flood, the
formation is ignited, and by con-
tinued production of air, the
burning front is advanced through
the reservoir (from Stosur and
Slater, 1987).
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 429

basins that also have significant late stepped growth using steam-floods from horizontal wells, which have
include the Ventura, San Ardo, Cat Canyon, and Wil- again significantly spiked the reserves (Figure 15A, B).
mington oil fields (Figure 15A, B). A further complexity in Midway-Sunset field is the
At the Midway-Sunset field, the oil gravities are checkerboard nature of the leases. Operators do not
mainly heavy (11 –148 API), with some extra heavy-oil always have contiguous acreage, and it is difficult to
(88 API). Reservoir sands have porosities of 30 to 35%, get benefits of thermal projects if offsetting producers
and permeabilities are from hundreds to thousands of are not steaming their reservoirs. Another heavy-oil
millidarcys. Most of the field is quite shallow, with a field that showed significant reserve growth related to
low natural drive (Lennon, 1976, 1990). The Midway- technological development is the South Belridge field
Sunset oil field is the largest of these fields that showed (Tennyson, 2005). The production at South Belridge is
late growth (Figure 15A). Pilot cyclic steam injection from two main reservoirs, the upper shallow Pliocene–
projects began in the early 1960s, which were so suc- Pleistocene heavy-oil sands of the Tulare Formation,
cessful that they were deployed throughout the field which contain heavy oil (10–148 API) and the lower
(Rintoul, 1995, 1999). Limited success with in-situ com- more intensely folded Belridge diatomites (Figure 5)
bustion or fire-flooding was done in 1970s, but the most with lighter, medium oil (20–328 API). Two advances in
significant reserves growth were related to the develop- unconventional recovery technology led to the increase
ment of steam-flooding in the 1960s and 1970s (Tennyson, in reserves growth of the South Belridge oil field. The
2005). Recently, since about 2000, operators have been first was related to the piloting of steam-flooding in the

Figure 15. Estimated ultimate


recovery over time for those giant
fields in California that showed
late-stage stepped growth, inter-
preted as reflecting advances in
in-situ technological develop-
ments. The Midway-Sunset, Kern
River, Elk Hills, South Belridge,
and Coalinga oil fields are in the
San Joaquin Basin; the Wilming-
ton oil field in the Los Angeles
Basin; and the Ventura, San Ardo,
and Cat Canyon oil fields in the
coastal California area (Figure 3)
(from Tennyson, 2005).
430 Hein

upper Pliocene–Pleistocene sands (which was successful has not been, in general, done on the many significant
and then done fieldwide); and the second was advances oil-sand deposits in the state (Table 1).
related to hydraulic fracturing of the deeper diatomite. Elsewhere, integration of LWD technology has im-
From its initial discovery in 1911 to 1950, the South proved unconventional shale gas production for shale
Belridge field reserves grew mainly by expansion and gas reservoirs through accurate well placement. The
the addition of new pools. In the early 1950s, an in-place Monterey Formation is actually similar to many of these
combustion pilot (fire-flood) operation showed that up other unconventional shale plays such as the Woodford
to 40 to 60% of the heavy oil could be recovered by in- Shale of southeastern Oklahoma, which has complex
situ thermal methods (Miller and McPherson, 1992). layering, faults, and extensive sealed and open frac-
Early CSS began in the 1960s, followed by the diatomite tures. Three-dimensional seismic and geologic analyses
fracturing technology in the 1970s, with the estimated are required to characterize the faults, dips, direction,
ultimate recovery tripling by 1990 to 1.1 billion bbl. Since and intensity of natural fractures for optimal lateral
the early 1990s, further infill drilling and expansion of orientation of horizontal wells and completions
steam-flood and water-flooding in the fractured diato- strategies.
mite have increased the expected ultimate recovery to 1.9 In-situ technologies, such as SAGD and CSS, are rou-
billion bbl (Tennyson, 2005) (Figure 15A). tinely used in Alberta to produce the vast oil-sands de-
The Wilmington field, in the Los Angeles Basin, con- posits in northern Canada (Hein et al., 2013). These
sists of weakly consolidated to unconsolidated, Plio- technologies have evolved to the extent that small-scale
cene deep-sea submarine fan turbidites. Oil gravities SAGD is being proposed for many of the smaller leases.
are variable from heavy (12–148API) in the shallower Expansion of this type of production to California may
pools to light (25–328API) in the deeper pools; reservoir be linked into existing infrastructures that are associ-
sands have porosities of 26 to 32% and permeabilities ated with heavy-oil production; in fact, the technologies
from 80 to 1600 md. Most of the field is quite shallow, are very similar to what has been done to rejuvenate the
with low natural drive. The Wilmington field grew by old Wilmington oil field horizontal drilling of multiple
the addition of pools for the first 20 yr, with an estimated wells from a single pad and creation of larger steam
ultimate recovery of 1 billion bbl. By the mid-1960s, the chambers in the subsurface that coalesce along the well
ultimate oil recoverability increased to about 3 billion pairs to give a better conformance to the scheme. In
bbl because of additions from offshore expansion and many areas of California, the oil sands, heavy oil, and
increased productivity associated with the fieldwide diatomites occur as separate play types within given
water-flooding schemes. In the late 1980s, steam-flooding fields or in close proximity to one another. These close
was introduced, and since 1988, both water-flooding associations of the diatomites and diatomaceous shales
and steam-flooding have produced most of the oil from (which are most amenable to hydraulic fracking) make
Wilmington (Tennyson, 2005). Recently, new technol- it possible to have full integration of the heavy-oil and
ogies, including horizontal drilling, LWD, and 3-D geo- oil-sand reservoirs, with deeper development by
logic modeling and visualization, along with continued multistage, multifracing (diatomites), and overlying or
secondary and tertiary recovery techniques, have con- laterally adjacent heavy oil and oil sands by SAGD, CSS
tinued to improve production efficiencies and extend with or without water-flood and steam-flood.
the production life of the field (Clarke, 1999; Clarke and At present, steam-enhanced heavy-oil production
Phillips, 2003, 2004). in California is in decline. Several other known Cali-
fornia heavy-oil or oil-sand deposits have never been
developed because it was initially thought that their
CONCLUSIONS AND FUTURE DEVELOPMENT viscosity was too high for steam or fire-flood to be ef-
fective. In-situ technologies such as SAGD or CSS are
Steam injection continues to be the favorite method for routinely used to develop the more highly viscous oil
thermal recovery of heavy oil from California. With in- sands of Canada. Recently, the use of SAGD in the Ox-
creased horizontal drilling, SAGD may be applied in nard field near Ventura, California, has been proposed.
combination with other thermal technologies, such as Other known heavy-oil and oil-sand deposits in Cali-
CSS and steam-flood, to maximize production through fornia, where such technologies may be used, include
technology sequencing in many of these old heavy-oil the large Foxen oil (tar) sand, Santa Maria Basin, and
reservoirs that are nearing the end of their production. the Arroyo Grande (Edna) in the Kern River area (among
Until now, most of the California oil sands have not been others, Table 1). Other prospective sources for Califor-
commercially developed. Although much of the in-situ nia’s future oil production include those offshore un-
steam technology was developed and applied in Cali- developed oil fields, such as those offshore the Santa
fornia to its heavy-oil reservoirs, to date this technology, Maria Basin, north of Point Conception, or undeveloped
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 431

fields north of the Channel Islands, San Miguel, Santa Beeson, D., and J. Singer, 2011, Surveillance well utility in
Rosa, and Santa Cruz. 4-D modeling of oil saturation changes at Kern River
In the past, in California, technological advances al- field, California (abs.): CD-ROM.
lowed oil production to continue when conventional Berman, B., and D. D. Clarke, 1987, THUMS oil operations,
in D. D. Clarke and C. P. Henderson, eds., Geologic guide
production declined by switching to heavy-oil produc-
to the Long Beach area: AAPG Pacific Section Guidebook
tion. Today, application of other technologies, such as 58, p. 99 – 127.
SAGD, may be applied to the common oil-sand and Beyer, L. A., 1987, Porosity of unconsolidated sand, diato-
seep deposits of California and, with technology se- mite, and fractured shale reservoirs, South Belridge and
quencing with heavy-oil production, may increase or at West Cat Canyon oil fields, California, in R. F. Meyer, ed.,
least maintain California’s oil production. These on- Exploration of heavy crude oil and natural bitumen:
shore resources, along with additions from offshore AAPG Studies in Geology 25, p. 395 – 414.
leases, may allow California to again become self- Biddle, K. T., 1991, The Los Angeles Basin: An overview, in
sufficient with its own strategic resources. K. T. Biddle, ed., Active basin margins: AAPG Memoir 52,
p. 5 – 24.
Blesener, J. A., and C. P. Henerson, 1996, New technologies
in the Long Beach Unit, in D. D. Clarke, G. E. Otott Jr.,
ACKNOWLEDGMENTS
and C. C. Phillips, eds., Old oil fields and new life: A visit
to the giants of the Los Angeles Basin: AAPG Pacific Sec-
The Energy Resources Conservation Board, Calgary, tion, p. 45 – 50.
provided technical assistance and support. I thank Bowersox, J. R., 2004, Late Neogene paleobathymetry, rel-
Kevin Parks, Doug Boyler, Jon Schwalbach, and Daniel ative sea level, and basin margin subsidence, northwest
Schwartz for helpful suggestions to improve the man- San Joaquin Basin, California: AAPG Search and Discov-
uscript and Dan Magee for digital graphics. Any use ery article 30029, 5 p.: http://www.searchanddiscovery
of trade, product, or firm names is for descriptive pur- .com/documents/2004/bowersox/index.htm (accessed
poses only and does not imply endorsement by the April 1, 2010).
author or the Alberta Government. Bridges, R. A., and J. W. Castle, 2003, Local and regional
tectonic control on sedimentology and stratigraphy in a
strike-slip basin: Miocene Temblor Formation of the
Coalinga area, California, U.S.A.: Sedimentary Geology,
REFERENCES CITED v. 158, p. 271 – 197, doi:10.1016/S0037-0738(02)00314-7.
Broderick, K. G., 2001, Controls on the location of hydrocar-
Alexander Jr., R. G., 1986, California and Saudi Arabia geo- bon seeps in the northern Santa Barbara Channel shelf,
logic contrasts, in M. T. Halbouty, ed., Future petroleum California: A geologic plumbing system (abs.): American
provinces of the world: AAPG Memoir 40, p. 291 – 315. Geophysical Union, Fall Meeting 2001, abstract OS31B-
Allan, M., M. Rahman, and B. Rycerski, 2006, Belridge giant oil 0436, 2 p.
field: Learnings from an unusual marine reservoir in an old California Department of Conservation, Division of oil and
field: AAPG Annual Convention, Houston, Texas, April 11: gas, 1981, Oil and gas production history in California:
AAPG Search and Discovery article 90052, http://www Sacramento, California, California Department of Conser-
.searchanddiscovery.com/documents/2006/06143allan vation, Division of Oil and Gas, 8 p.: ftp://ftp.consrv.ca
/index.htm; http://www.searchanddiscovery.com .gov/pub/oil/history (accessed April 1, 2010).
/documents/2006/06088houston_abs/abstracts/allan01 California Department of Conservation, Division of oil and
.htm (accessed April 1, 2010). gas, 1993, Petroleum seeps, geogenic sources, Section 9.2,
Allen, A. A., R. S. Schlueter, and P. G. Mikolaj, 1970, Natural Report, 5 p.
oil seepage at Coal Oil Point, Santa Barbara, California: California Department of Conservation, Division of oil, gas
Science, v. 170, no. 3961, p. 974 – 977, doi:10.1126/science and geothermal resources, 2006, Annual report of the state
.170.3961.974. oil and gas supervisor, 1994 through 2006 (December 31):
Arnold, R., and R. Anderson, 1910, Geology and oil resources California Department of Conservation, Sacramento, Cali-
of the Coalinga oil district, California: U.S. Geological Sur- fornia, 269 p.
vey Bulletin 398, 354 p. Clark, J. F., L. Washburn, J. S. Hornafius, and B. P. Luyendyk,
Bate, M. A., 1984, Temblor and Big Blue formations: Interpre- 1999, Distribution and source strength of dissolved hy-
tation of depositional environment sequence on Coalinga drocarbons from natural marine seeps, Coal Oil Point,
anticline, Fresno County, California: M.Sc. thesis, Stanford Santa Barbara, California (abs.): 1999 AAPG Pacific Sec-
University, Stanford, California, 123 p. tion Meeting, Monterey, California: AAPG Search and Dis-
Bate, M. A., and S. A. Graham, 1987, Significance of reservoir covery article 90920, http://www.searchanddiscovery.com
stratigraphy in production of heavy oil from Coalinga oil /abstracts/html/1999/pacific/abstracts/0681m.htm
field, in R. F. Meyer, ed., Exploration of heavy crude oil and (accessed April 1, 2010).
natural bitumen: AAPG Studies in Geology 25, p. 301–309. Clark, J. F., K. Schwager, and L. Washburn, 2005, Variability
432 Hein

of gas composition and flux intensity in natural marine and natural bitumen: AAPG Studies in Geology 25, p. 565–
hydrocarbon seeps: University of California Energy Insti- 570.
tute, Development and Technology, Paper EDT-008, 15 p. Demaison, G. J., 1977, Tar sands and supergiant oil fields:
Clark, M. S., L. F. Klonsky, and K. E. Tucker, 2001, Geologic AAPG Bulletin, v. 61, p. 1950 – 1961, reprinted, in E. A.
study and multiple 3-D surveys give clues to complex res- Beaumont, 1982, Energy minerals: AAPG Reprint Series
ervoir architecture of giant Coalinga oil field, San Joaquin 25, p. 218 – 229.
Valley, California: The Leading Edge, p. 744– 751, doi:10 Dibblee Jr., T. W., 1984, Geology and genesis of the Coast
.1190/1.1487285. Range province of California and its hydrocarbon de-
Clarke, D. D., 1999, At 68 Wilmington still has life: New posits, in R. F. Meyer, J. W. Earley, J. Barnea, and R. L.
technology revitalizes the old field: 1999 AAPG Pacific Johnston, eds., Exploration for heavy crude and bitumen:
Section Meeting, Monterey, California: AAPG Search and AAPG Research Conference, October 29–November, Santa
Discovery article 90920, http://www.searchanddiscovery Maria, California, 44 p.
.com/abstracts/html/1999/pacific/abstracts/0681p.htm Dibblee Jr., T. W., R. L. Johnston, J. W. Earley, and R. F.
(accessed April 1, 2010). Meyer, 1987, Field trip guidebook: Geology and hydrocar-
Clarke, D. D., and C. C. Phillips, 2003, Three-dimensional bon deposits of the Santa Maria, Cuyama, Taft-McKittrick,
geologic modeling and horizontal drilling bring more oil and Edna oil districts, Coast Ranges, California, in R. F.
out of the Wilmington oil field of southern California, in Meyer, ed., Exploration of heavy crude oil and natural bi-
T. R. Carr, E. P. Mason, and C. T. Feazel, eds., Horizontal tumen: AAPG Studies in Geology 25, p. 685 – 713.
wells: Focus on the reservoir: AAPG Methods in Explo- Dusseault, M. B., 2013, Screening criteria and technology
ration 14, p. 27 – 47. sequencing for in-situ viscous oil production, in F. J. Hein,
Clarke, D. D., and C. C. Phillips, 2004, Three-dimensional D. Leckie, S. Larter, and J. R. Suter, eds., Heavy-oil and oil-
geologic modeling and horizontal drilling bring more oil sand petroleum systems in Alberta and beyond: AAPG
out of the Wilmington oil field of southern California, in Studies in Geology 64, p. 655 – 668.
T. R. Carr, E. P. Mason, and C. T. Feazel, eds., Horizontal Eagan, J. M., M. M. Kolb, K. D. Knollenberg, and D. J.
wells: Focus on the reservoir: AAPG Search and Discovery Murphy, 1999, Three-dimensional surveillance in a ma-
article 20019, p. 27– 47: http://www.searchanddiscovery ture steam-flood Tulare Formation, South Belridge field,
.com/documents/2004/clarke/index.htm (accessed April 1, Kern County, California: 1999 AAPG Pacific Section Meet-
2010). ing, Monterey, California: AAPG Search and Discovery
Colvin, R. G., 1968, South Belridge oil field, in S. E. Karp, ed., article 90920, http://www.searchanddiscovery.com
Geology and oil fields, west side southern San Joaquin /abstracts/html/1999/pacific/abstracts/0681u.htm
Valley: AAPG — Society of Economic Geologists — (accessed April 1, 2010).
SEPM Guidebook, Pacific Section, p. 64 – 67. Earley, J. W., 1984, Economic factors in near-surface heavy
Croft, G. D., and T. W. Patzek, 2009, The future of California’s oil/tar sand mining, in R. F. Meyer, J. W. Earley, J. Barnea,
oil supply: Society of Petroleum Engineers Western Re- and R. L. Johnston, eds., Exploration for heavy crude and
gional Meeting, San Jose, California, March 24 – 26, SPE bitumen: AAPG Research Conference, Santa Maria, Cali-
Paper 120174-PP, 13 p. fornia, October 29 – November, 12 p.
Crowell, J. C., 1974a, Sedimentation along the San Andreas Fischer, P. J., 1978, Oil and tar seeps, Santa Barbara Basin,
fault, California, in R. H. Dott Jr. and R. H. Shaver, eds., California, in California State Lands Commission, Califor-
Modern and ancient geosynclinal sedimentation: SEPM nia offshore gas, oil and tar seeps, Sacramento, California,
(Society of Sedimentary Geology) Special Publication 19, p. 1 – 62.
p. 292 – 303. Garon, A. M., M. Komar, and G. C. Cala, 1984, Oxygen
Crowell, J. C., 1974b, Origin of Late Cenozoic in southern flooding: State of the art, in R. F. Meyer, J. W. Earley, J.
California, in W. R. Dickinson, ed., Tectonics and sedi- Barnea, and R. L. Johnston, eds., Exploration for heavy
mentation: SEPM (Society of Sedimentary Geology) Spe- crude and bitumen: AAPG Research Conference, Santa
cial Publication 22, p. 190 – 204. Maria, California, October 29 – November, 36 p.
Crowell, J. C., 1987, Coarse clastic facies of Miocene Ridge Gorsline, D. S., and K. O. Emery, 1959, Turbidity current de-
Basin adjacent to the San Gabriel fault, southern California: posits in San Pedro and Santa Monica basins off southern
Geological Society of America, Centennial Field Guide- California: Geological Society of America Bulletin, v. 70,
Cordilleran Section, p. 231 – 232. p. 279 – 290.
de Chadenedes, J. F., 1984, Geology and genesis of the Coast Guerard Jr., W. F., 1989, Heavy oil in California: California
Range province of California and its hydrocarbon depos- Department of Conservation, Division of Oil and Gas,
its: Surface tar sand deposits in California, in R. F. Meyer, Publication TR28, 12 p.
J. W. Earley, J. Barnea, and R. L. Johnston, eds., Explo- Gutierrez-Alonso, G., and M. R. Gross, 1997, Geometry of in-
ration for heavy crude and bitumen: AAPG Research Con- verted faults and related folds in the Monterey formation:
ference, Santa Maria, California, October 29– November, Implications for the structural evolution of the southern
15 p. Santa Maria Basin, California: Journal of Structural Ge-
de Chadenedes, J. F., 1987, Surface tar sand deposits in Cali- ology, v. 19, no. 10, p. 1303– 1321, doi:10.1016/S0191-8141
fornia, in R. F. Meyer, ed., Exploration of heavy crude oil (97)00039-4.
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 433

Hallmark, F. O., 1980, Unconventional petroleum resources of reservoir heterogeneity: Application to the character-
in California: California Department of Conservation, Di- ization of heavy oil reservoirs: U.S. Department of En-
vision of Oil and Gas, Sacramento, Publication TR25, 17 p. ergy Final Report, DE-FC26-00BC15301, 80 p.
Harding, T. P., 1976, Tectonic significance and hydrocarbon Ingle, J., 1981, Origin of Neogene diatomites around the
trapping consequences of sequential folding synchronous north Pacific rim, in R. E. Garrison and R. G. Douglas.,
with San Andreas faulting, San Joaquin Valley, California: eds., The Monterey Formation and related siliceous rocks
AAPG Bulletin, v. 60, p. 356 – 378. of California: Pacific Section of the SEPM (Society of Sed-
Hartman, B., and D. E. Hammond, 1981, The use of carbon imentary Geology), Research Symposium, Los Angeles,
and sulfur isotopes as correlation parameters for the source p. 159 – 179.
identification of beach tar in the southern California bor- International Energy Agency (IEA), 2010, World Energy
derland: Geochimica et Cosmochimica Acta, v. 45, p. 309– Outlook, Report (November), 50 p.
319, doi:10.1016/0016-7037(81)90241-6. Interstate Oil Compact Commission (IOCC), 1984, Major tar
Head, I. M., D. M. Jones, and S. R. Larter, 2003, Biological sand and heavy oil deposits of the United States: IOCC
activity in the deep subsurface and the origin of heavy oil: Report, p. 85 – 116.
Nature, v. 426, p. 344 – 352, doi:10.1038/nature02134. Isaacs, C. M., 1981a, Porosity reduction during diagenesis of
Hein, F. J., 2006, Heavy oil and oil (tar) sands in North Amer- the Monterey Formation, Santa Barbara coastal area, in
ica: An overview and summary of contributions: Natural R. E. Garrison and R. G. Douglas, eds., The Monterey For-
Resources Research, v. 15, no. 1, p. 67 – 84, doi:10.1007 mation and related siliceous rocks of California: Pacific
/s11053-006-9016-3. Section SEPM (Society for Sedimentary Geology) Special
Hein, F. J., and D. K. Higley, 2010, EMD oil sands commodity Publication 15, p. 257– 271.
report: AAPG Energy Minerals Division, 25 p., http://aapg Isaacs, C. M., 1981b, Guide to the Monterey Formation in the
/members-only/energy minerals division/oil sands/ California coastal area, Ventura to San Luis Obispo: AAPG
(accessed December 15, 2010). Pacific Section, 81 p.
Hein, F. J., D. Leckie, S. Larter, and J. R. Suter, 2013, Heavy Isaacs, C. M., 1987, Sources and deposition of organic matter
oil and bitumen petroleum systems in Alberta and in the Monterey Formation, south-central coastal basins
beyond: The future is nonconventional and the future is of California, in R. F. Meyer, ed., Exploration for heavy
now, in F. J. Hein, D. Leckie, S. Larter, and J. R. Suter, eds., crude oil and natural bitumen: AAPG Studies in Geology
Heavy-oil and oil-sand petroleum systems in Alberta and 25, p. 193 – 205.
beyond: AAPG Studies in Geology 64, p. 1 – 21. Jaramenko, D., 2011, Oil sands overseas: Oil Sands Review,
Heizer, R. F., 1940, Aboriginal use of bitumen by the Cali- v. 6, no. 5, p. 18 – 22.
fornia Indians, in Geological formations and economic Jeffrey, A. W. A., H. M. Alimi, and P. D. Jenden, 1991, Geo-
development of the oil land gas fields of California: State chemistry of Los Angeles Basin oil and gas systems, in
Division of Mines Bulletin, v. 118, p. 74. K. T. Biddle, ed., Active margin basins: AAPG Memoir 52,
Henderson, C. P., 1987, The stratigraphy of the Wilmington p. 197 – 219.
oil field, in D. D. Clarke and C. P. Henderson, eds., Geo- Jennings, C. W., 1957, Asphalt and bituminous rock, in L. A.
logic field guide to the Long Beach area, AAPG Pacific Wright, ed., Mineral commodities of California: Califor-
Section, p. 57 – 68. nia Division of Mines and Geology Bulletin 176, p. 5 – 22.
Herron, E. H., 2003, U.S. heavy oil and natural bitumen re- Karner, G. D., and J. F. Dewey, 1986, Rifting: Lithospheric
sources: Petroleum Equities Inc., 2 p.: www.petroleumequities versus crustal extension as applied to the Ridge Basin of
.321z.com (accessed December 15, 2010). southern California, in M. T. Halbouty, ed., Future petro-
Hills, L. V., ed., 1974, Oil sands fuel of the future: Canadian leum provinces of the world: AAPG Memoir 40, p. 317 –
Society of Petroleum Geologists Memoir 3, 263 p. 337.
Hodgson, S. F., 1980, Onshore oil and gas seeps in California: Kunitomi, D. S., S. P. Mulqueen, and B. H. Hesson, 2007, Oil
California Department of Conservation, Division of Oil on their shoes: Famous and little known oil seeps of Los
and Gas, Sacramento Publication TR26, 20 p. Angeles and Ventura counties: AAPG Pacific Section,
Hornafius, J. S., D. Quigley, and B. P. Luyendyk, 1999, The 80 p.
world’s most spectacular marine hydrocarbon seeps (Coal Kuuskraa, V. A., E. C. Hammershaimb, and M. Paque, 1987,
Oil Point, Santa Barbara Channel, California): Quantifica- Major tar sand and heavy-oil deposits of the United
tion of emissions: Journal of Geophysical Research, v. 104, States, in R. F. Meyer, ed., Exploration of heavy crude oil
no. C9, p. 20,703 – 20,711, doi:10.1029/1999JC900148. and natural bitumen: AAPG Studies in Geology 25, p. 123–
Hostettler, F. D., R. J. Rosenbauer, T. D. Lorenson, and J. 135.
Dougherty, 2004, Geochemical characterization of tarballs Kvenvolden, K. A., and C. K. Cooper, 2003, Natural seepage
on beaches along the California coast: Part I. Shallow of crude oil into the marine environment: Geo-Marine
seepage impacting the Santa Barbara Channel Islands, Letters, v. 23, no. 3 – 4, p. 140 – 146.
Santa Cruz, Santa Rosa and San Miguel: Organic Geo- Kvenvolden, K. A., F. D. Hostettler, R. J. Rosenbauer, T. D.
chemistry, v. 35, no. 6, p. 725–746, doi:10.1016/j.orggeochem Lorenson, and P. R. Carlson, 1999, Oil seeps as sources of
.2004.01.022. some coastal bitumen at the California shoreline (abs.):
Imhof, M. G., and J. W. Castle, 2004, Seismic determination 1999 AAPG Pacific Section Meeting, Monterey, California:
434 Hein

AAPG Search and Discovery article 90920, http://www national Conference, Caracas, Venezuela, February 7 – 17,
.searchanddiscovery.com/abstracts/html/1999/pacific 1982: New York City, New York: UNITAR/UNDP, 350 p.
/abstracts/0682y.htm (accessed April 1, 2010). Meyer, R. F., E. D. Attanasi, and P. A. Freeman, 2007, Heavy
Kvenvolden, K. A., F. D. Hostettler, T. D. Lorenson, and R. J. oil and natural bitumen resources in geological basins of
Rosenbauer, 2002a, Natural oil seeps and coastal tar in the world: U.S. Geological Survey, U.S. Geological Sur-
California: An extension of the Kaplan legacy (abs.): Geo- vey Open-File Report 2007-1084: Reston, Virginia, U.S. De-
logical Society of America, Annual Meeting, Denver, Pro- partment of the Interior, U.S. Geological Survey, 36 p.
gram with Abstracts, Paper 95-6, 1 p. Miller, D. D., and J. G. McPherson, 1992, South Belridge field,
Kvenvolden, K. A., W. S. Reeburgh, and T. D. Lorenson, 2002b, U.S.A., San Joaquin Basin, California, in E. A. Beaumont
Naturally occurring methane seepage (abs.): EOS, Trans- and N. H. Foster, eds., AAPG treatise of petroleum geol-
actions American Geophysical Union, v. 82, p. 457. ogy, atlas of oil and gas fields, structural traps III, tectonic
Lennon, R. B., 1976, Geological factors in steam-soak projects fold and fault traps, p. 221 – 241.
on the west side of the San Joaquin Basin: Journal of Pe- Minerals Management Service, 2005, Fingerprinting seeps in
troleum Technology, v. 28, no. 7, p. 741 – 748, doi:10.2118 the Santa Maria Basin: Natural seeps sampled: Newslet-
/5540-PA. ter, May-June 2005, p. 5.
Lennon, R. B., 1990, Midway-Sunset field, U.S.A., San Mulqueen, S. P., 2007, Petroleum seeps: Structural setting,
Joaquin Basin, California, in E. A. Beaumont and N. H. energy drive and path of migration, in D. S. Kunitomi,
Foster, eds., AAPG treatise of petroleum geology, atlas of S. P. Mulqueen, and B. H. Hesson, eds., Oil on their shoes:
oil and gas fields, structural traps III, tectonic fold and Famous and little known oil seeps of Los Angeles and
fault traps, p. 221 – 241. Ventura counties: AAPG Pacific Section, p. 1 – 26.
Lillis, P. G., and R. G. Stanley, 1999, Petroleum systems of Natural History Museum of Los Angeles County, 2005, Pit 91
the La Honda Basin, California: 1999 AAPG Pacific Sec- excavation, Page Museum, La Brea Tar Pits, 2 p.: www
tion Meeting, Monterey, California: AAPG Search and .tarpits.org/research/pit91/pitfacts.html (accessed April 1,
Discovery article 90920, http://www.searchanddiscovery 2010).
.com/abstracts/html/1999/pacific/abstracts/0683f.htm Norton, T. F., G. E. Otott Jr., 1996, The stratigraphy of the
(accessed April 1, 2010). Wilmington oil field: AAPG Pacific Section, Old oil fields
Lorenson, T., 2001, The ecology of oil seeps in central Cali- and new life: A visit to the giants of the Los Angeles
fornia, U.S. Geological Survey: Sound Waves Monthly Basin, p. 23 – 35.
Newsletter, May 2001, 2 p. Oakeshott, G. B., M. D. Turner, and C. W. Jennings, 1954,
Magoon, L. B., K. E. Peters, A. H. Scheirer, C. Lampe, D. L. Correlation chart of sedimentary formations in southern
Gautier, and P. G. Lillis, 2004, Modeling petroleum sys- California, in California Department of Natural Resources,
tems in the San Joaquin Basin, California (abs.): AAPG Division of Mines, Bulletin 170, v. 2, cont. 1, plate 1.
Proceedings Annual Meeting, Dallas, Texas, http://www Olsen, D. K., and E. D. Ramzel, 1995, State of heavy oil
.searchanddiscovery.com/abstracts/html/2004/annual production and refining in California: Cooperative Agree-
/abstracts/Magoon.htm (accessed April 1, 2010). ment DE-FC22-83FE60149, U.S. Department of Energy Re-
May, M. T., 2013, Oil-saturated Mississippian – Pennsylvanian port 1995-018, p. 159 –172.
sandstones of south-central Kentucky, in F. J. Hein, D. Orcutt, W. W., 1924, Early oil development in California:
Leckie, S. Larter, and J. R. Suter, eds., Heavy-oil and oil- AAPG Bulletin, v. 8, p. 61 – 72.
sand petroleum systems in Alberta and beyond: AAPG Pickering, K. T., R. N. Hiscott, and F. J. Hein, 1989, Deep
Studies in Geology 64, p. 373 – 405. marine environments: Clastic sedimentation and tecton-
Mayuga, A. N., 1970, Geology and development of Califor- ics: Unwin Hyman, London, 436 p.
nia’s giant Wilmington oil field, in M. T. Halbouty, ed., Quigley, D. C., J. S. Hornafius, B. P. Luyendyk, R. D. Francis,
Geology of giant petroleum fields: AAPG Memoir 14, J. Clark, and L. Washburn, 1999a, Decrease in natural
p. 158 – 184. marine hydrocarbon seepage near Coal Oil Point, Cali-
Meyer, R. F., 1987, Exploration of heavy crude oil and natural fornia, associated with offshore oil production: Geology,
bitumen: AAPG Studies in Geology 25, 535 p. v. 27, p. 1047– 1050, doi:10.1130/0091-7613(1999)027<1047
Meyer, R. F., and E. D. Attanasi, 2003, Heavy oil and natural :DINMHS>2.3.CO;2.
bitumen strategic petroleum resources: U.S. Geological Quigley, D. C., B. P. Luyendyk, J. S. Hornafius, R. D. Francis,
Survey, Fact Sheet 70-03, August 2003, online version 1.0: J. Clark, and L. Washburn, 1999b, Decrease in natural
http://pubs.usgs/fs/fs070-03/fs070-03.html (accessed hydrocarbon seepage offshore Coal Oil Point, California,
December 30, 2009). associated with oil production: 1999 AAPG Pacific Section
Meyer, R. F., J. W. Earley, J. Barnea, and R. L. Johnston, Meeting, Monterey, California: AAPG Search and Discov-
1984a, Exploration for heavy crude and bitumen: AAPG ery article 90920, http://www.searchanddiscovery.com
Research Conference, October 29 – November, Santa /abstracts/html/1999/pacific/abstracts/0683y.htm
Maria, California, 518 p. (accessed April 1, 2010).
Meyer, R. F., J. C. Wynn, and J. C. Olson, eds., 1984b, The Reid, S. A., and J. L. McIntyre, 2001, Monterey Formation
future of heavy crude and tar sands: United Nations In- porcelanite reservoirs of the Elk Hills field, Kern County,
stitute for Training and Research (UNITAR), 2nd Inter- California: AAPG Bulletin, v. 85, p. 69 – 189.
Overview of Heavy Oil, Seeps, and Oil (Tar) Sands, California 435

Requejo, A. G., and H. Halpern, 1989, An unusual hopane Tennyson, M. E., 1998, Estimates of ultimate recovery of oil
biodegradation sequence in tar sands from the Pt. Arena from California’s giant fields (>100 million barrels), 1923– Next
(Monterey) Formation: Nature, v. 342, p. 670 – 673, doi:10 1995: U.S. Geological Survey Open-File Report 98-513-A,
.1038/342670a0. 165 p.
Rintoul, B., 1995, Midway-Sunset keeps producing oil with a Tennyson, M. E., 2005, Growth history of oil reserves in major
little help from steam injection: Hart’s Oil and Gas World, California oil fields during the twentieth century, in T. S.
v. 87, no. 10, p. 101– 102. Dyman, J. W. Schmoker, and M. Verma, eds., Geologic,
Rintoul, B., 1999, Steam-flooding marked production break- engineering, and assessment studies of reserve growth:
through: Bakersfield Californian, April 27, 1999, 2 p. U.S. Geological Survey Bulletin 2172-H, 15 p.
Roadifer, R. E., 1987, Size distribution of the world’s largest Terrell, H., 2011, What’s new in production, solar power en-
known oil and tar accumulation, in R. F. Meyer, ed., hances recovery in old oil fields: World Oil (March), p. 19.
Exploration of heavy crude oil and natural bitumen: AAPG Truex, J. N., 1972, Fractured shale and basement reservoir,
Studies in Geology 25, p. 3–23. Long Beach Unit, California: AAPG Bulletin, v. 56, p. 1931–
Roberts III, W. H., 1987, Common conditions for heavy oils, 1938.
in R. F. Meyer, ed., Exploration of heavy crude oil and nat- Truex, J. N., 1974, Structural evolution of Wilmington, Cali-
ural bitumen: AAPG Studies in Geology 25, p. 453 –472. fornia, anticline: AAPG Bulletin, v. 58, p. 2398 – 2410.
Sanford, S. J., and N. A. Wildman, 1999, Use of 3-D model- U.S. Department of Energy, 2010, Fact sheet, U. S. heavy oil
ing and visualization tools to add value in California heavy resource potential, strategic unconventional fuels: U.S.
oil projects: Cymric and Coalinga fields: 1999 AAPG Pacific Department of Energy, Office of Petroleum Reserves, 1 p.
Section Meeting, Monterey, California: AAPG Search and U.S. Geological Survey, 1999a, The effect of seeps on the
Discovery article 90920, http://www.searchanddiscovery environment, 2 p.: http://seeps.wr.usgs.gov/seeps
.com/abstracts/html/1999/pacific/abstracts/0684d.htm /environment.html (accessed June 15, 2010).
(accessed April 1, 2010). U.S. Geological Survey, 1999b, Some examples of oil and gas
San Joaquin Geological Society, 2002a, The Kern County oil seeps, 2 p.: http://seeps.wr.usgs.gov/seeps/examples.html
industry, 2 p.: www.sjgs.com (accessed May 1, 2010). (accessed June 15, 2010).
San Joaquin Geological Society, 2002b, The history of the oil U.S. Geological Survey, 2000a, What are oil and gas seeps?,
industry (with emphasis on California and Kern County), 2 p.: http://seeps.wr.usgs.gov/seeps/what.html (accessed
12 p.: www.sjgs.com (accessed May 1, 2010). June 15, 2010).
San Joaquin Geological Society, 2005, Geology of the McKittrick U.S. Geological Survey, 2000b, Seeps and underground oil dis-
tar pits, 2 p.: www.sjgs.com/tarpits_geol.html (accessed coveries, 2 p.:, http://seeps.wr.usgs.gov/seeps/discoveries
May 1, 2010). .html (accessed June 15, 2010).
Schamel, S., 1998, Reactivation of an idle lease to increase U.S. Geological Survey, 2005, U.S. Geological Survey Web
heavy oil recovery through application of conventional site postings on oil and gas seeps, including detailed quad-
steam drive technology in a low dip slope and basin res- rangle maps: http://pubs.usgs.gov/seeps (accessed June
ervoir in the Midway-Sunset field, San Joaquin Basin, 15, 2010).
California: Quarterly Report to U.S. Department of En- Veil, J. A., and J. A. Quinn, 2008, Water issues associated
ergy, Report DE-FC22-95BC14937, 325 p. with heavy oil production: Argonne National Laboratory,
Schamel, S., 2013, Unconventional oil resources of the Uinta Environmental Sciences Division, Report ANL/EVS//R
Basin, Utah, in F. J. Hein, D. Leckie, S. Larter, and J. R. -08/4, 59 p.
Suter, eds., Heavy-oil and oil-sand petroleum systems Walters, E. J., 1974, Review of the world’s major oil sand de-
in Alberta and beyond: AAPG Studies in Geology 64, posits, in L. V. Hills, ed., Oil sands fuel of the future: Ca-
p. 437 – 480. nadian Society of Petroleum Geologists Memoir 3, p. 240–
Schwartz, D. E., 1988, Characterizing the lithology, petro- 263.
physical properties, and depositional setting of the Washburn, L., J. F. Clark, and P. Kyriakidis, 2005, The
Belridge diatomite, South Belridge field, Kern County, spatial scales, distribution, and intensity of natural ma-
California, in S. A. Graham, ed., Studies of the geology of rine hydrocarbon seeps near Coal Oil Point, California:
the San Joaquin Basin: SEPM Pacific Section, v. 60, p. 281 – Marine and Petroleum Geology, v. 22, p. 569 – 578, doi:10
301. .1016/j.marpetgeo.2004.08.006.
Slatt, R. M., S. Phillips, J. M. Boak, and M. B. Lagoe, 1993, Wilkinson, E. R., 1972, California offshore oil and gas seeps:
Scales of geologic heterogeneity of a deep-water sand giant California Department of Conservation, Division of Oil
oil field, Long Beach Unit, Wilmington field, California, and Gas, 11 p.
in E. G. Rhodes and T. F. Moslow, eds., Frontiers in sed- Yeates, R. S., and J. M. Beall, 1991, Stratigraphic controls on
imentary geology, marine clastic reservoirs, examples and oil fields in the Los Angeles Basin, in K. T. Biddle, ed.,
analogs: New York, Springer-Verlag, p. 263 – 292. Active margin basins: AAPG Memoir 52, p. 221 – 235.
Stosur, J. J. G., and J. A. Slater, 1987, Recovery methods for Zieglar, D. L., and J. K. Cassell, 1978, The synthesis of OCS
heavy oil and tar sands, in R. F. Meyer, ed., Exploration well information, offshore central and northern Califor-
of heavy crude oil and natural bitumen: AAPG Studies in nia, Oregon and Washington: Proceedings of the AAPG
Geology 25, p. 679 – 683. Pacific Section 53rd Annual Meeting, April 28 – 29, 27 p.
Next

You might also like