Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/290822084

Sedimentary Facies and Sedimentology of the Late Quaternary Santa Barbara


Basin, Site 893

Article · October 1995


DOI: 10.2973/odp.proc.sr.146-2.276.1995

CITATIONS READS

69 63

1 author:

Richard Behl
California State University, Long Beach
88 PUBLICATIONS   2,541 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Millennial-scale climate variability in the Pleistocene Pacific View project

Quantitative compositional characterization of the Miocene Lark Formation, Danish North Sea and Norwegian Margin View project

All content following this page was uploaded by Richard Behl on 12 April 2016.

The user has requested enhancement of the downloaded file.


Kennett, J.P., Baldauf. J.G., and Lyle, M. (Eds.), 1995
Proceedings of the Ocean Drilling Program, Scientific Results, Vol. 146 (Pt. 2)

22. SEDIMENTARY FACIES AND SEDIMENTOLOGY OF THE LATE QUATERNARY


SANTA BARBARA BASIN, SITE 8931

Richard J. Behl2

ABSTRACT

Distribution and character of sedimentary facies in a 196.5-m core from Ocean Drilling Program (ODP) Site 893 in Santa
Barbara Basin (southern California) reflect global and regional controls during the latest Quaternary. Stratigraphic trends in
grain size, sediment composition, sedimentary fabric, and degree of bioturbation record changes in oceanic deep-water circula-
tion, climate, sea level, basin geometry, sediment sources, and transport pathways through the last two glacial and interglacial
episodes (-160 k.y.). The sedimentary sequence in Hole 893A is divided into six facies: four record varied levels of bottom-
water oxygenation by the degree of bioturbation of primarily laminated olive-gray hemipelagic mud, whereas two reflect
coarse- and fine-grained event deposition. Varved sediments typical of the Holocene Santa Barbara Basin compose only 21%
of the entire latest Quaternary sequence.
An ultra-high-resolution record (<lO-yr sample spacing) of bottom-water oxygenation, as shown by the preservation or
destruction of annual varves, reveals cyclic variation over decadal to Milankovitch time scales and apparently reflects the
changing age and ventilation of Pacific Intermediate Waters. Over the past -75 k.y., bioturbation and basin oxygenation corre-
spond closely to glacial-interglacial oscillation, with warm intervals being oxygen-depleted and laminated and cold intervals
being oxygenated and bioturbated. Preliminary results suggest a correlation between basin stratigraphy and deep-water circula-
tion proxy records from elsewhere in the Pacific and Atlantic oceans, indicating that sediments of Santa Barbara Basin are a
sensitive recorder of global and regional paleoceanographic and paleoclimatic change.

INTRODUCTION
For the past four decades, varved sediments from the central part
of Santa Barbara Basin (SBB) of the Southern California Continental
Borderland have yielded ultra-high-resolution paleoceanographic
and paleoclimatic records based on variation in fish (sardine) abun-
dances (Soutar, 1967; Soutar and Isaacs, 1969; Baumgartner et al.,
1992),planktonicmicrofossils(Pisias 1978, 1979; Langeet al., 1987,
1990), pollen (Heusser, 1978), stable isotopes (Dunbar, 1983; Schim-
melmann and Tegner, 1991), biomarkers (Lajat et al., 1990; Kennedy
and Brassell, 1992), and varve thickness (Soutar and Crill, 1977). For
a review of previous work in SBB, see Kennett, Baldauf, et al. (1994)
or Lange et al. (in press). All these prior studies were derived from
shallow-penetration box or piston cores, and were restricted to the
Holocene (generally the last several hundred years). It remained un-
known whether the bathymetric and oceanographic conditions re-
quired for preservation of the detailed, varved record were unique to
the Holocene or typical of the entire late Quaternary.
To answer this question, two advanced hydraulic piston cores
were obtained from near the basin center (34°17.25'N, 120°02.2'W) Figure 1. Bathymetric map and setting of Santa Barbara Basin and Site 893.
(Fig. 1; Kennett, Baldauf, et al., 1994), recovering 196.5 m (Hole Bathymetry in meters.
893A) and 68.8 m (Hole 893B) of intermittently laminated upper
Quaternary sediments. Micropaleontologic and stable isotope stratig-
raphy indicate that Hole 893A extends into oxygen isotope Stage 6, promise of extending the high-resolution chronostratigraphy beyond
approximately 160 ka (Kennett, this volume). Accelerator mass spec- the range accessible by radiocarbon dating.
trometry (AMS) radiocarbon dates to approximately 28,000 calendar For the most part, the cored material was recovered in excellent
years before present were obtained from the upper 43 m (Ingram and condition except for irregularly distributed bedding-plane partings
Kennett, this volume), providing a high-resolution chronostratigra- produced by degassing and expansion of the methane-rich sediment,
phy for the last glacial/interglacial transition. Correlation of the mag- contributing to a total recovery of 104% to 107% of the length of the
netic secular variation record (Kennett, Baldauf, et al., 1994) offers cored sequence (Kennett, Baldauf, et al., 1994). The parting can in-
terfere with some in-situ measurements of physical properties that re-
quire undisturbed fabrics over several centimeters, such as bulk
density or magnetic susceptibility, however, it is not detrimental for
'Kennett. J.P., Baldauf, J.G., and Lyle, M. (Eds.), 1995. Proc. ODP, Sci. Results,
146 (Pt. 2): College Station, TX (Ocean Drilling Program).
most stratigraphic studies of a compositional nature or for samples
2
Marine Science Institute and Department of Geological Sciences, University of specifically selected for microfabric analysis. All stratigraphic posi-
California, Santa Barbara, CA 93106-6150, U.S.A. tions reported herein are corrected to account for the gaps formed by
R..I. BEHL

sediment degassing and are based on observations and analyses of the and mid-water suspensate plumes (Figs. 1 and 9; Drake et al., 1972;
196.5-m core from Hole 893A. Fleischer, 1972; Thornton, 1981; Stein, this volume); (2) coarse-
The high sedimentation rate (-120 cm/k.y.), organic-rich varved grained siliciclastic sediments transported eastward into the basin by
mud, proximal setting recording both oceanic and continental sig- coastal longshore drift (Trask, 1952; Emery, 1960), and then, in part,
nals, and the close relationship with the California Current system deflected seaward to the middle to outer shelf and returned to the west
make Site 893 extremely important for paleoceanographic and pale- by strong bottom currents (Kolpack, 1986); and (3) biogenic compo-
oclimatic studies, many of which are presented in this volume (e.g., nents (e.g., diatoms, radiolarians, and foraminifers) derived from
Kennett; Kennett and Ingram; Heusser). This paper presents the sed- overlying surface waters and, in part, benthic organisms, chiefly for-
imentology and facies distribution of Hole 893A to provide a frame- aminifers, echinoderms, and sponges. Slope deposits, covering 96%
work for paleoceanographic and paleoclimatic time series, and of the basin deeper that the shelf break, are redistributed by mass
additionally explores the effects of paleoceanographic and paleocli- flows, including sediment creep, slides, and slumps (Thornton,
matic change on sedimentation in Santa Barbara Basin. 1984). Short-term deployment of sediment traps suggests that most
of the flux to the sediment can be accounted for by biopackaging, ei-
ther as fecal pellets (Dunbar and Berger, 1981) or marine snow (All-
SETTING OF THE MODERN SANTA dredge and Gottschalk, 1989).
BARBARA BASIN
Santa Barbara Basin is the northernmost basin of the southern METHODS
Californian Continental Borderland (Emery, 1960; Gorsline and
Teng, 1989), bounded by the mainland on the north and the San Sedimentologic and stratigraphic observations for this study were
Miguel-Anacapa island chain (known as the northern Channel Is- made from three distinct data sets. The stratigraphic distribution of
lands) to the south. The east-west-trending, ~600-m-deep basin opens facies was determined by a detailed visual examination of the 196.5-
over a deep sill to the west (-475 m depth) and a shallower sill to the m core from Hole 893A, locally supplemented by examination of X-
east (-230 m depth). Sediment supply to the basin is highly seasonal, radiographs. The face of the archive half of the split core was careful-
with the bulk of siliciclastic sedimentation derived from riverine run- ly scraped with a negatively charged blade to reveal details of even
off during winter and early spring rainstorms (Soutar and Crill, 1977; subtle or obscured sedimentary structures (Kennett, Baldauf, et al.,
Thornton, 1981, 1984). Consequently, planktonic sedimentation dur- 1994). Sedimentary facies were identified and logged at 0.5-cm (<5-
ing the late spring, summer, and fall is largely undiluted by terrige- yr) resolution (see data report; Behl, this volume). Although analysis
nous material. Below sill depth, the modern basin contains oxygen- of X-radiographs would allow more subtle features to be detected,
depleted waters which derive from the upper part of the Pacific Inter- only limited selected intervals of the core were imaged at Texas
mediate Water mass and the Oxygen Minimum Zone (OMZ) off cen- A&M University and Scripps Institution of Oceanography.
tral California. Once within the basin, this low-oxygen water mass is Textural analysis of fine-grained sediments was performed using
further depleted of oxygen by degradation of organic matter from the a SediGraph 5100. All samples were treated with 15% H2O2 and 1%
highly productive surface waters, rendering the seafloor inhospitable HC1 to remove organic matter and carbonate, dispersed in 0.1%
to benthic faunas and preventing bioturbation (Emery and Hülse- Na6(PO_,)6, sieved to remove the >63-µm fraction, then redispersed in
mann, 1962). In cores from the basin center (below -525 m water 0.1 % Na6(PO3)6 before analysis. The SediGraph uses attenuation of
depth), the seasonal variations in sedimentation (siliciclastic-domi- X-rays to determine the settling rate of dispersed particles; therefore,
nated vs. biogeneous-dominated) are preserved as thin, millimeter- it measures equivalent spherical sedimentation diameters (ESSD),
scale laminations, or varves. At the sediment-water interface, sulfate- rather than intercepted diameter of the particles, cross-sectional area,
reducing, filamentous bacteria form extensive, stabilizing mats, be- or volume. As these sediments were deposited in a marine environ-
low which continued bacterial oxidation of organic material renders ment controlled largely by hydrodynamic conditions, ESSD was se-
shallow pore waters anoxic. Thickness of the varves correlates close- lected as the most meaningful measurement of grain size. Most of the
ly with historic and tree-ring records of precipitation (Schimmel- hemipelagic sediments have large fine-grained tails of grains smaller
mann and Tegner, 1991; Soutar and Crill, 1977). Although there is than can be reliably measured (1 µm or 10 Φ); therefore, particle sort-
evidence for short-term (subseasonal) disruption of the basin's strat- ing is calculated by the log quartile deviation (Krumbein and Petti-
ification (Sholkovitz and Gieskes, 1971; Reimers et al., 1990), sub- john, 1938).
oxic conditions apparently prevail throughout most of the year, Compositional data were collected from smear slides of the >63-
preserving primary laminations and other high-frequency paleocean- µm (sand-sized) fraction separated from fine-grained sediments prior
ographic and paleoclimatic signals. In the last -2000 yr, only one ep- to SediGraph measurement of grain size. This fraction typically
isode of prolonged oxygenation of the basin floor is known to have makes up <2% of the hemipelagic mud samples and defines a distinct
occurred (Schimmelmann et al., 1992). compositional population from that of the silt and clay portion. At
The spatial distribution of sedimentary fabrics throughout SBB is least 300 grains were counted during line-count transects of the slides
directly controlled by the dissolved oxygen content of the overlying with frequency presented in number % of intersections (not weight %
bottom water (Savrda et al., 1984). In core-top samples, continuously or area %). The samples were previously cleaned of organic matter
laminated sediment occurs only in the central basin within the anaer- and carbonate as described above; therefore, these data are presented
obic zone where bottom waters contain <O. 1 mL/L of dissolved oxy- on a carbonate-free basis. An additional measurement—the presence
gen (O2). Fringing the laminated sediments is a narrow zone or absence of echinoderm spines—was made by Ingrid Hendy (UC
containing both varved and bioturbated layers, interbedded on a scale Santa Barbara) on washed, disaggregated samples untreated for car-
of centimeters; this facies lies beneath waters containing -0.1 mL/L bonate removal.
O2and likely results from short-term fluctuation in the position of the
dysaerobic-anaerobic boundary (at 0.1 mL/L O2). Completely mas- SEDIMENTARY FACIES
sive (bioturbated) sediment occurs only beneath waters with >O. 1-
0.2 mL/L O2. Observations
The modern SBB contains no significant active submarine fans or With the exception of infrequent, thin sand layers, the SBB core
deltas to control sediment dispersal (Thornton, 1984). Instead, the ba- is composed entirely of hemipelagic mud consisting of terrigenous
sin is fed from three separate sources: (1) fine-grained siliciclastics silty clay to clayey silt (80%-90%) with minor amounts (generally
chiefly derived from the Santa Clara and Ventura rivers as surface <10% each) of diatoms and nannofossils (bulk smear-slide analysis
SEDIMENTARY FACIES AND SEDIMENTOLOGY

B C
Facies 1: Fades 2: Facies 3: Facies 4:
Well-laminated hemipelagic mud Indistinctly laminated hemipelagic mud Trace-laminated hemipelagic mud Massive hemipelagic mud

} \
(/ ^iliillifliiii
to

oo

Facies 5: Gray-layer fine-grained event deposits Facies 6:


Homogeneous suspensate deposits Graded clayey silt turbidites Sand turbidite

1 1 cm

1 cm

Figure 2. Representations of characteristic features of sedimentary facies at Site 893. Graphic scale is approximate.

in Kennett, Baldauf, et al., 1994). Trace amounts (<3%) of foramini- tions with diffuse stratigraphic contacts or by (b) an interval contain-
fers, radiolarians, sponge spicules, echinoderm spines, and framboi- ing only a few discrete patches of laminations surrounded by
dal pyrite are consistently present in smear slides. Amorphous homogeneous sediment (Fig. 2C). In addition to the homogenized re-
organic matter is present throughout most of the core (1-4 weight gions, small, distinct 2- to 6-mm-diameter burrows {Chondritesl) are
percent [wt%] total organic carbon; see Gardner and Dartnell, this locally present. These are primarily expressed as elongate horizontal
volume; Stein and Rack, this volume). to inclined ellipses (up to 50° from horizontal). The vertical extent of
The sequence from Hole 893A was described at a subcentimeter the burrows is typically <l cm. Facies 2 and 3 locally contain
scale and six sedimentary facies are identified, including both hemi- "blooms" of benthic foraminifers in far greater abundance (>2 tests/
pelagic and event deposits (Fig. 2). These facies are interbedded over cm2 visible on the scraped core face) than observed in Facies 1. These
a range of scales, from <l cm to >2 m. blooms are most common in centimeter-thick layers located immedi-
Facies 1-4 define a gradual progression from well-laminated to ately above well-laminated sediment (Facies 1).
massive (bioturbated) hemipelagic mud; all are chiefly olive-gray Facies 4: massive hemipelagic mud. Facies 4 is composed of ol-
(5Y 4/2) with minor variation in value. The hemipelagic Facies 1-4 ive-gray mud containing no discernible primary fabric. Lower
are all texturally similar, composed of 36 to 54 wt% silt (43% mean), boundaries of Facies 4 layers are locally discordant and irregular on
with median grain sizes ranging from 2.2 to 5.5 µm (Fig. 3). Sorting, a <l-centimeter scale. These massive intervals are commonly mot-
as shown by the log quartile deviation, is very poor, ranging from tled by texture or color and locally display discrete burrows 3-10 mm
0.49 to 0.62 with a mean value of 0.53. thick and up to 30 mm long. These burrows are most visible at layer
Facies 1: well-laminated hemipelagic mud. Facies 1 is composed boundaries where strong contrasts in sediment color exist (Fig. 2D).
of unbioturbated, fine-grained sediment displaying distinct, continu- Some burrows originating in massive layers show vertical branching
ous laminations across the central part of the core (Fig. 2A). Individ- and extend 1-3 cm in total length. Internal structure of the burrows is
ual laminations are generally <l.O to 2.0 mm thick and arranged as not clear when observed on the surface of the core.
couplets of dark terrigenous-rich (winter) and light diatom-rich Facies 5: gray layer deposits. Facies 5 is characterized by the dis-
(spring-summer) layers (Soutar and Crill, 1977). tinct lack of olive coloration and a decided gray appearance (5 Y 5/1).
Facies 2: indistinctly laminated hemipelagic mud. Facies 2 is Gray layers typically have sharp bases and often display sharp tops.
identified by the presence of diffuse, discontinuous or irregular lam- Gray layers are locally micaceous or silty, especially at their bases;
inations (Fig. 2B). Individual laminations are for the most part thick- many are normally graded in texture or in color (dark to light) while
er than in Facies 1, typically ranging from 1.5 to 4 mm thick. Distinct others are massive. Gray layers range from a few millimeters to 20
burrows are extremely rare; when present they appear as 1- to 3-mm- cm in thickness; some units are amalgamations of up to four distinct
thick, structureless ellipses that are principally oriented in the hori- graded deposits. Although gray layers display a broader range of
zontal direction {Chondritesl). grain sizes than the hemipelagic facies (1.8 to 6.5 µm median grain
Facies 3: hemipelagic mud with traces of laminations. Facies 3 is size, 29% to 60% silt), their mean values are similar (3.2 µm, 42%
identified by (a) olive-gray mud displaying extremely faint lamina- silt). Most gray layers are better sorted than the hemipelagic facies

297
R.J. BEHL

volume). The sand is unconsolidated and uncemented with individual


Grain diameter (µm) deposits ranging from millimeters to >2 m in thickness. The thicker
1 10 1C beds are typically disturbed by the HPC coring process, yet some dis-
b i i i i i i •i| l play normal size grading. Thinner deposits (<2 cm) are generally
— A well-preserved, display sharp bases and locally show gradational or

••
4 massive mud tops (Fig. 2F). A few sand layers are planar laminated.
3 -
Facies Discussion
2 —
In general, the diameter and depth of penetration of macrofaunal
1 4 burrows correlates with the oxygenation level of overlying bottom
* •• waters (Pearson and Rosenberg, 1978; Savrda et al., 1984); that is,
I 1
the greater the oxygenation, the greater and deeper the biological
b I I 1 ll| 1 1 I I I I ll|
mixing. In this study, I assume that the relation between oxygen lev-
B _ els and burrow size extends to smaller, meiofaunal and microfaunal
4
benthic organisms (e.g., foraminifers, nematodes, etc.) that are capa-
3 ble of sublamination-scale burrowing. The distinction between dif-
2 -
• ferent lamination styles and degree of bioturbation in Facies 1-4 in
this study serves as a "bioturbation index," which reflects the dis-
1

•••
solved oxygen content of bottom water during or shortly after depo-
n I i i ill i I i i i i ill sition (Byers, 1977; Savrda etal., 1984; Droser and Bottjer, 1986). In
b I l 111| i i i i i 1111 i 1 1 1 1 1 11 the modern central SBB, Savrda et al. (1984) found laminated sedi-
ments under anaerobic conditions (<O. 1 mL/L O2), homogenized
4 - • c _ (bioturbated) sediments within dysaerobic or aerobic environments
3 - - (>O. 1-0.2 mL/L O2), and interbedded laminated and homogeneous
sediment layers at the anaerobic-dysaerobic boundary. Some lami-
2 - nated sediments adjacent to the anaerobic-dysaerobic boundary dis-
play a few cross-cutting burrows, but most of the primary fabric is
1
• preserved. These relationships are the basis of the following interpre-
I i i i
0 tations:
5 i i i 111 i i i i i i • 11 1 1 1 1 1 1 1 Facies 1 (well-laminated hemipelagic mud) was clearly deposited
D _

_-•
• ••
4 - • • •
* m within anaerobic conditions (<O.l mL/L O2) that effectively excluded
burrowing macrofauna capable of disturbing the fine primary lamina-
3 * • - tion. Although distinct burrows are rare in Facies 2 (indistinctly lam-
inated hemipelagic mud), it is reasonable that the diffuse and
2
•4 irregular laminations reflect a minor degree of microfaunal to meio-
1 faunal bioturbation of the distinct varves of Facies 1. In Facies 2, the
1 i i i i i i i l l very small-scale bioturbation was only sufficient to broaden and dis-
0 turb the primary laminations, but not sufficient to obliterate them.
/ i i i i 1 1 1 1 1 1 1

•'JJJH.
Facies 3, containing small burrows and either (a) faint, broad, dif-
E -
1

6 • fuse laminations or (b) isolated patches of laminated fabric surround-


1

5 ed by structureless mud, records an increased degree of infaunal



• mixing. Bioturbation was sufficient to homogenize part of the sedi-
1

4
ment (vertically or laterally), yet insufficient to obliterate evidence of
1

3 •

the primary pattern of seasonal sedimentation. This facies was likely



2 •
f

1 • formed close to the anaerobic-dysaerobic boundary, which Savrda et


i i i MI 1 al. (1984) place beneath waters containing approximately 0.1 mL/L
0
1 10 100 O 2 . In particular, Facies 3(b) probably resulted from temporal chang-
es in bottom-water oxygenation in which originally laminated sedi-
Figure 3. Representative grain-size spectra for Facies 1-5. A. Facies 1: well- ment was subsequently partially homogenized by downward
laminated hemipelagic mud (Sample 146-893A-3H-6, 29-30 cm). B. Facies penetration of burrows during an episode of oxygenation. The unusu-
2: indistinctly laminated hemipelagic mud (Sample 146-893A-21H-4, 29-30 al blooms of benthic foraminifers that locally occur at the bases of Fa-
cm). C. Facies 3: hemipelagic mud with traces of laminations (Sample 146- cies 2, 3 or 4 layers where found directly above well-laminated
893A-14H-3, 29-30 cm). D. Facies 4: massive hemipelagic mud (Sample sediment (Facies 1) may reflect the sudden availability of previously
146-893A-14H-7, 25-26 cm). E. Facies 5: gray layer event deposits (Sample ungrazed, organic matter-rich laminated sediments with regression of
I46-893A-3H-2, 24-25 cm). Note better sorting of (E) (event deposit) com- the anaerobic-dysaerobic boundary and the corresponding increase in
pared with (A-D) (hemipelagic mud facies). Vertical axes show mass percent the dissolved oxygen content of bottom water (see Savrda and Bottjer
in each 1/4 Φ interval. [ 1987] for a similar stratigraphic occurrence of shelled macrofauna in
Miocene rocks).
(0.45 to 0.51 log quartile deviation, 0.48 mean) and their grain size Grimm et al. (unpubl. data) suggest that diffuse lamination, as
distributions (i.e., narrow distribution with minor fine-grained tails) found in Facies 3(a), may be a primary sedimentary feature reflecting
are typical of sediment that settled as individual particles according lower seasonality than known for the late Holocene. It is unlikely,
to Stoke's Law rather than as aggregates or floes (Fig. 3E; Kranck, however, that hemipelagic sedimentation over tens to hundreds of
1984). years would remain only subtly changed in composition and texture
Facies 6: sand deposits. Facies 6 is composed of very fine to fine- from season-to-season and year-to-year, without any distinctive sea-
grained quartzose to quartzofeldspathic sand (Marsaglia et al., this sonal plankton blooms or rainfall run-off deposits. For this reason, I
SEDIMENTARY FACIES AND SEDIMENTOLOGY

Trace-laminated hemipelagic mud (4%) Increased bioturbation or oxygenation


Indistinctly laminated hemipelagic mud (7%)

Well-laminated hemipelagic mud (10%)

Sand turbidites (4%)


lie mug Gray-layer event deposits (4%)
nFaci 4)

Figure 4. Pie chart showing relative stratigraphic abundances of sedimentary


facies at Hole 893A.

find it less astonishing to assume that faint, diffuse laminations result


from synsedimentary, millimeter-scale horizontal bioturbation of
originally distinct seasonal laminations.
Due to the lack of primary structures and the presence of distinct
burrows, and the gradational character of Facies I through 4, Facies
4 (massive olive gray mud) is considered to result from synsedimen-
tary or postdepositional biological homogenization beneath waters
with >O. 1-0.2 mL/L O2. Common mottling, distinct burrows, and the
discordant and irregular character of the lower boundaries of many 100
Facies 4 units attest to their origin by bioturbation. Complete homo-
genization generally mixes <3 cm deep as indicated by the presence Φ
of numerous thin, massive layers intercalated between laminated in-
tervals and by the maximum vertical extent of observed burrows. An
undetermined portion of this massive facies, however, may be due to
event deposition of remobilized slope sediment (Thornton, 1986)
with texture and composition identical or indistinguishable from that
of the basinal facies. Some redeposited sediment can be identified by
subtle compositional grading or the presence of thin, mica-rich lay-
ers. 150
Gray layers (Facies 5) are moderately well sorted, commonly
graded, sharp-based, and have a color and mineralogic composition
that is distinct from the olive-green hemipelagic mud (Stein, this vol-
ume). Consequently, gray layers are interpreted as fine-grained event
deposits. Thornton (1984) interpreted the massive and graded gray
layers as flood suspensate deposits from major storms and graded
"silty" turbidites, respectively. The sand deposits (Facies 6) are also
sharp-based, locally graded and intercalated with fine-grained hemi-
pelagic muds. They are interpreted as coarse-grained turbidites.
Overall, laminated hemipelagic mud (Facies 1-3) makes up 21% 200
of the total stratigraphic thickness of the core, massive olive-gray
mud (Facies 4) accounts for 70%, and the event deposits (Facies 5 vyelt
and 6) account for the remainder (approximately 4% each) (Fig. 4). laminated
This percentage distribution of facies is markedly different from that
previously reported from the Holocene section, which is composed of Figure 5. Degree of bioturbation ("bioturbation index") shown as a 99-cm
70% laminated olive-gray mud, 4% massive olive-gray mud, and running average of 1-cm observations vs. corrected depth downcore. Major
26% "gray layer" event deposits (Thornton, 1984). This distribution lithostratigraphic subunits are presented at left (after Kennett, Baldauf, et al.,
shows that anaerobic conditions responsible for preservation of lam- 1994); Oxygen isotope stages presented at right (Kennett, this volume).
inated sediment in the Holocene SBB were typical of only a portion
of the last 160 k.y.

part of oxygen isotope Stage 6 (OIS 6); (IE) a lower nonlaminated se-
STRATIGRAPHIC TRENDS quence corresponding to Termination II and OIS 5e (Eemian); (ID) a
Facies and Textural Trends lower laminated sequence corresponding to OIS 5e to 5c; (IC) a thick
upper intermittently laminated sequence spanning OIS 5c through
Hemipelagic mud OIS 3; (IB) an upper nonlaminated sequence corresponding to OIS 2;
and (IA) an upper laminated sequence which correlates with most of
The shore-based scientific party (Kennett, Baldauf, et al., 1994) Termination I and OIS 1. In the upper 100 m (last 75 k.y.) the broad
initially defined six lithostratigraphic subunits based on the large- pattern of laminations is largely synchronized with glacial-intergla-
scale distribution of laminated sediments (Fig. 5); these units have cial changes (Fig. 6), with laminated sediments associated with warm
been placed in a chronostratigraphic framework by AMS I4C dating intervals and massive sediments with cool to cold intervals (see Ken-
and oxygen isotope stratigraphy (Ingram and Kennett, this volume; nett, this volume, and Kennett and Ingram, this volume). This corre-
Kennett, this volume). The lithostratigraphic subunits include: (IF) a lation between SBB laminations and the glacial-interglacial record
lower intermittently laminated sequence corresponding to the latter breaks for the lower half of the core, suggesting that the interaction
R.J. BEHL

Bioturbation index Oxygen isotopes most gray layers are 0.5-2 cm thick and bioturbation within massive
intervals probably mixes intervals as large as 3 cm thick. Gray layers
1 I
1 that do occur within otherwise massive sequences are thicker and
—^< — generally have diffuse, gradational tops and, locally, gradational
- " ^ mm
m

^― - bases. In some locations, a subtle graying of massive olive-gray mud


2 is the only remaining indication of a bioturbated gray layer. Digital
color analysis quantitatively records this color shift (Merrill and
>
< Beck, this volume), and integration of the deviation from the mean
3 -
value for the massive hemipelagic mud may allow quantification of
the volumetric contribution of gray event deposits, even in heavily
> bioturbated sediments.
- Herein, the stratigraphic distribution of gray layers is plotted in
4
Figure 8 as the running average of their percent thickness (i.e., centi-
< meters of gray layers/meter of core). In addition to the association
- 5a - with laminated intervals, gray layers are most frequent during periods
of relatively high sea level, for example, throughout OIS 5 (Fig. 8).
- 5b -
This relationship suggests that deposition of even the clearly turbid-
itic deposits are related to major flood events or other changes in sed-
5c iment supply rather than to off-shelf transport during sea-level falls.
- As basinal SBB sediments are chiefly derived from the Santa Clara
5c '

and Ventura rivers at the eastern boundary of the depocenter (Fleis-

cher, 1972; Thornton, 1981, 1984; Stein, this volume; Marsaglia et
——
H
al., this volume), the highstand predominance of gray layers likely re-
flects the effect of changing basin geometry on the transport of flood
6 -
> suspensates (Fig. 9). With falling sea level, the eastern passage into
the basin narrowed, reducing the transport of fine-grained suspen-
>» sates into the basin from the Santa Clara and Ventura rivers. Trans-
160 - port of flood-related sediment into SBB may have taken significantly
i I 1 longer at lower sea levels than the ~2 yr required for flood deposits
4 3 2 1 4 3 to be completely deposited in the central basin during the recent times
Faciθs δ18O%oPDB (Drake et al., 1972). Slower or diminished transport would spread out
Massive < > Laminated the climatic signal to such a degree that the input pulse might be
Oxygenated Anoxic "washed out" by typical hemipelagic sedimentation. Decreased flux
into the basin at lowstands did not result in decreased sediment accu-
Figure 6. Comparison between the bioturbation index and the oxygen isotope mulation rates (Gardner and Dartnell; Stein, both this volume) be-
stratigraphy (after Kennett, this volume). Note close correlation of laminated cause the total depositional area of the basin was simultaneously
(oxygen depleted) intervals with warm stages and interstadials (OIS I, 3, reduced (Fig. 10).
parts of 5) and massive (oxygenated) intervals with cold stages and stadials Although earthquakes may trigger resuspension and redeposition
(OIS 2 and 4) since-75 ka. of slope sediments in this seismically active region, the distinct com-
position of gray layers (Fleischer, 1972; Stein, this volume) indicates
a unique provenance (i.e., floods from the Santa Clara/Ventura riv-
between the SBB configuration and the oceanic circulation system ers) rather than simple reworking of typical hemipelagic slope sedi-
must have been different prior to OIS 4. ment.
A more detailed and complex history of SBB bottom-water oxy-
Sand Turbidites
genation is recorded by the fine-scale variation in the lamination and
degree of bioturbation of the hemipelagic mud. Rather than simply Sand turbidites are largely restricted to two intervals, 107-115
switching from laminated to nonlaminated modes for extended peri- mbsf (-90-82 ka, OIS 5b) and 50-70 mbsf (-48-33, OIS 3) (Fig. 8).
ods of time, the degree of bioturbation varies on a variety of scales These sequences were chiefly deposited during intermediate eustatic
from centimeters to tens of meters, representing environmental levels, when sea level was 70-90 m below that at present, (Bard et al.,
change over decades to millennium (Fig. 7). In general, the lamina- 1990) when the shelf area was sharply reduced in size. The thickest
tion data or "bioturbation index" is presented as the 99 cm (-800 yr) sand bed at Site 893 was deposited during the minor sea-level drop
running average of the centimeter-resolution raw data to emphasize from OIS 5c to 5b when sea level was -20 m below that at present.
the major features of the record. For example, the interval from 30 to The stratigraphic distribution of sand turbidites is plotted as the run-
60 ka (roughly corresponding to OIS 3) displays an interval of inter- ning average of their percent thickness (i.e., centimeter of sand/meter
mittent lamination designated as the upper part of lithostratigraphic of core).
Subunit IC. On closer examination, the sediments record >l 2 distinct
intervals of lamination (oxygen-depletion), each lasting -1000 yr. At Grain Size
a finer scale, these intervals are shown to be composed of numerous Stratigraphic variation in the grain size of hemipelagic mud (Fa-
oscillations between oxygenated and dysaerobic conditions at peri- cies 1-4) is presented as weight % silt (Fig. 8). There is much sample-
ods < 100 yr (Fig. 7). to-sample variation that does not correlate with sedimentary facies or
lithostratigraphic unit. Grain size variation appears to have been most
Gray Layers
sensitive, however, to rates of sea-level change. Intervals of low per-
Gray layer event deposits are chiefly restricted to stratigraphic in- centages of silt at -165-150 mbsf and 20-10 mbsf are associated
tervals that are laminated (Fig. 8). This relationship suggests that de- with rapid sea-level rises at Terminations II and I, respectively. Inter-
tectable occurrences of gray layers are largely controlled by vals of high percentages of silt are less well correlated with sea-level
preservation due to the lack of bioturbation. This is not surprising, as lowering at OIS transitions 5e to d and 5a to 4.
SEDIMENTARY FAC1ES AND SEDIMENTOLOGY

200
3 2 1 4 3 2 1 4 3 2 1 4 3 2 1
Facies Facies Facies Facies
Massive < Laminated Massive < Laminated Massive ++ Laminated Massive < Laminated
Figure 7. Bioturbation index vs. age presented at a number of scales and degrees of smoothing. Major intervals of preserved lamination (i.e., oxygen-depletion
and nonbioturbation) are shown to consist of many, higher frequency oscillations between laminated and massive intervals. A. 99-cm running average. B. 19-
cm running average. C. 9-cm running average. D. 1-cm raw data.

Thornton (1984) studied modern core-top sediments in SBB and Figure 11 presents the stratigraphic variation in abundance of sil-
found a decreasing trend in percentage of silt toward the center of the iceous plankton, mica, sponge spicules, and total siliciclastic grains
basin from the major sources at the basin's margins. The strati graphic compared to oxygen isotopes as a proxy for eustatic sea level. The
trends of grain size at Hole 893 A do not reflect the changing proxim- abundance of siliceous plankton was apparently influenced by sea-
ity of the source areas as modulated by changing sea level. Instead, level fluctuations with peak abundances occurring at 150, 60, and 5
grain size must be controlled by the efficient transport of the silt- mbsf, corresponding to eustatic highstands during OIS 5e, 3 and 1.
sized fraction across the shelf into the deep basin center and the se- The influence of sea-level variation is also partially reflected in the
lective transport of certain compositional fractions (e.g., mica flakes) abundance of mica flakes, especially in the lower portion of the core
as outlined below. where three peaks correlate with sea-level highstands at OIS 5e, 5c,
and 5a. The abundances of sponge spicules and total siliciclastic
Compositional Trends grains are inversely related because the two components generally
make up 85%-98% of the samples. There is little direct correlation
Compositional data discussed here are derived from line counts of between variation in the abundances of these components and sea
smear slides of the carbonate-free, sand-sized fraction of the hemipe- level or other environmental indicators except that sponge spicules
lagic mud (Facies 1^1). Only 25 samples have been counted and are most abundant between 140 and 90 mbsf and 10-0 mbsf, intervals
these data must be considered very preliminary. Nevertheless, the that correspond to the general highstands of OIS 5 and 1. These in-
data suggest that several components of the >63-µm fraction may be creases during highstands may reflect either decreased transport of
useful proxies for change in sea level or environmental conditions. sand off the shelf or increased production of sponge spicules when
Total siliciclastics (i.e., quartz, feldspar, mica, and rock fragments) shelf area expanded.
are the most common component, making up 30%-98% of the sam- Figure 12 compares the presence and absence of macrofaunal de-
ples, averaging 69%; mica alone accounts for 8%-42% of the total bris—echinoid spines and sponge spicules—with the "bioturbation
samples (25% mean). Sponge spicules compose 1 %-51 % of the sam- index" and the oxygen isotope record. The presence of echinoid
ple (21% mean), whereas siliceous plankton (diatoms, radiolarians, spines in the hemipelagic mud is directly correlated with the absence
silicoflagellates) make up <l%-39% of the total carbonate-free sam- of laminations, indicating that the spines were derived from echino~
ple (8.5% mean). Presence or absence of echinoid spines was deter- derms that actually occupied the oxygenated basin floor, rather than
mined on separate samples that still contained carbonate material (I. being transported downslope from shallower waters. Echinoderms
Hendy, unpubl. data). are only present in the modern SBB in waters with >~0.3 mL/L O2

301
R.J. BEHL

Bioturbation index Gray layers (%) Sand layers (%) Grainsize Oxygen isotopes
l i i r

50 -

100 -
Q.
Φ
Q

150 -

200 I I j L

Facies 4 G 0 50 100 0 50 100 35 40 45 50 55


Massive < - Laminated (cm/m) (cm/m) Silt (wt %) Benthic foraminifers
Oxygenated • Anoxic (‰ PDB)

Figure 8. Stratigraphic distribution of bioturbation index, and percentages of gray layers (Facies 5), sand layers (Facies 6), and silt in the hemipelagic mud
facies, compared with oxygen isotopes (after Kennett, this volume). The oxygen isotopes of benthic foraminifers are presented here as a proxy signal for sea
level.

(interpreted from Savrda et al., 1984). Sponge spicules show no sim- posited during the glacial maxima (OIS 2, 4, and portions of 6). Al-
ilar relation to laminations and ancient oxygen levels, but instead are though a maximum glacio-eustatic lowering of sea level -120 m
weakly correlated with changes in sea level as described above, sug- would drop the present OMZ sufficiently to prevent oxygen-depleted
gesting that they are chiefly allochthonous sediments transported to waters from entering the SBB, oxygen levels in SBB have varied sig-
the central basin from the upper slope and shelf. nificantly without major sea-level change. The oxygen content of
SBB appears to most strongly reflect global to regional changes in
deep- and intermediate-water circulation (Fig. 13). Perhaps the clear-
DISCUSSION est example of supra-regional control of oxygenation of the SBB is
Paleoceanographic Controls of Sedimentation found at the Younger Dryas cooling event (-13-11 ka). The laminat-
ed mud sequence deposited under anaerobic conditions since the last
The sediment of SBB is clearly a sensitive recorder of paleocean- deglaciation is interrupted from 17.5 to 20.5 mbsf by a massive inter-
ographic and paleoclimatic change (e.g., Pisias, 1978, 1979; Dunbar, val of bioturbated sediment deposited during the Younger Dryas
1983; Baumgartner et al., 1992; Kennett, this volume). As detailed (Figs. 5 and 6; see also Kennett and Ingram, this volume). This strati-
above, changing levels of bottom water oxygenation (from <O.l mL/ graphic pattern in the silled SBB is nearly identical to that observed
L O2 to >0.3 mL/L O2) are shown by the degree of bioturbation of in the open slope environment at DSDP Site 480 in the Gulf of Cali-
hemipelagic Facies 1-4. This "bioturbation index" varies sharply and fornia (Keigwin and Jones, 1990). There was no significant sea-level
cyclically over a range of time scales from subdecadal to that of the drop at this time (merely a pause in sea-level rise) to affect the posi-
Milankovitch band (Fig. 7). Episodes of basin oxygenation or oxy- tions of water masses entering the SBB. Furthermore, with the same
gen-depletion apparently reflect both regional and global influences. signal recorded in different locations >IOOO km apart, this oxygen-
For example, laminations are absent or infrequent in sediments de- ation event likely reflects the increased ventilation (and freshening)

302
SEDIMENTARY FACIES AND SEDIMENTOLOGY

120°40•W 120°20' 119°40' 119°20

Surface currents Distribution of flood Nearshore sand transport


suspensates

Figure 9. A. Modern sediment transport pathways in Santa Barbara Basin.


Modified from Kolpack (1971) and Thornton (1984). B. Lowstand (-121 m)
configuration of the Santa Barbara Basin with speculative sediment transport
paths.

of intermediate water along the eastern Pacific rather than a local creased sediment bypass of the SBB, however, had the inverse effect
bathymetrically influenced signal. Indication of a wider—global— on sediment accumulation rates which remained relatively constant
control of oxygenation in SBB comes from the approximately dozen through high- and lowstands in SBB (see Gardner and Dartnell, and
intervals of preserved lamination (i.e., oxygen-depletion) recorded Stein and Rack, both this volume). The Santa Clara and Ventura riv-
during OIS 3 (Fig. 6). These events closely correspond in spacing to ers were probably captured by the Hueneme Canyon with most
the subtle peaks in the stacked deep-sea oxygen isotope record pre- coarse-grained material shunted directly into the adjacent Santa
sented by Martinson et al. (1987) as well as to the oxygen isotope Monica Basin (cf. Malouta et al., 1981). With falling sea level, the
record at Site 893 (Kennett, this volume). They have an even more eastern passage into the basin shoaled and narrowed, reducing the
startling correlation to the warm interstadials (Dansgaard-Oeschger flux of fine-grained suspensates into the basin. Furthermore, the
events) recorded in the oxygen isotope record of the Greenland ice lengthened Santa Clara-Ventura River system was restricted from
cores (Dansgaard et al., 1993). Variations in the carbonate composi- significant westward migration into SBB by the presence of >300 m
tion of piston cores from the North Atlantic likewise show similar highlands along the present coast (Fig. 9B). On the other side of the
number and spacing of pulses in OIS 3 (Keigwin and Jones, 1994). basin, longshore transport of sand from the west was intercepted by
Direct correspondence between these records links exceedingly rapid submarine canyons west of Point Conception (which currently head
switches in climate and deep- and intermediate-water circulation at -100 m water depth), routing most coarse-grained sediment away
across the globe to the oxygenation of SBB. from SBB into the deep Arguello Canyon. The western sill likely
shoaled above the depth of any potential oxygen-depleted water
Basin Geometry mass, and surface circulation was diminished with the coalescence of
the northern Channel Island chain into one large sheltering landmass.
Variation in sea level has a dramatic effect on the geometry of the
marginal SBB as shown by a comparison between the configuration
of the modern basin and during the last glacial maximum (Figs. 1 and CONCLUSIONS
9). The basin had approximately half the present depositional area at
maximum lowstand (Fig. 10) which, with input fluxes held constant, The distribution and character of sedimentary facies at Hole 893A
would double linear sedimentation and mass accumulation rates. In- was influenced by both global and regional paleoceanographic vari-

303
R.J. BEHL

5. Major intervals of lamination are composed of numerous


20 smaller-scale cycles in the degree of bioturbation reflecting
millennial to decadal variation in oxygen content of the basin.
0 6. Most sand transport to the basin center occurred at intermedi-
ate sea levels (-70 to -90 m) and was chiefly limited to two
-20 pulses at 90 to 82 ka (oxygen isotope Stage 5b) and 48 to 33
(m)

ka (oxygen isotope Stage 3).


-40 7. Grain size of the hemipelagic facies, as shown by the percent-
Φ age of silt, reflects eustatic variation in sea level, becoming
-60
lev

finer during rapid transgressions at Terminations I and II, and


coarser at transitions from oxygen isotope Stages 5e to d and
cσ -80 from 5a to 4.
Φ
CD 8. The composition of the sand-sized fraction of the hemipelagic
-100 - mud is chiefly controlled by relative sea level. Siliceous plank-
ton are more abundant (less diluted) at eustatic sea-level high-
-120 U stands. Mica flakes and sponge spicules were preferentially
transported into the basin during periods of higher sea level.
-140 9. The presence of echinoid spines in the sand-sized fraction of
2500 3000 3500 4000 4500 5000 hemipelagic mud correlates with bioturbated intervals in Santa
Barbara Basin, indicating that the massive sediment correlates
Area of basin (km2) with >0.3 mL/L of dissolved oxygen in the central basin.
Figure 10. Variation of the total depositional area of Santa Barbara Basin
with change in sea level constructed from present bathymetric contours.
ACKNOWLEDGMENTS

ation. Stratigraphic trends in grain size and sediment composition This research was supported by a postdoctoral fellowship from
were strongly influenced by effects of changing sea level on the dep- the Marine Science Institute, University of California at Santa Bar-
ositional geometry of the basin, which controlled the width of the bara. JOI/USS AC provided funds for travel to College Station, where
shelves, locations of sediment sources, sediment transport pathways, the assistance of the Ocean Drilling Program Gulf Coast Repository
and selective transport of sedimentary components. Although the Ho- staff was greatly appreciated. Echinoderm data were provided by In-
locene Santa Barbara Basin is well known for its varved sediments, grid Hendy. Thanks to James P. Kennett for invigorating discussions
the degree of bioturbation and preservation of laminations varied dur- and for providing useful criticism of this manuscript. This manuscript
ing the latest Quaternary, recording changes in the oxygenation level was greatly improved by the thoughtful reviews of Jack Baldauf,
of the basin. Chuck Savrda, and Scott Thornton—thank you.
An ultra-high-resolution record (<lO-yr sample spacing) of bot-
tom-water oxygenation ("bioturbation index") is constructed from REFERENCES
the centimeter-scale distribution of facies in Santa Barbara Basin and Alldredge, A.L., and Gottschalk, C.C., 1989. Direct observation of the mass
displays cyclic variation over decadal to Milankovitch time scales. flocculation of diatom blooms: characteristic settling velocities and form
Correlation of this stratigraphy with paleoceanographic and paleocli- of diatom aggregates. Deep-Sea Res., 36:159-171.
matic records elsewhere in the Pacific and Atlantic oceans suggests Bard, E., Hamelin, B., and Fairbanks, R.G., 1990. U-Th ages obtained by
that oxygenation of Santa Barbara Basin reflects the changing age mass spectrometry in corals from Barbados: sea level during the past
and ventilation of Pacific Intermediate Water rather than the local 130,000 years. Nature, 346:456^58.
systematics of a marginal basin. Baumgartner, T.R., Soutar, A., and Ferreira-Bartrina, V., 1992. Reconstruc-
tion of the history of the Pacific sardine and northern anchovy popula-
Specific findings of this study are:
tions over the past two millennia from sediments of the Santa Barbara
basin, California. Calif. Coop. Oceanic Fisheries Invest. Rep., 33:24-40.
1. The stratigraphy at Site 893 is defined by six sedimentary fa- Byers, C.W., 1977. Biofacies patterns in euxinic basins: a general model. In
cies: well-laminated hemipelagic mud (Facies 1). indistinctly Cook, H.E., and Enos, P. (Eds.), Deep-water Carbonate Environments.
laminated hemipelagic mud (Facies 2), trace laminated hemi- Spec. Publ.—Soc. Econ. Paleontol. Mineral., 25:5-17.
pelagic mud (Facies 3), massive hemipelagic mud (Facies 4), Dansgaard, W., Johnsen, S.J., Clausen, H.B., Dahl-Jensen, D., Gundestrup,
gray-layer event deposits (Facies 5), and sand turbidites (Fa- N.S., Hammer, C.U., Hvidberg, C.S., Steffensen, J.P., Sveinbjörnsdottir,
cies 6). A.E., Jouzel, J., and Bond, G., 1993. Evidence for general instability of
2. The Pleistocene distribution of sedimentary facies differs past climate from a 250-kyr ice-core record. Nature, 364:218-220.
Drake, D.E., Kolpack, R.L., and Fischer, P.J., 1972. Sediment transport on
markedly from that of the Holocene, in which laminated sedi-
the Santa Barbara-Oxnard shelf, Santa Barbara Channel, California. In
ments compose >70% of the stratigraphic thickness. At Hole Swift, D.J.P., Duane, D.B., and Pilkey, O.H. (Eds.), Shelf Sediment
893A, massive hemipelagic mud makes up 70% of the se- Transport: Process and Pattern: Stroudsburg, PA (Dowden, Hutchinson,
quence, whereas all laminated muds (Facies 1-3) combine to and Ross), 307-331.
compose only 21%. Droser, M.L., and Bottjer, D.J., 1986. A semiquantitative field classification
3. Increased sorting of gray layers (Facies 5) compared to the of ichnofabric. J. Sediment. Petrol., 56:558-559.
hemipelagic facies indicates that they settled chiefly as indi- Dunbar, R.B., 1983. Stable isotope record of upwelling and climate from
vidual particles from suspended flows, rather than as aggre- Santa Barbara Basin, California. In Thiede, J., and Suess, E. (Eds.),
gates in fecal pellets or marine snow. Coastal Upwelling, Its Sediment Record, Part B. Sedimentary Records of
Ancient Coastal Upwelling: New York (Plenum), 217-246.
4. For the past 75 k.y., the stratigraphic pattern of bioturbation
Dunbar, R.B., and Berger, W.H., 1981. Fecal pellet flux to modern bottom
(oxygenation) vs. lamination in SBB corresponds to glacial- sediment of Santa Barbara basin (California) based on sediment trapping.
interglacial oscillation in the global oxygen-isotope record, in Geol. Soc. Am. Bull., 92:212-218.
which warm episodes were oxygen-depleted and laminated Emery, K.O., 1960. The Sea off Southern California: A Modern Habitat of
and colder episodes were oxygenated and massive. Petroleum: New York (Wiley).
SEDIMENTARY FACIES AND SEDIMENTOLOGY

Emery, K.O., and Hülsemann, J., 1962. The relationships of sediments, life Pearson, T.H., and Rosenberg, R., 1978. Macrobenthic succession in relation
and water in a marine basin. Deep-Sea Res. Part A, 8:165-180. to organic enrichment and pollution of the marine environment. Ocean-
Fleischer, P., 1972. Mineralogy and sedimentation history, Santa Barbara ogr. Mar. Bioi, 16:229-311.
Basin, California. J. Sediment. Petrol., 42:49-58. Pisias, N.G., 1978. Paleoceanography of the Santa Barbara Basin during the
Gorsline, D.S., and Teng, L.S.-Y., 1989. The California Continental Border- last 8000 years. Quat. Res., 10:366-384.
land. In Winterer, E.L., Hussong, D.M., and Decker, R.W. (Eds.), The , 1979. Model for paleoceanographic reconstructions of the Cali-
Geology of North America (Vol. N): The Eastern Pacific Ocean and fornia Current during the last 8000 years. Quat. Res., 11:373-386.
Hawaii. Boulder (Geol. Soc. Am.), 471-487. Reimers, C.E., Lange, C.B., Tabak, M., and Bernhard, J.M., 1990. Seasonal
Heusser, L., 1978. Pollen in Santa Barbara Basin, California: a 12,000 year spillover and varve formation in the Santa Barbara Basin, California.
record. Geol. Soc. Am. Bull., 89:673-678. Limnol. Oceanogr., 35:1577-1585.
Keigwin, L.D., and Jones, G.A., 1990. Deglacial climatic oscillations in the Savrda, C.E., and Bottjer, D.J., 1987. The exaerobic zone, a new oxygen-
Gulf of California. Paleoceanography, 5:1009-1023. deficient marine biofacies. Nature, 327:54-56.
, 1994. Western North Atlantic evidence for millennial-scale Savrda, C.E., Bottjer, D.J., and Gorsline, D.S., 1984. Development of a com-
changes in ocean circulation and climate. J. Geophys. Res., 99:12397- prehensive oxygen-deficient marine biofacies model: evidence from
12410. Santa Monica, San Pedro and Santa Barbara basins, California Continen-
Kennedy, J.A., and Brassell, S.C., 1992. Molecular records of twentieth cen- tal Borderland. AAPG Bull., 68:1179-1192.
tury El-Nino events in laminated sediments from the Santa Barbara Schimmelmann, A., Lange, C.B., Berger, W.H., Simon, A., Burke, S.K., and
basin. Nature, 357:62-64. Dunbar, R.B., 1992. Extreme climatic conditions recorded in Santa Bar-
Kennett, J.P., Baldauf, J.G., et al., 1994. Proc. ODP, [nit. Repts., 146 (Pt. 2): bara Basin laminated sediments: the 1835-1840 Macoma event. Mar.
College Station, TX (Ocean Drilling Program). Geol., 106:279-299.
Kolpack, R.L., 1971. Oceanography of the Santa Barbara Channel. In Kol- Schimmelmann, A., and Tegner, M.J., 1991. Historical oceanographic events
pack, R.L. (Ed.), Biological and Oceanographical Survey of the Santa reflected in L1C/I2C ratio of total organic carbon in Santa Barbara basin
Barbara Channel Oil Spill 1969-1970 (Vol. 2): Los Angeles (Allan Han- sediment. Global Biogeochem. Cycles, 5:173-188.
cock Found., Univ. Southern Calif.), 90-180. Sholkovitz, E.R., and Gieskes, J.M., 1971. A physical-chemical study of the
, 1986. Sedimentology of the mainland nearshore region of Santa flushing of the Santa Barbara Basin. Limnol. Oceanogr., 16:479-489.
Barbara Channel, California. In Knight, R.J., and McLean, J.R. (Eds.), Soutar, A., 1967. The accumulation of fish debris in certain California
Shelf Sands and Sandstones. Mem.—Can. Soc. Pet. Geol., 11:57-72. coastal sediments. Calif. Coop. Oceanic Fisheries Invest. Rep., 11:136-
Kranck, K., 1984. Grain-size characteristics of turbidites. In Stow, D.A.V., 139.
and Piper, D.J.W. (Eds.), Fine-grained Sediments: Deep-water Processes Soutar, A., and Crill, P.A., 1977. Sedimentation and climatic patterns in the
and Fades: Geol. Soc. Spec. Publ. London, 15:83-92. Santa Barbara Basin during the 19th and 20th centuries. Geol. Soc. Am.
Krumbein, W.C., and Pettijohn, F.J., 1938. Manual of Sedimentary Petrog- Bull., 88:1161-1172.
raphy: New York (Appleton-Century-Crofts). Soutar, A., and Isaacs, J.D., 1969. History offish populations inferred from
Lajat, M., Saliot, A., and Schimmelmann, A., 1990. Free and bound lipids in fish scales in anaerobic sediments off California. Calif. Coop. Oceanic-
recent (1835-1987) sediments from Santa Barbara basin. Org. Geochem., Fisheries Invest. Rep., 13:63-70.
16:793-803. Thornton, S.E., 1981. Suspended sediment transport in surface waters of the
Lange, C.B., Berger, W.H., Burke, S.K., Casey, R.E., Schimmelmann, A., California Current off southern California: 1977-78 floods. Geo-Mar.
Soutar, A., and Weinheimer, A.L., 1987. El Nino in Santa Barbara basin: Lett., 1:23-28.
diatom, radiolarian, and foraminiferan responses to the "1983 El Nino" , 1984. Basin model for hemipelagic sedimentation in a tectoni-
event. Mar. Geol., 78:153-160. cally active continental margin: Santa Barbara Basin, California conti-
Lange, C.B., Burke, S.K., and Berger, W.H., 1990. Biological production off nental borderland. In Stow, D.A.V., and Piper, D.J.W. (Eds.), Fine-
southern California is linked to climatic change. Clim. Change, 16:319- grained Sediments: Deep-water Processes and Fades. Geol. Soc. Spec.
329. Publ. London, 15:377-394.
Lange, C.B., Schimmelmann, A., Yasuda, M.K., and Berger, W.H., in press. -, 1986. Origin of mass flow sedimentary structures in hemipelagic
Paleoclimatic significance of marine varves off southern California. In basin deposits: Santa Barbara basin, California Borderland. Geo-Mar.
Wigand, P., and Rose, M. (Eds.), Southern California Climate: Trends Lett., 6:15-19.
and Extremes of the Past 2000 Years: Los Angeles (Los Angeles Co. Trask, P.D., 1952. Source of beach sand at Santa Barbara, California as indi-
Nat. Hist. Mus.). cated by mineral grain studies. Univ. Calif. Tech. Rep. Eng., Ser. 14, no.
Malouta, D.N., Gorsline, D.S., and Thornton, S.E., 1981. Processes and rates 11.
of Recent (Holocene) basin filling in an active transform margin: Santa
Monica basin, California borderland. J. Sediment. Petrol., 51:1077-1095.
Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, T.C., Jr., and Date of initial receipt: 8 September 1994
Shackleton, N.J., 1987. Age dating and the orbital theory of the ice ages: Date of acceptance: 3 April 1995
development of a high-resolution 0 to 300,000-year chronostratigraphy. Ms 146SR-276
Quat. Res., 27:1-29.

305
R.J. BEHL

Oxygen isotopes Plankton (%) Mica (%) Sponge spic. (%) Siliciclastics (%
I I I I

100 -
α>
Q

150 -

200 J I I I I I I I I I I I I I I I
0 10 20 10 20 30 40 0 20 40 60 50 70 90
Benthic foraminifers >63 µm fraction >63 µm fraction >63 µm fraction >63 µm fraction
(‰ PDB) hemipelagic mud hemipelagic mud hemipelagic mud hemipelagic mud

Figure 11. Compositional trends in the sand-sized (>63 µm) fraction of hemipelagic mud from Santa Barbara Basin compared with the oxygen isotopic compo-
sition of benthic foraminifers at Site 893 (as a proxy for sea-level change).

306
SEDIMENTARY FACIES AND SEDIMENTOLOGY

Bioturbation index Echinoid spines Sponge spicules Oxygen isotopes


i i r

.Q

100 -
Q.
Φ
Q

150 -

200 I I
Facies 4 G Absent Present 0 20 40 60 4
Massive - Laminated >63 µm fraction >63 µm fraction Benthic foraminifers
Oxygenatθd -*• Anoxic (‰ PDB)
Figure 12. Stratigraphic trends in the presence or absence of echinoid spines and percentage of sponge spicules compared with the oxygenation/bioturbation
index and the oxygen isotopic composition of benthic foraminifers from Hole 893A. Echinoid data are smoothed (3-point running average).

307
R.J. BEHL

Dissolved oxygen
mL/L Santa Barbara Basin
0 12 3 4 5 6 7

IntermediateWater
(Oxygen Minimum Zone)

B
Dissolved oxygen
mL/L Santa Barbara Basin
0 12 3 4 5 6 7
Sea level
California Current Water Surface Productivity

( ( ( (
:::; Transitional Wa 1(1,
intermediate Water
genated)

Figure 13. End-member water-mass models explaining oxygenation levels of the Santa Barbara Basin. A. Oxygen-depleted intervals when laminations are pre-
served at a sea-level highstand. B. Oxygenated intervals when sediments are bioturbated and massive at a sea-level lowstand. Oscillation between basin-oxy-
genation states generally correlates to, but is not primarily driven by, eustatic variation. Oxygenation of intermediate water is probably the dominant control.
See text for discussion.

View publication stats

You might also like