Fulthorpe, C.S., Hoyanagi, K., Blum, P., and the Expedition 317 Scientists
Proceedings of the Integrated Ocean Drilling Program, Volume 317
Site U13521
Expedition 317 Scientists2
Chapter contents
Background and objectives . . . . . . . . . . . . . . . . 1
Operations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Lithostratigraphy. . . . . . . . . . . . . . . . . . . . . . . . 4
Biostratigraphy . . . . . . . . . . . . . . . . . . . . . . . . 22
Paleomagnetism . . . . . . . . . . . . . . . . . . . . . . . 29
Physical properties . . . . . . . . . . . . . . . . . . . . . 31
Geochemistry and microbiology. . . . . . . . . . . 35
Heat flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Downhole logging . . . . . . . . . . . . . . . . . . . . . 43
Stratigraphic correlation . . . . . . . . . . . . . . . . . 47
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
Tables. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Background and objectives
Hole U1352A
Position: 44°56.2440′S, 172°1.3615′E
Start hole: 1145 h, 30 November 2009
End hole: 1530 h, 30 November 2009
Time on hole (d): 0.16
Seafloor (drill pipe measurement from rig floor, m DRF): 354.8
(APC mudline)
Distance between rig floor and sea level (m): 11.0
Water depth (drill pipe measurement from sea level, m): 343.8
Total depth (drill pipe measurement from rig floor, m DRF):
397.0
Total penetration (m DSF): 42.2
Total length of cored section (m): 42.2
Total core recovered (m): 43.92
Core recovery (%): 104
Total number of cores: 5
Hole U1352B
Position: 44°56.2558′S, 172°1.3630′E
Start hole: 1530 h, 30 November 2009
End hole: 1615 h, 5 December 2009
Time on hole (d): 5.03
Seafloor (drill pipe measurement from rig floor, m DRF): 354.6
Distance between rig floor and sea level (m): 11.0
Water depth (drill pipe measurement from sea level, m): 343.6
Total depth (drill pipe measurement from rig floor, m DRF):
1185.5
Total penetration (m DSF): 830.9
Total length of cored section (m): 830.9
Total core recovered (m): 613.87
Core recovery (%): 74
Total number of cores: 94
1 Expedition 317 Scientists, 2011. Site U1352. In
Fulthorpe, C.S., Hoyanagi, K., Blum, P., and the
Expedition 317 Scientists, Proc. IODP, 317: Tokyo
(Integrated Ocean Drilling Program Management
International, Inc.).
doi:10.2204/iodp.proc.317.104.2011
2 Expedition 317 Scientists’ addresses.
Proc. IODP | Volume 317
Hole U1352C
Position: 44°56.2662′S, 172°1.3630′E
Start hole: 2015 h, 5 December 2009
End hole: 2200 h, 20 December 2009
Time on hole (d): 15.07
doi:10.2204/iodp.proc.317.104.2011
Expedition 317 Scientists
Seafloor (drill pipe measurement from rig floor, m
DRF): 354.5 (tagging seafloor)
Distance between rig floor and sea level (m): 11.0
Water depth (drill pipe measurement from sea
level, m): 343.5
Total depth (drill pipe measurement from rig floor,
m DRF): 2282.0
Total penetration (m DSF): 1927.5
Total length of cored section (m): 1296.4
Total core recovered (m): 655.02
Core recovery (%): 51
Total number of cores: 146
Hole U1352D
Position: 44°56.2326′S, 172°1.3611′E
Start hole: 0615 h, 21 December 2009
End hole: 2100 h, 21 December 2009
Time on hole (d): 0.61
Seafloor (drill pipe measurement from rig floor, m
DRF): 345.2 (APC mudline)
Distance between rig floor and sea level (m): 11.0
Water depth (drill pipe measurement from sea
level, m): 344.2
Total depth (drill pipe measurement from rig floor,
m DRF): 472.2
Total penetration (m DSF): 127.0
Total length of cored section (m): 127.0
Total core recovered (m): 130.84
Core recovery (%): 103
Total number of cores: 14
Integrated Ocean Drilling Program (IODP) Site
U1352 (proposed Site CB-04B) is located on the upper slope (344 m water depth) within the Canterbury
Bight and is the most basinward site of the Canterbury Basin drilling transect. This location was chosen as a primary site in response to an Environmental Protection and Safety Panel (EPSP) request
(December 2005) to avoid the high seismic amplitudes observed at 1.6–1.7 s two-way traveltime at Site
CB-04A. Site U1352 is located downdip from Site CB04A on dip seismic Profile EW00-01-60 (Figs. F1, F2).
Because of the move from Site CB-04A, there is no
crossing strike profile at Site U1352.
Site U1352 penetrates seismic sequence boundaries
U6–U19 where sediments are finer grained and pelagic microfossils are more abundant than at shelf
sites. This provides good age control for sequences
drilled on the shelf.
An additional target, requiring deep penetration, was
the Marshall Paraconformity, which has been dated
at its onshore type section using strontium isotopes
as representing a hiatus of ~3.4 m.y. (32.4–29 Ma)
(Fulthorpe et al., 1996). The paraconformity probably records intensified current erosion or nondeposition at all water depths that accompanied the develProc. IODP | Volume 317
Site U1352
opment of ocean circulation following the opening
of the seaway south of Tasmania (Carter, 1985; Fulthorpe et al., 1996; Carter et al., 2004). Seismic interpretation supports a current-related origin by indicating that the paraconformity forms the base of the
interval of sediment drift deposition. Indeed, immediately post–Marshall Paraconformity sedimentation involves sediment drift deposition in shallow
(Ward and Lewis, 1975), intermediate (Fulthorpe and
Carter, 1991; Lu et al., 2003), and deep water settings
(Shipboard Scientific Party, 1999a; Carter et al.,
2004). Drilling during Ocean Drilling Program (ODP)
Leg 181 indicates that the paraconformity developed
in deep (bathyal) water ~1–2 m.y. earlier than in
shallow water (McGonigal and Di Stefano, 2002).
Dating the paraconformity in the offshore Canterbury Basin at Site U1352 provides a further test of
this hypothesis.
Because of time constraints, drilling into one of the
large elongate sediment drifts of the Canterbury Basin, specifically drift D11 (Lu et al., 2003; Lu and
Fulthorpe, 2004), became a secondary objective.
Therefore, sites that were originally proposed for
drilling into drift D11 (proposed Sites CB-05B to CB05E) became contingency sites to be drilled only if
drilling at shelf sites was not possible. Nevertheless,
insights into sediment drift deposition and paleoceanography are expected from drilling at Site
U1352. The largest mounded elongate drifts lie
within the northeastern part of the shelf-slope sediment prism. Drift geometries become gradually less
pronounced along strike toward the southwest, and
mounded drifts are absent at Site U1352 (Lu and
Fulthorpe, 2004). However, the generation of
mounded drifts requires specific conditions that are
not well understood; a slope contour current alone is
insufficient, as indicated by the fact that such drifts
are not forming under the present current regime.
Current reworking of sediments is evident at Site
U1352, and currents may have left a paleoceanographic record of glacial–interglacial cycles, as at
ODP Site 1119 (Carter et al., 2004), without producing distinctive geometries.
The principal objectives at Site U1352 were
1. To sample slope sediments basinward of clinoform breaks of progradational seismic sequence
boundaries, particularly U6–U9, U11, and U13–
U19 (late Miocene to Pleistocene) to provide sequence boundary ages;
2. To penetrate the Marshall Paraconformity and
the top of the underlying Amuri Limestone (late
Eocene at total depth); and
3. To provide insights into the role of contour current deposition in a location where prominent
sediment drift geometries are absent.
2
Expedition 317 Scientists
Operations
Transit to Site U1352
After a 3 h, 8 nmi move from Site U1351 in dynamic
positioning (DP) mode, the R/V JOIDES Resolution
was positioned at Site U1352 at 0500 h (all times are
ship local time, Universal Time Coordinated [UTC] +
13 h) on 30 November 2009. The position reference
was a combination of Global Positioning System
(GPS) and an acoustic beacon on the seafloor,
weighted heavily toward the acoustic beacon (80%).
A positioning beacon (FSI BAP-547W, SN 1025, 14.0
kHz, 200 dB) was deployed at 0455 h on 30 November.
Site U1352 overview
Four holes were drilled at this site (Table T1). Hole
U1352A was cored with the advanced piston corer
(APC) system to 42.2 m drilling depth below seafloor
(DSF) with the objective of providing whole-round
samples for microbiology, chemistry, and geotechnical studies. Hole U1352B was cored with the APC/extended core barrel (XCB) coring system to XCB refusal at 830.9 m DSF. Hole U1352C was drilled and
cored with the rotary core barrel (RCB) system to the
target, record depth of 1927.5 m DSF. Hole U1352D
was cored with the APC system to 127 m DSF while
we waited for the weather to improve before transiting to the next site.
Logging was attempted with partial success in Holes
U1352B and U1352C. The third-generation advanced piston corer temperature tool (APCT-3) was
deployed five times with good results. Overall recovery for Site U1352 was 102% with the APC, 59% with
the XCB, and 51% with the RCB. The total cored interval for Site U1352 was 2296.5 m, and total recovery was 1444.1 m (63%).
Rig floor operations commenced at 0500 h on 30 November when the vessel shifted to DP control and
ended at 2100 h on 21 December when the vessel
was secured for transit.
Hole U1352A
The bottom-hole assembly (BHA) was made up with
an APC/XCB bit, and coring in Hole U1352A began
at 1145 h on 30 November. The mudline was established at 354.8 m drilling depth below rig floor (DRF)
(343.8 meters below sea level [mbsl]). APC coring
continued through Core 317-U1352A-5H (42.2 m
DSF), using nonmagnetic coring assemblies. Core
orientation was measured on all cores, and a temperature measurement was taken with Core 317-
Proc. IODP | Volume 317
Site U1352
U1352A-4H. Contamination testing was done on all
cores with perfluoromethylcyclohexane (PFT) and
microspheres. A total of 43.92 m of core was recovered from Hole U1352A (104%). The drill string was
pulled back to the seafloor, and the bit cleared the
seabed at 1530 h on 30 November, officially ending
operations in Hole U1352A.
Hole U1352B
The vessel was offset 20 m south of Hole U1352A,
and Hole U1352B was piston cored to Core 317U1352B-36H at 297.0 m DSF, with a total recovery of
298.85 m (101%). Core orientation was measured
with the Flexit tool on the first 17 cores before the
tool was pulled because of more severe piston coring
conditions. The XCB coring system was deployed for
Cores 317-U1352B-37X through 94X (297.0–830.9 m
DSF). Recovery with the XCB system was very good
to ~500 m DSF. Below that depth, a steadily increasing number of cores had poor recovery. A total of
533.9 m was cored, and 315.02 m was recovered
(59%). Toward the bottom of the hole, the XCB cutting shoe overheated and the formation caused excessive torque. The risk of damaging the core barrel
was too high to continue with the XCB system. A 50
bbl high-viscosity mud sweep was pumped to clean
the hole.
The drill string was tripped back to 80 m DSF, and
the triple combination (triple combo) logging tool
string was rigged up and run into the hole (RIH). The
first logging run failed to pass 487 m wireline log
depth below seafloor (WSF) because the hole evidently had collapsed during logging preparations.
Two passes were successfully recorded. The caliper
readings indicated that the borehole was too large to
allow an adequate clamp for the Versatile Seismic
Imager (VSI) tool to obtain reliable data, so the vertical seismic profile (VSP) plan was canceled. The Formation MicroScanner (FMS)-sonic tool string was
subsequently deployed, and it also encountered a
borehole obstruction, this time somewhat higher in
the hole. FMS-sonic logs were recorded from 442 m
WSF. After the logging tools were rigged down, a 12
bbl cement plug was pumped at 154.4 m DSF. The
drill string was tripped back to the surface and
cleared the rotary table at 1615 h on 5 December,
ending operations in Hole U1352B.
PFT and microspheres were deployed every ~50 m
throughout Hole U1352B for microbiological contamination testing. Temperature measurements with
the APCT-3 tool were taken with Cores 317-U1352B6H, 10H, 15H, and 20H. All temperature measurements had good decay curves. The cored interval in
3
Expedition 317 Scientists
Site U1352
Hole U1352B was 830.9 m, and total recovery was
613.87 m (74%).
predecessor programs. The cored interval in Hole
U1352C was 1296.4 m, with 655.02 m of core recovered (51%).
Hole U1352C
The hole was swept clean with a 50 bbl sweep of
high-viscosity mud, the RCB coring bit was released,
and the drill string was tripped out to ~900 m DSF.
The top drive was picked back up, and 400 bbl of
high-viscosity logging mud was displaced into Hole
U1352C. The top drive was set back, and an attempt
was made to continue the trip out of the hole. After
one stand, excessive drag required the top drive to be
reinstalled, and the trip out of the hole continued
with rotation until the end of the string reached 545
m DRF. The top drive was then set back and the drill
string was pulled back to 439 m DRF. The upper
guide horn was removed, and the vibration-isolated
television (VIT) camera was deployed to observe and
document the cone of cuttings at the seafloor during
logging. The drill string was set to a logging depth of
458 m DRF. The logging string was rigged up to run a
modified version of the triple combo tool string. The
first logging run indicated that the hole had collapsed, and the tool string was unable to pass 207.5
m WSF. The tool string was pulled back to the surface
and rigged down. The drill string was tripped back to
the surface, clearing the rotary table at 2200 h on 20
December, ending Hole U1352C.
The ship was offset 20 m south of Hole U1352B, and
operations in Hole U1352C began with the makeup
of a new BHA for the RCB system, which was fitted
with a mechanical bit release to facilitate logging after coring was completed. Drilling in Hole U1352C
began at 2015 h with a center bit installed. At ~300
m DSF, the center bit was recovered, inspected, and
reinstalled. The hole was then advanced to 574.7 m
DSF, at which point the center bit was pulled and an
RCB was dropped. The hole was cored with the RCB
system from 574.7 to 603.6 m DSF (Cores 317U1352C-2R through 4R), and 12.79 m of core (44%)
was recovered. The center bit was reinstalled, and
drilling continued to 660.0 m DSF. The center bit
was again pulled, and rotary coring resumed for
Cores 317-U1352C-6R through 41R (660.0–1009.9 m
DSF), with 20 bbl high-viscosity mud sweeps every
50 m of coring to clean the cuttings from the hole.
At 1900 h on 8 December, a 50 bbl sweep was
pumped and a wiper trip of the drill string was made
from 1007 to ~200 m DSF and back. Coring resumed
~8 h later for Cores 317-U1352C-42R through 118R
(1009.9–1661.5 m DSF), with sections of good and
poor recovery. The deplugger was run, and coring intervals were reduced to 5 m several times in attempts
to improve core recovery. On 15 December at 2030
h, a 50 bbl sweep was pumped and a wiper trip of the
drill string was made from 1662 to ~1000 m DSF and
back; coring resumed ~7 h later. The interval for regular 20 bbl hole cleaning mud sweeps was decreased
to 40 m.
On 17 December, it became evident that the scientific target for the hole, the Marshall Paraconformity,
was deeper than anticipated. Permission was requested and received to exceed the original EPSP
limit of 1913 m DSF by up to 250 m (to 2163 m DSF).
The Marshall Paraconformity was recovered in Core
317-U1352C-140R (1851–1861 m DSF) on 18 December. Coring continued to Core 317-U1352C-148R
(1927.5 m DSF) in order to provide sufficient depth
to log across the Marshall Paraconformity. The last
core from Hole U1352C was received on deck at
1740 h on 19 December.
Starting at 796 m DSF, PFT and microspheres were
deployed at ~50 m intervals throughout Hole
U1352C for microbiological contamination testing.
The penetration depth of 1928 m for Hole U1352C
established a new single-bit, single-expedition record
for the JOIDES Resolution and made Hole U1352C the
deepest sediment hole ever drilled by IODP or its
Proc. IODP | Volume 317
Hole U1352D
When operations in Hole U1352C concluded, the
winds and swells were too high to begin operations
at the next site, which was in extremely shallow
(<100 m) water. We anticipated a weather delay of
~24 h and decided to core Hole U1352D. An APC/
XCB assembly was made up, and piston coring in
Hole U1352D began at 0615 h on 21 December and
continued through Core 317-U1352D-14H (0–127.0
m DSF). Core recovery totaled 130.84 m (103%).
Nonmagnetic coring assemblies were used, and core
orientation was measured on all cores.
The drill string was pulled back to the rotary table,
the BHA was racked back, and the rig floor was secured for transit at 2100 h on 21 December, ending
Hole U1352D and Site U1352.
Lithostratigraphy
Four holes were drilled at Site U1352, reaching a total depth of 1927 m core depth below seafloor (CSFA; unless otherwise noted, all depths in this section
are reported in m CSF-A) and spanning the Holocene
to late Eocene. APC drilling was used in Holes
U1352A and U1352B to 297 m and in Hole U1352D
for the entire cored interval. XCB drilling was used
4
Expedition 317 Scientists
in Hole U1352B to 822 m (total depth), and RCB
drilling was used in Hole U1352C.
The succession was divided into three lithologic
units (Table T2; Figs. F3, F4). Recovery was variable,
ranging from an average maximum of 103% in Subunit IA (Holes U1352A, U1352B, and U1352D) to
38% in Subunit IIA (15% in Hole U1352B and 43%
in Hole U1352C). Subunit IB had an average recovery of 97%, whereas recovery for Subunit IC averaged 45% (15% in Hole U1352C and 60% in Hole
U1352B). The remaining units occur only in Hole
U1352C, where Subunit IIB averaged 53% recovery,
Subunit IIC averaged 67%, and Unit III averaged
54%.
Unit I spans the Holocene to middle Pliocene and
contains predominantly mud-rich sediment consisting of calcareous sandy mud; interbedded sand,
mud, and clay; massive sand; mottled sandy mud;
homogeneous mud; shelly mud; and marl.
Although the boundary between Units I and II was
placed at 710 m, the lithologic transition between
Units I and II is gradual, reflecting a progressive
change in water depth to deeper slope depositional
environments. Unit II spans the middle Pliocene
through early Miocene and contains hemipelagic to
pelagic sediment consisting of calcareous sandy
mud, sandy marls, chalk, sandy marlstone, and
sandy limestone, with minor amounts of calcareous
mudstone and sandstone. Unit II contains a gradual
progression from uncemented calcareous sandy mud
and marl to lithified marlstone and limestone. Notably, this unit generally lacks clay-sized material, except in discrete mudstone beds present in the lower
half of the unit. Glauconitic laminae and layers also
occur in the lowermost part of this unit. Packages of
recumbent and isoclinal folds, tilted beds, contorted
strata, and fluid escape features are present within
the middle portion of Unit II.
A hiatus representing ~12 m.y. occurs at the base of
Unit II at 1853 m, where there is an abrupt change to
lithologic Unit III, which is composed of hemipelagic to pelagic foraminifer-bearing nannofossil limestone of early Oligocene to late Eocene age. Except
for minor abundances of quartz and clay minerals,
Unit III lacks siliciclastic components. This unit is
correlative with the onshore Amuri Limestone.
Site U1352 represents a late Eocene to early Oligocene and nearly complete Neogene continental slope
sedimentary record dominated by pelagic to hemipelagic sedimentation with minor traction and gravityflow sediments. The gradual downhole transition in
lithofacies from more siliciclastic rich Pleistocene to
Holocene muddy facies into pelagic limestones and
glauconitic marls and marlstones appears to reflect
Proc. IODP | Volume 317
Site U1352
the transition seen on seismic profiles from an upper-slope location on a clinoformal margin with a
sharp shelf-slope break in the Pleistocene to Holocene toward a toe-of-slope to basin-floor position on
a more ramplike margin in the Miocene (Lu and
Fulthorpe, 2004).
Like those at other Expedition 317 sites, the lithologic units defined at Site U1352 exhibit gradual
changes downhole, making definition of unit
boundaries difficult. Lithologic units (Table T2) were
defined primarily by their lithologies and, importantly, by the repetitive assemblages of facies that occur in each unit. The boundaries between units and
subunits were chosen based on these criteria and
not, in general, on correlation to any other data set
(such as whole-round physical property measurements or downhole logs). However, in intervals of
relatively poor recovery, such data sets were used in
conjunction with lithologic logs and percentages to
determine the locations or depth ranges of unit
boundaries. Note that downhole logs exist only for
the uppermost 487 m of Hole U1352B and for the
uppermost 207 m of Hole U1352C because of hole
cave-in below these depths (see “Downhole logging”).
Description of lithologic units
Unit I
Intervals: Cores 317-U1352A-1H through 5H; 317U1352B-1H through 81X; and 317-U1352C-2R
to Section 11R-1, 92 cm
Depths: Hole U1352A: 0–43.11 m (total depth);
Hole U1352B: 0–710.78 m; and Hole U1352C:
574.7–709.32 m
Age: Holocene to mid-Pliocene
Unit I is divided into subunits according to the occurrence of two sandy lithologies. The first subunit
boundary (IA/IB) is placed at 98 m, below which
sharp-based, thin gray sand beds are absent. The second subunit boundary (IB/IC) is placed at 447 m to
mark the first downhole occurrence of thick graygreen calcareous muddy sand beds, which become
sandy marls with >30% carbonate content toward
the base of the unit.
Subunit IA
Intervals: Cores 317-U1352A-1H through 5H and
317-U1352B-1H to Section 11H-4, 23 cm
Depths: Hole U1352A: 0–43.11 m (total depth) and
Hole U1352B: 0–98.41 m
Age: Holocene to Pleistocene
Subunit IA is dominated by homogeneous mottled
and interbedded mud facies that generally contain a
few percent of very fine sand, often with thin clay-
5
Expedition 317 Scientists
rich intervals (Figs. F5, F6). Sand beds occur either as
thin, very fine to fine, dark gray, sharp-based beds
on a scale of a few centimeters or as thin to thick,
sharp-based, shelly, greenish very fine to fine muddy
sand or sandy mud (Fig. F7). The mud is gray and
sometimes mottled or color banded and has either
abundant shells (including Tawera, Chlamys patagonica delicatula, and other bivalve, gastropod, barnacle
plate, bryozoa, and worm tube fragments) or is devoid of shells (especially in intervals of interbedded
mud, sand, and clay). The alternation between shelly
and nonshelly mud beds occurs on a scale of 10–40
m. Deformation occurs in the uppermost 20 m of
Subunit IA, including normal faulting (interval 317U1352A-2H-2, 0–50 cm) and folding (intervals 2H-4,
99–130 cm, and 317-U1352B-2H-3, 10–57 cm) (Fig.
F8). The ichnofabric index ranges from 1 to 5 but is
typically moderate to complete (3–5) in the mud intervals. The thin gray sand beds are generally not
bioturbated, although sand-filled burrows sometimes
extend a few centimeters below their sharp basal
contacts (e.g., Section 317-U1352B-5H-3, 102 cm).
Subtle changes occur within this subunit. At the top
of Hole U1352A and in Cores 317-U1352B-1H
through 7H there is a marked alternation between
green calcareous muddy sand or sandy mud facies
and thicker beds composed of gray clay-rich mud
beds alternating with sharp-based, thin dark gray
sand beds with sand-filled burrows below (Fig. F9).
The boundaries of these beds are very distinct, although they are often deformed by drilling (Fig. F8).
Below Core 317-U1352B-7H the boundaries of the
clay, mud, and thin sand beds are no longer distinct,
but individual beds can still be identified. The lack of
distinct bedding contacts is presumably the result of
increased bioturbation. Below Section 317-U1352B11H-3, green sand beds are less common, and the
thin gray sand beds that occur frequently within the
interbedded mud and clay above are absent. Section
317-U1352B-11H-3 also marks a change in magnetic
susceptibility logs, with an apparent baseline shift
from higher to lower values across the Subunit IA/IB
boundary.
Mineralogy from smear slides in Subunit IA shows
that the dominant detrital components are quartz,
feldspar, rock fragments, mica (dominated by muscovite; see “Site U1352 smear slides” in “Core descriptions”), ferromagnesian minerals (e.g., hornblende),
and dense minerals (e.g., epidote and zircon). The
authigenic fraction is dominated by pyrite in the upper part of the succession and by microcrystalline
carbonate in the lower part of the succession and in
cemented zones. The percentage of carbonate components is generally characterized by a wide range of
values in Subunit IA, with especially high values in
Proc. IODP | Volume 317
Site U1352
smear slides taken from localized yellow patches
within the muddy units (partly cemented burrow
fills). The carbonate content of smear slides appears
to decrease downhole from the mudline and then increase toward the base of Subunit IA (Fig. F10). Rock
fragments observed in this subunit represent a range
of lithologies, including grains of siltstone, sandstone, and foliated metamorphic rocks (phyllite and
schist), as well as indeterminate rock fragments. Biogenic components are composed dominantly of foraminifers, nannofossils, and undifferentiated or
highly altered, abraded, or fragmented bioclasts.
A thin section of a burrowed lithified concretion in
interval 317-U1352B-2H-2, 32–34 cm, contains subequal amounts of quartz, feldspar, sedimentary to
metasedimentary rock fragments, and mica grains
locally suspended in a micritic to clay-rich matrix.
The shallow depth of this concretion (10 m) suggests
a very early precipitation of authigenic carbonate,
perhaps in association with methanogenesis.
X-ray diffraction (XRD) analyses of samples selected
from the dominant lithologies show that in Subunit
IA peak intensities of quartz and total clay are relatively high, whereas calcite is largely zero (Fig. F11;
see also XRD in “Supplementary material”). In this
subunit, the micas (muscovite/biotite) and chlorite
have relatively low, constant peak intensities with
depth relative to their intensities in the rest of the
hole. Plagioclase is variable and is not correlated
with quartz, as it is at Site U1351, and hornblende is
relatively low. Siderite is present in trace amounts,
but no appreciable pyrite or dolomite was observed.
Subunit IB
Interval: Sections 317-U1352B-11H-4, 23 cm, to
53X-1, 18 cm
Depth: 98.41–446.88 m
Age: Pleistocene
Subunit IB is dominated by gray to greenish gray and
dark greenish gray mud containing rare to locally
abundant shells; these shells become less abundant
and occur as broken fragments with depth. Decimeter-thick, greenish gray, calcareous very fine to fine
sand, sandy mud, and muddy sand are intercalated
with the mud, the latter typically forming fining-upward packages. Shells are also commonly associated
with the basal boundaries of green calcareous muddy
sand beds. These sharp basal surfaces of greenish
gray calcareous sandy mud or muddy sand beds are
often clustered, occurring as frequently as three
times in one core (e.g., Cores 317-U1352B-16H, 23H,
and 28H), separated by thicker uninterrupted gray
mud beds. Burrows often extend ~50 cm below the
base of the graded sand beds. Shells include bivalves
6
Expedition 317 Scientists
(including Chlamys patagonica delicatula), gastropods, echinoid spines, serpulid tubes, coral (Flabellum), and foraminifers. The alternating lithologies
change below a sharp-based green sandy marl in Section 317-U1352B-37X-2, becoming dominated by
color-banded mud with only very thin green sandy
layers. The green sandy layers at the base of each fining-upward succession become increasingly calcareous throughout Subunit IB. Below Core 317-U1352B35H, very dark greenish gray, poorly sorted very fine
to medium sandy marls first appear. These become
more frequent with depth, as indicated by smear
slide data. Calcareous nodules occur throughout this
subunit and are relatively common below 350 m.
The ichnofabric index ranges from no bioturbation
to heavy bioturbation (1–4), and identifiable traces
include Thalassinoides.
Smear slide analysis of Subunit IB reveals components similar to those found in Subunit IA, although
with higher carbonate percentages (Fig. F10). The
highest carbonate values occur close to the base of
Subunit IB, between 380 and 440 m. Smear slide data
through Subunit IB document an increase in siliceous biogenic components (diatoms, siliceous
sponge spicules, and other minor siliceous debris) in
tandem with increasing carbonate content (Fig. F10)
to a depth approaching 300 m. Below 300 m, siliceous components are less common, although they
are present throughout the remainder of Subunit IB.
Below ~205 m, micrite is a common component in
smear slides. Below 350 m, concretions or nodules
are relatively common; notably, this depth also coincides with the start of an increasing trend in calcium
concentration in interstitial water analysis (see
“Geochemistry and microbiology”).
Thin sections of lithified intervals in Subunit IB have
variable lithologies (see “Site U1352 thin sections”
in “Core descriptions”). The most numerous stratigraphic examples are concretions similar in texture
to those found near the top of Subunit IA (intervals
317-U1352B-16H-4, 130–134 cm; 17H-4, 69–72 cm;
and 21H-1, 0–2 cm; Fig. F12A). These concretions
contain bioturbated, micritized marlstone (muddy
limestone) with sparse, angular silt- to sand-sized terrigenous components, organic debris, and bioclasts
(including echinoderm fragments); degraded and
micritized foraminifers; and molds of dissolved mollusks. Localized concentrations of pyrite were found
within burrows and moldic pores. Lower in Subunit
IA, the lithified intervals are more variable and lithification appears to be related to a combination of
compaction and cementation rather than micrite
formation within the matrix. Some samples are foraminiferal limestones (e.g., Sample 317-U1352B-35H-
Proc. IODP | Volume 317
Site U1352
2, 40–41 cm [294.7 m]) having textures that vary between grain (foraminifer) and matrix supported (e.g.,
Sample 42X-5, 5–7 cm [347 m]; Fig. F12C). Textural
variations in the latter sample are related to bioturbation, and platy fossil fragments show no preferred
orientation. This sample also contains evidence of
significant compaction. Additionally, glauconite is
more prevalent, occurring as a pale green fill in foraminifer tests, as well as in darker green discrete lobate pellets, alteration products of biotite, and likely
some epigenetic glauconite in the matrix. The lowermost thin section in the subunit is of a matrix-rich,
burrowed marlstone containing a mixture of diagenetic and compactional features.
The interval from 280 to 300 m contains some unusual lithologies in the cores, including a >3 m thick
bed of shelly sand that was partly liquefied during
the coring process (Sections 317-U1352B-32H-5
through 33H-1) and a sandy marl bed (Sections 36H2 through 37X-2). In addition to high values of carbonate content, estimates from smear slide and XRD
data reveal marked fluctuations in mica, quartz and
feldspar, glauconite, and clay between 250 and 300 m
(Figs. F10, F11). As at Site U1351, the relative intensity of the calcite peak correlates with the total calcium carbonate concentration determined by coulometry, indicating that calcite is the dominant
carbonate mineral. However, minor amounts of Mg
calcite and aragonite based on peak heights were also
observed, always co-occurring with calcite (e.g., Sample 317-U1352B-32H-CC, 53–55 cm).
XRD data show a distinct mineralogical transition
within Subunit IB (Fig. F11). Quartz content determined from XRD analysis is significantly lower (p =
0.05 for this and other t-tests and correlations) below
the Subunit IA/IB boundary, which is matched by a
decrease in sand content, as seen in smear slides. Additionally, quartz content decreases significantly
within Subunit IB below ~250 m; this trend is also
observed in smear slides. Plagioclase remains fairly
constant within this subunit, but, in contrast to Subunit IA, a significant correlation (N = 40) exists between quartz and plagioclase. Total clays remain relatively high with no apparent depth trends, as also
seen in smear slides, but mica and hornblende increase significantly below the Subunit IA/IB boundary, with the increase in mica supported by smear
slide observations. Calcite peak intensity also significantly increases below the Subunit IA/IB boundary,
supported by both coulometry and smear slide observations. Pyrite and siderite peak intensities are
highest in Subunit IB, especially below 250 m. Dolomite was infrequently observed, with peak highs
close to detection limits.
7
Expedition 317 Scientists
Subunit IC
Intervals: Section 317-U1352B-53X-1, 18 cm,
through Core 81X and Core 317-U1352C-2R to
Section 11R-1, 92 cm
Depths: Hole U1352B: 446.88–710.78 m and Hole
U1352C: 574.7–709.32 m
Age: Pleistocene to mid-Pliocene
The top of this subunit was placed at the top of a 6.5
m thick bed of sandy marl, marking the top of an interval of ~40 m where this lithology is dominant. Below this, core recovery in Hole U1352B dropped dramatically, with the recovered intervals containing
homogeneous mud (largely lacking in shells), calcareous sandy mud and sandy marl, and sandy marlstone. Many of these sandy mud and sandy marl
beds fine upward. Recovery from Hole U1352C
(starting at 574.7 m) was also poor; however, the
cores from this hole contain largely lithified intervals, generally of very fine sandy marlstone (with
scattered small shells, including foraminifers). Bioturbation is particularly evident in the sandy marlstones (ichnofabric index of 1–4, where sediments
without recognizable depositional structures were recorded as 1). Poor recovery in Hole U1352B may be
caused by the intermittent presence of lithified layers composed of thick sandy marlstone between layers of unlithified sandy marl and homogeneous
mud.
The quartz, feldspar, and mica mineralogy in Subunit IC, as estimated from smear slides, is similar to
that of Subunits IA and IB, with higher and more
variable average carbonate concentrations and therefore lower percentages of other minerals (Fig. F10).
The presence of shells in the muddy sediments and
sedimentary rocks below ~575 m is extremely rare.
All mineral percentage estimates are highly variable
throughout Subunit IC, which distinguishes this
subunit somewhat from Subunits IA and IB and Unit
II. This high variability is also reflected in XRD data
(Fig. F11). Glauconite is more common toward the
top of Subunit IC and decreases in concentration
downhole. The proportion of siliceous bioclasts increases toward the base of the subunit before dropping markedly in the unit below.
Seven thin sections were prepared from Subunit IC.
One thin section (Sample 317-U1352B-72X-CC, 27–
29 cm) contains micritic limestone/marlstone and is
compositionally and texturally very similar to concretions observed at the top of the hole in Subunit
IA, which is likely attributable to downhole contamination. Three other thin sections show slightly different microfacies from those described in Subunit
IB: Samples 317-U1352B-54X-4, 4–6 cm (~459 m);
Proc. IODP | Volume 317
Site U1352
65X-1, 14–17 cm (~563 m); and 317-U1352C-4R-5,
52–54 cm (~600 m). In general, these microfacies are
finer grained and better sorted and have a smaller,
better sorted population of bioclasts. The sand fraction is mainly quartz, feldspar, mica, chlorite, and
dense minerals. These samples have variable
amounts of calcareous matrix and no distinct lamination, and they are somewhat homogenized by bioturbation, with rare large (centimeter scale) discrete
burrows.
The low core recovery in Subunit IC hampered a detailed analysis of depth trends in XRD mineralogy
(Fig. F11). In general, this subunit can be characterized as having no clear trends with depth in the minerals discussed above, although there is a notable increase in the variance of all these minerals. Quartz,
total clays, and micas co-vary and appear to be inversely related to calcite content. Interestingly,
quartz and plagioclase are not well correlated, as was
also the case in Subunit IB. Calcite has the highest
peak intensities of Unit I within this interval, which
is also supported by both coulometry and smear
slide observations. Pyrite and siderite peak intensities are highly variable, and dolomite was not observed.
Unit I/II boundary
As at other Expedition 317 sites, the Unit I/II boundary is transitional. All of Subunit IC represents a
transition between Unit I and Unit II lithologies,
marked by increasing percentages of carbonate and
decreasing amounts of clay (Fig. F11), although the
noncalcareous homogeneous mud common in Subunits IA and IB occurs as deep as Core 317-U1352B74X. The Unit I/II boundary was placed at the bottom of Core 317-U1352B-81X (and at the base of
Section 317-U1352C-11R-1) at the base of calcareous
muddy sand where (1) an abrupt change of baseline
occurs in magnetic susceptibility and gamma ray
whole-round measurements (see “Physical properties”), (2) the chemistry of interstitial water analyses
shows a local peak (see “Geochemistry and microbiology”), and (3) the analysis of XRD mineralogy
data shows a major shift in composition from claydominated to carbonate-dominated lithologies.
When total clays are normalized to calcite to account for carbonate-dilution effects, clay content decreases notably downhole beginning at 709 m (Fig.
F13). Similar increases in magnetic susceptibility and
natural gamma ray data were observed at the same
depth (see “Physical properties”). Below this depth,
sediments are carbonate dominated and clay content
only occasionally increases.
8
Expedition 317 Scientists
Unit II
Intervals: Cores 317-U1352B-81X through 94X
and Sections 317-U1352C-11R-1, 92 cm, to
140R-2, 47 cm
Depths: Hole U1352B: 710.78–822.13 m (total
depth) and Hole U1352C: 709.32–1852.63 m
Age: middle Pliocene to early Miocene
Unit II is dominated by homogeneous calcareous
sandy mud or sandy marl (uncemented intervals)
and sandy marlstone (cemented intervals) throughout Holes U1352B and U1352C. Subtle differences in
Hole U1352C allow for unit subdivision based on the
presence of dark muddy intervals (noncalcareous),
current-generated structures, and laminated sandstone beds in Subunit IIB, as opposed to the more
thorough degree of bioturbation present in Subunit
IIA and the presence of glauconite-rich beds in Subunit IIC (Fig. F14).
Subunit IIA
Intervals: Cores 317-U1352B-81X through 94X
and Sections 317-U1352C-11R-1, 92 cm, to 61R1, 30 cm
Depths: Hole U1352B: 710.78–822.13 m (total
depth) and Hole U1352C: 709.32–1189.30 m
Age: middle Pliocene to early Pliocene
Subunit IIA is dominated by dark greenish gray to
greenish gray homogeneous sandy marlstone along
with the less lithified sandy marl recovered in Hole
U1352B. Subordinate lithologies such as calcareous
very fine sandy mud, muddy sand, mud, and mudstone were recovered from both holes. The partially
to fully lithified sandy marl and sandy marlstone were
heavily bioturbated, and Chondrites, Helminthopsis,
Terebellina, Scolicia, Planolites, Zoophycos, and Thalassinoides ichnogenera were identified in the lithified
layers. The ichnofabric index for Hole U1352B cores
is generally 1 (partly because drilling overprint obscured the ichnofabric), but in Hole U1352C it
ranges between 1 and 5 (typically 3–4). Distinct bedding planes are not visible; however, in some intervals the overprinting of shallow and deeper trace fossils (e.g., Thalassinoides over Chondrites) suggests
erosion or other changes at the sediment/water interface (Fig. F15). Shells are extremely rare or are absent in the lower part of Subunit IIA and include
fragments of bivalves, brachiopods, and foraminifers. Small ovoid crystalline calcareous fragments or
“blebs” first occur in Core 317-U1352C-38R. Calcareous concretions occur sporadically throughout the
subunit.
Although variable, carbonate content is more uniform in Subunit IIA than it is in overlying units, and
estimated percentages from smear slide observations
range generally between 10% and 40% (Fig. F10).
Proc. IODP | Volume 317
Site U1352
Rare outliers, estimated from smear slide observations at >70% (chalk or limestone) carbonate, occur
in discrete, paler (light greenish gray) intervals (note
that coulometry analyses are not generally available
from these minor lithologies). The percentage of carbonate components appears to increase downhole in
the uppermost 200 m of Unit II, and values below
900 m seem to fall largely between 30% and 60%
carbonate based on smear slide estimations (15%–
50% based on coulometry data; Fig. F10). Mica percentages from smear slide and XRD data appear in
general to be higher than values in Unit I, whereas
clay concentrations are lower. Both glauconite and
siliceous bioclastic material become extremely rare
with depth. The percentage of ferromagnesian minerals is generally very low throughout Subunit IIA.
Subtly different facies, such as paler greenish gray
layers, begin to appear at ~1026 m (Core 317U1352C-44R) and become more common with
depth. These paler intervals, also bioturbated, are
generally more cemented and occasionally contain
>70% carbonate (estimated from smear slide observation), placing them in the chalk or limestone classification (Fig. F15). The dominant lithotype continues to be dark greenish gray very fine to fine sandy
marlstone, but the pale layers give the rock a colorbanded appearance in some places (e.g., Core 317U1352C-44R). In addition to these paler layers,
sandy laminae, sand lenses, and wavy or ripple laminae are occasionally present in the cores, increasing
in frequency downhole.
Thin sections from Subunit IIA contain distinctly
microburrowed marlstone facies. Although the overall composition is marlstone to sandy marlstone,
lighter trace-fossil fills and laminae observed in cores
were seen in thin section to be arkosic, micaceous,
and bioclastic siltstone to fine sandstone that is variably cemented by carbonate or that contains calcareous matrix. Darker burrow fills are mainly calcareous
marlstone, a combination of carbonate (nannofossils
and/or micrite) and clay with traces of organic matter. Minor terrigenous components include phyllite
lithic fragments, glauconite, and dense mineral
grains. A few samples contain higher percentages of
siliceous debris, including diatoms and sponge spicules, the latter locally concentrated in burrows (e.g.,
Sample 317-U1352C-11R-2, 21–23 cm). Some samples contain a higher proportion of clay and organicrich matrix (e.g., Sample 317-U1352C-40R-3, 2–3
cm). Diagenesis is limited to carbonate cementation
and pyrite formation as both pore-filling cement and
grain replacement, as well as replacement of opaline
sponge spicules by birefringent chert/quartz.
Based on XRD analysis, the mineralogy of Subunit
IIA is similar to that of Subunit IC, with relatively
9
Expedition 317 Scientists
moderate to high amounts of quartz, micas, and plagioclase feldspars relative to the remainder of the
succession (Fig. F11). These mineral groups have no
overall trend with depth but do have pronounced
variability with depth in this subunit. Total clays are
highest near the top of the subunit and decrease
from 660 to ~800 m within Hole U1352B. Hornblende is relatively low and variable. Calcite is lowest
near the top of the subunit and increases slightly
from 660 to 800 m, below which it has no trend with
depth but reveals the same pattern of variability
shown by silicate minerals, with the relative amount
of silicates inversely proportional to the amount of
calcite (significant at p = 0.05). The micas (muscovite/biotite) and chlorite are highly positively correlated (R = 0.92), and quartz and the micas are positively correlated, with a stronger correlation between
quartz and chlorite (R = 0.4). Plagioclase and quartz
abundances do not correlate in Subunit IIA, as was
also the case in Subunit IC. XRD mineralogy and
smear slide observations agree well, with both showing fluctuating trends in micas and sand, a relative
decrease in clays, and an increase in carbonate content near the top of the subunit.
Subunit IIB
Interval: Sections 317-U1352C-61R-1, 30 cm, to
123R-1, 142 cm
Depth: 1189.30–1693.92 m
Age: early Pliocene to late early Miocene
Darker gray and brown lithologies appear in the succession below 1170 m (Fig. F15). Sandy marlstone is
the dominant lithology, along with subordinate dark
gray, centimeter-thick, bioturbated very fine sandy
mudstone beds and centimeter-thick, horizontally
laminated fine sandstone. Based on smear slide observations, the latter lithologies are calcareous to
noncalcareous but are uniformly less calcareous than
the sandy marlstone. The Subunit IIA/IIB boundary
was placed at the base of an irregular surface that
separates lighter (above) from darker (below) very
fine sandy marlstone. Below this boundary, distinct
dark brown mudstone beds are more common and
are associated with more frequent sandstone beds,
wavy laminations, ripples, and sand lenses. Ferromagnesian minerals disappear from smear slides below Core 317-U1352C-62R (1204 m) and are below
detection in XRD analyses from approximately Core
317-U1352C-59R (1170 m) and below.
Subhorizontal wavy laminae marked by millimeterthick, brownish very fine sandy mudstone are increasingly more frequent downcore, becoming
prominent from Core 317-U1352C-70R downhole.
Below Core 317-U1352C-110R, these laminae are frequently intercalated with lighter gray sandy marlProc. IODP | Volume 317
Site U1352
stone, which sometimes gives the rock a dark–light
pin-striped appearance (Fig. F15F–F15G). Millimeterdiameter white calcareous blebs were observed
throughout the marlstone. Light blue-gray sandfilled burrows are common within the marlstone,
particularly in the lower portion of the subunit. Bioturbation is abundant throughout Subunit IIB, ranging between 1 and 5 (mostly 3–4) on the ichnofabric
index, and includes traces of Zoophycos, Planolites,
Chondrites, Thalassinoides, Paleophycus, Helminthopsis,
Terebellina, Scolicia, and Teichichnus.
The Pliocene/Miocene boundary occurs within Core
317-U1352C-73R, probably either at the top or the
base of a distinctive paler calcareous unit, the sharp
top of which occurs at Section 73R-2, 33 cm
(1276.83 m), and the sharp and bioturbated base of
which occurs at 73R-4, 20 cm (1279.7 m) (see “Biostratigraphy”).
In Sections 317-U1352C-88R-1 through 88R-4
(1371–1376 m), 94R-CC through 95R-1 (1437–
1439.5 m), and 105R-2 through 106R-5 (1528–1541
m), variable stratification attitudes, folding, shells,
and a greater proportion of carbonaceous material
mark several distinct stratigraphic intervals containing mass-transport complexes (slumps) that often include internal planar surfaces that divide areas with
different deformation characteristics or bedding angles (Fig. F16).
The common occurrence of dark brown muddy layers stops in Core 317-U1352C-90R (1392 m), although these layers continue to occur sporadically
downhole, and a notable change in mineralogy occurs at ~1400 m, as indicated by XRD analyses (Fig.
F11). These changes may be associated with a hiatus
observed in Core 317-U1352C-91R that is possibly
associated with a highly carbonate rich layer at
1409.12 m (Section 91R-7, 73 cm), where the faunal
age changes (between Sections 90R-CC and 91R-CC)
from latest Miocene to early late Miocene (upper
Kapitean to lower Tongaporutuan, ~5.30–11.01 Ma;
see “Biostratigraphy”).
Carbonaceous material is rare near the top of Unit II,
appears more frequently below 1400 m in association with the slumped material mentioned above,
and appears again in greater concentration in dark
brownish gray layers in Cores 317-U1352C-110R and
111R (1575–1590 m). Glauconite is a common constituent below Core 317-U1352C-105R, where a
graded bed has a glauconitic basal layer.
A banded appearance becomes evident in Core 317U1352C-113R (1603 m) and below. Distinctly wavy,
darker gray layers (laminae to thin beds) alternate
with greenish gray very fine sandy marlstone layers
(Fig. F15). The darker layers are commonly finer and
have the appearance of incipient stylolites (although
10
Expedition 317 Scientists
little dissolution appears to have occurred across
these surfaces). This banding increases in intensity
with depth, and glauconite begins to appear at low
concentrations in the marlstones in Core 317U1352C-117R.
The major minerals of Subunit IIB, as indicated by
XRD analyses, are similar to those of Subunit IIA,
with the exception of hornblende, which was not
observed apart from trace amounts in two samples,
and quartz, which decreases significantly in this
unit, especially below 1400 m (Fig. F11). Micas (and
chlorite) and plagioclase gradually decrease from the
top of the subunit to 1400 m and then increase notably to a maximum for this subunit at ~1550 m, below which they decline throughout the remainder of
the subunit. Total clays are also higher within this
subunit, especially in the uppermost part. Below
1580 m, clays were not detected by XRD. As in Subunit IIA, micas and chlorite are positively correlated
and total clays and calcite are inversely correlated;
however, there is no correlation between total clays
and micas, micas and quartz, or quartz and plagioclase.
Thin section observations show that Subunit IIB
contrasts with the overlying subunit in having more
matrix in its upper part and mainly a sandy marlstone lithology and some marlstone lithologies. No
other difference in the types or ratios of components
between Subunits IIA and IIB were observed. Siliceous sponge spicules, organic debris (plant matter),
and phosphatic debris (fish remains and fecal pellets?) are also present. In the upper part of Subunit
IIB, the matrix is a mixture of nannofossils and clay
that changes to micrite and clay below Section 317U1352C-78R-1 (1314 m). Most samples are bioturbated, but original laminations composed of alternating matrix-rich and well-sorted matrix-poor laminae are locally preserved. The latter are slightly
coarser and are carbonate cemented locally by poikilotopic calcite crystals (Sample 317-U1352C-90R-1,
0–3 cm). Sand and silt grains are generally angular.
Differential compaction is evident between more
competent cemented zones (usually better sorted
burrow fills) and less competent muddy zones. Longer mica flakes are bent and fractured across more
competent grains. Evidence for pressure solution is
more pronounced within and below Sample 317U1352C-95R-6, 39–42 cm (1447 m), where glauconite and phosphatic and organic debris are more
common. In particular, compaction (fracturing and
interpenetration) and truncation (press-solved
boundaries) of foraminifers were observed. Dark
seams visible in the core were shown to be created by
aligned and compacted platy organic matter, which
Proc. IODP | Volume 317
Site U1352
organic geochemical analyses indicate to be of terrestrial origin (see “Geochemistry and microbiology”). The irregular and stylolitic appearance of the
seams is a result of differential compaction around
more competent grains, as well as some apparent
pressure solution. Note that the more calcareous
samples below this interval (e.g., Cores 317-U1352C114R, 116R, and 118R) show evidence for pressure
solution only where minor organic matter is present.
Opaline siliceous sponge spicules are variably preserved and locally replaced by silica (chert) and microcrystalline calcite. In some samples (e.g., Cores
317-U1352C-70R and 89R), all three phases, original
opal, and secondary chert or carbonate are present.
Carbonate cement is present throughout as intraparticle filling of foraminifer chambers and as interparticle filling in better sorted grainstone/siltstone/
sandstone laminae and burrow fills. Zeolites were observed as pore-filling cement in foraminifer chambers in thin sections from Cores 317-U1352C-112R
and 113R. Zeolite crystals are locally encased by later
carbonate cement, indicating that they were an earlier diagenetic process.
Subunit IIC
Interval: Sections 317-U1352C-123R-1, 142 cm, to
140R-2, 47 cm
Depth: 1693.9–1852.6 m
Age: early Miocene
Subunit IIC is characterized by thin, intercalated
glauconitic sandstone interspersed with sandy marlstone and sandy limestone and, less frequently, with
mudstone lithotypes (Fig. F17). Sandstone first occurs as a minor lithology in Core 317-U1352C-115R,
and highly glauconitic sandstone first occurs in Core
123R (top of the subunit). Thereafter, these glauconitic layers become common, and the background
sandy marlstone also becomes glauconitic.
The marlstone consists of greenish gray, well-sorted
very fine to fine sandy marlstone. A glauconitic, bioturbated section at 317-U1352C-125R-5, 134 cm
(1714.34 m), was estimated by smear slide analysis to
contain >70% carbonate, making this interval glauconitic sandy limestone. The sandy marlstone in
Core 317-U1352C-126R changes to greenish gray,
well-sorted very fine to fine sandy limestone (i.e.,
>70% carbonate) by Core 132R and remains limestone to the base of the subunit. Stylolites are more
common with depth and are marked by millimeterthick muddy layers. A subordinate lithology consists
of dark greenish gray very fine sandy calcareous
mudstone. Bioturbation ranges between an ichnofabric index of 1 and 5 and includes Zoophycos and
Rosselia forms.
11
Expedition 317 Scientists
Glauconitic sandstones become more common with
depth to the base of the subunit. They consist of
moderately to well-sorted, very fine to fine (mostly
very fine), planar- to ripple-laminated calcareous
glauconitic sandstone. Ripple sets are <2 cm thick,
and the units themselves are typically <10 cm thick,
although one thicker unit (44 cm) occurs in Core
317-U1352C-135R. In the same core, the glauconitic
sandstones include both bedding-parallel (i.e., in situ
beds) and glauconitic sandstones that truncate bedding and burrows (i.e., sedimentary dikes and sills;
Fig. F17). The latter units generally appear to be subhorizontal but also occur at a high angle to stratification. They also bifurcate around centimeter-thick
sandy limestone units. In both types of glauconitic
sandstone, planar lamination is common. In the
glauconitic sandstones that cut primary stratification, two colors of sandstone occur: light blue-gray
and dark greenish gray. These layers also contain
small clasts of the surrounding limestone. Beginning
in Core 317-U1352C-139R, pyrite nodules as large as
1 cm in diameter occur.
Thin sections (see “Site U1352 thin sections” in
“Core descriptions”) from this subunit are mainly
glauconitic calcareous sandstone and glauconitic
sandy marlstone at the top of the subunit, transitioning to foraminifer micritic limestone at the base
of the subunit. Glauconite occurs as pellets, altered
bioclasts, and altered mica grains. Phosphatic fish remains and fecal pellets are locally present in minor
amounts. The matrix is a mix of micrite with an indeterminate amount of admixed clay minerals, and
the proportion of matrix changes within a sample
because of the presence of darker laminae and burrow fills. At the top of the subunit, the darker laminae are arguably depositional features enriched in
organic matter and clay minerals. Downsection,
these pass to striking dark bands that crosscut the
limestone fabric and are composed of glauconitic
calcareous sandstone to sandy marlstone. Some of
these crosscutting laminae fork, and others taper and
pass laterally into stylolites.
Authigenic phases are mainly carbonate (as intraparticle and interparticle cement) and pyrite. Diagenesis
includes patchy silicification, and low-birefringence
sulfate(?) cement is present in Samples 317-U1352C135R-5, 38–41 cm, and 137R-2, 58–60 cm (Fig. F12E,
F12F).
Evidence for compaction includes grain fracturing,
interpenetration, deformation of micrite-filled burrow margins, pressure solution, development of dark
seams, and stylolites. The darker intervals (laminae/
burrow fills) are associated with organic matter,
which appears to enhance pressure solution, creating
darker seams and leading to the formation of styloProc. IODP | Volume 317
Site U1352
lites. The amount of organic matter essentially declines downhole throughout the subunit, and below
Section 317-U1352C-137R-2 there is less compaction
or pressure solution and little evidence for stylolitization of the limestone lithologies.
The mineralogy of Subunit IIC from XRD analyses is
substantially different from that of other portions of
this unit. All silicates decrease significantly, and carbonate increases between Subunits IIB and IIC. As in
Subunit IIB, there is no detectable total clay content,
with the exception of Sample 317-U1352C-136R-3,
127–129 cm (1816.86 m), which also contains elevated quartz and mica (Fig. F11). Also similar to Subunit IIB, micas and chlorite are positively correlated,
but no correlation exists between total clays and calcite, total clays and micas, micas and quartz, or
quartz and plagioclase.
Unit II/III boundary
The boundary between Units II and III is an abrupt
lithologic change from glauconitic sandy limestones
and marlstones in Unit II to clean foraminiferal
limestone in Unit III. The boundary coincides with
the Marshall Paraconformity, which represents some
12 m.y. of missing time at this locality (see “Biostratigraphy”). The unconformity occurs as a zone
of rubble (apparently mostly blocks of the underlying limestone) in interval 317-U1352C-140R-2, 44–
48 cm.
Unit III
Interval: Section 317-U1352C-140R-2, 47 cm,
through Core 148R
Depth: 1852.6–1924.3 m (total depth)
Age: early Oligocene to late Eocene
Lithologic Unit III consists of white, cemented, finegrained slightly sandy or silty (a few percent) limestone. This unit was recovered largely in pieces as
long as 10 cm. Bioturbation is common throughout
and includes abundant Zoophycos burrows, together
with Chondrites and Planolites. Ichnofabric index values range between 1 and 5. Stylolites are very common and are better developed than in Subunit IIC.
Medium gray siliceous nodules as large as 5 cm in diameter occur sporadically throughout the unit.
Color differences include slightly darker intervals
and burrow fills as well as a purple coloration that is
associated with burrows or appears as circular or
curved features. Pyrite nodules in Unit III are associated with burrow fills or stylolites. Below Core 317U1352C-147R, well-laminated, light greenish gray
very fine sandy marlstones occur in beds as thick as 3 cm
and may contain pyrite specks or fragments. These
marlstone beds appear to occur more frequently
downhole.
12
Expedition 317 Scientists
Smear slide and thin section observations show that
the limestone matrix is recrystallized with preserved
fabric, suggesting that the original composition was
a foraminifer-bearing nannofossil ooze, with a few
percent (generally silt sized) grains of quartz and
feldspar. Of the seven thin sections prepared from
Unit III, six are composed of foraminifer micritic
limestone. The stylolites are marked by concentrations of mainly opaque minerals (including pyrite;
Figs. F12, F18), but they also contain clay, silt-sized
quartz, and organic matter and are oriented at several different angles in the cores, commonly crosscutting each other. They have variable amplitudes,
although they are better developed at the top of the
section, where they exhibit higher amplitudes,
thicker clay seams, and distinct crosscutting relationships. In contrast, the amplitude becomes very low
in the muddy limestone below the Oligocene–Eocene unconformity (Cores 317-U1352C-147R and
148R). One stylolite in Sample 317-U1352C-146R-3,
21–24 cm, appears to have had a later dilational
phase with intra-stylolite precipitation of microcrystalline carbonate and sulfate(?). The same sample
contains a partly silicified bioclast, and some chert
nodules and patches are present. In this section
(Sample 317-U1352C-145R-1, 36–37 cm) the siliceous rocks appear to be silicified equivalents of the
limestone, having been replaced by microquartz/
chalcedony (chert) and opal-CT (porcellanite).
The mineralogy of Unit III from XRD analyses is notably different from that of the overlying units at
this site and is dominated by carbonate minerals
(calcite) with only a minor amount of quartz below
1900 m.
Downhole trends in sediment composition
and mineralogy
The composition and mineralogy of Site U1352 is
generally comparable to the other Expedition 317
sites. However, because of more complete core recovery, deeper penetration, and the subsequent collection of older strata, several distinct downhole
changes in the concentration and relative proportions of components and mineralogy can be recognized between the three lithologic units. Primarily,
carbonate concentrations increase and clay content
decreases with depth in the hole.
Unit I is a clay-rich unit, composed primarily of
quartz, clay, feldspar, rock fragments, mica, hornblende, rare dense minerals, carbonate, and siliceous
bioclasts. Quartz and feldspar content is relatively invariant within Unit I, with somewhat higher plagioclase peak intensities within Subunit IC. A distinct
feature of Unit I is the co-varying changes between
Proc. IODP | Volume 317
Site U1352
total clay content, total micas, chlorite, and ferromagnesian minerals. Subunits IA and IB have relatively high total clay and ferromagnesian content,
with increasing mica and chlorite peak intensities
with depth. Subunit IC has distinctly lower ferromagnesian and total clay peak intensities and higher
mica and chlorite peak intensities. Glauconite is relatively low throughout Unit I but becomes relatively
more prevalent between 250 and 500 m, occurring as
fossil infills, often within cemented layers or concretions. Dolomite and siderite were occasionally observed, but peak intensities are close to the limit of
detection (see XRD in “Supplementary material”).
Overall, the depth trends in calcite peak intensity
follow smear slide and bulk CaCO3 concentrations.
Within Unit I, calcite concentrations are lowest in
Subunit IA, although there are local patches of high
authigenic microcrystalline carbonate concentrations, abundant large shells and shell fragments, and
concretions that occur as shallow as 10 m, indicating
early precipitation of authigenic carbonate. Micrite
becomes increasingly common below 205 m, and
concretions or nodules (marlstones to fossiliferous
muddy limestones) become common below 350 m,
corresponding to an increasing calcium concentration in interstitial water analysis. Subunit IC, below
450 m, has a higher carbonate concentration, which
is due to both high volumes of calcareous microfossils and also higher concentrations of authigenic carbonates. In contrast with the other three sites, Site
U1352 has high proportions of siliceous biogenic
components, with relatively elevated concentrations at ~300 and ~575 m, below which siliceous
fragments become abruptly less common.
The Unit I/II boundary at 710 m coincides with a
downhole shift from clay-dominated to carbonatedominated lithologies, which corresponds to
changes in magnetic susceptibility and gamma ray
measurements and interstitial water chemistry. Below this point, the sediments are significantly dominated by carbonate components (p = 0.05, N = 74),
and Unit II is subdivided largely according to the
presence or absence of other noncalcareous lithologies. Total clay from XRD is significantly lower in
Unit II relative to Unit I (p = 0.05, N = 74). Carbonate
concentrations increase gradually downhole through
Unit II, whereas hornblende and siliceous bioclasts
become rare and clay concentrations also decrease.
Below ~800 m, depth trends become less obvious,
but the mineralogy remains highly variable. An increase in carbonaceous material at ~1375, 1439, and
1535 m corresponds to lithologies interpreted as
mass-transport deposits. A decrease in quartz peak
intensities coinciding with an increase in mica and
plagioclase occurs at 1400 m, associated with a de-
13
Expedition 317 Scientists
crease in the number of dark muddy layers occurring
downhole. These changes may be associated with an
unconformity that is possibly associated with a carbonate-rich layer at 1409 m, representing ~5 m.y. of
Miocene time missing. Below this point, previously
rare carbonaceous material becomes more common
and glauconite content begins to increase below
1550 m, whereas mica and plagioclase decrease from
this point and clays were not detected below
1580 m. Pressure solution is evident in samples below 1440 m and is often associated with minor concentrations of organic matter. Some replacement of
siliceous bioclasts by secondary chert or carbonate
occurs, and carbonate cement is common, especially
in better sorted sandy layers and burrow fills.
A notable change in composition and mineralogy
occurs below 1700 m and the transition into Subunit
IIC, where the concentration of silicate minerals decreases significantly and carbonate content increases. Clay content was not detectable, except for a
few isolated intervals that also contain elevated
quartz and mica in rare muddy layers. The content
of glauconite increases downhole, although much of
this is associated with discrete layers that are interpreted as intrusions from below.
Unit III comprises recrystallized limestone, probably
originally a foraminifer-bearing nannofossil ooze,
with a few percent of generally silt-sized grains of
quartz and feldspar. Total clay, total micas, chlorite,
and ferromagnesian minerals were not observed in
Unit III. Below 1900 m, pyrite is associated with
thin-bedded very fine sandy marlstones, which increase in frequency downhole. Stylolites are well developed in the upper part of this unit and decrease in
frequency and amplitude downhole.
Correlation with wireline logs
Only a short interval of downhole wireline logging
data could be obtained from Hole U1352B because of
a blockage at ~500 m within Subunit IC, presumably
from hole cave-in. The caliper tool indicated that the
hole above this point was very wide (washed out)
(see “Downhole logging”), but a few narrow sections provided points at which caving material could
accumulate. In Hole U1352C, the triple combo tool
string was blocked at ~200 m. However, enough
downhole logging data were acquired to allow some
correlation to Unit I lithology (Fig. F19). In addition
to the limited downhole logging data, good core recovery over most of the drilled interval at Site U1352
makes it possible to compare lithologic trends to
other physical property data (e.g., magnetic susceptibility) acquired from whole-round measurements.
Note that the mineralogy of the sediments (feldspar
and mica in the sand fraction and quartz in the clay
Proc. IODP | Volume 317
Site U1352
fraction) reduced the ability of gamma ray logging
data to unambiguously resolve sand-to-mud transitions (see “Lithostratigraphy” in the “Site U1351”
chapter for further discussion). Also, the hole was
enlarged, so gamma ray logs should be used with
caution (see “Downhole logging”).
In Subunit IA, the lithology alternates between an
interbedded sand/clay/mud lithology and calcareous, often shelly, green muddy sands. The green
sandy beds correlate with low values of magnetic
susceptibility measured on the recovered cores and
also with low values on the downhole gamma ray
log. Mud beds or interbedded lithologies have relatively high magnetic susceptibility and gamma ray
values. A particularly high peak in magnetic susceptibility occurs between 58 and 61 m (Cores 317U1352B-6H through 7H).
In Subunit IB, the succession alternates between homogeneous mud intercalated with green sand. The
green sand beds are again calcareous and rich in
shells. As in Subunit IA, the sand beds in Subunit IB
have low magnetic susceptibility and gamma ray values; however, some low values in the downhole
gamma ray data (e.g., 255–265 m) do not apparently
correlate with lithologic beds and may relate to mineralogical differences within the mud beds or with
conditions in the hole.
The lithology of Subunit IC is characterized by more
frequent and more highly calcareous beds of green
sand and gray or green mud. The sand beds contain
shell fragments, biogenic clasts, and calcareous concretions. The low magnetic susceptibility and
gamma ray values correspond to calcareous sand
beds, as above; however, a relatively high value of
magnetic susceptibility between 630 and 668 m does
not correspond to any observed lithologic change in
the cores, although it does correspond to a peak in
clay and mica concentrations, observed in XRD
analyses, as well as a minimum in calcite concentrations (Fig. F11; also see “Physical properties”).
No downhole logging data are available for Units II
and III. Physical property measurements done on the
cores, including magnetic susceptibility and natural
gamma radiation (NGR), show trends associated
with lithologic changes within Units II and III (see
“Physical properties”). Magnetic susceptibility
shows another peak at the Unit I/II boundary at
~710 m and then decreases gradually until ~810 m,
where there is an abrupt upward baseline shift in the
readings and an increase in variability of the measurements, although there is no obvious change in
lithology. Below 900 m, variability in magnetic susceptibility and NGR decreases, concurrent with an
increase in recovery that is possibly related to increasing carbonate content or cementation. Peaks in
14
Expedition 317 Scientists
magnetic susceptibility and NGR occur at 1200 m
(near the Subunit IIA/IIB boundary) and 1440 m,
and another upward baseline shift occurs between
1450 and 1500 m (poor core recovery over this interval means the exact location of this shift is unknown). The peaks correspond to lost recovery intervals and to peaks in clay and mica concentrations in
XRD measurements and quartz and feldspar contents in smear slide data, possibly indicating that
these layers are less cemented, more terrigenous rich
layers.
Below 1500 m, NGR and magnetic susceptibility
trends decrease steeply, probably related to the rapidly increasing carbonate content of the sediments
in the lower part of Subunits IIB and IIC. The Marshall Paraconformity at the Unit II/III boundary
shows up as a dramatic downward shift and decrease
in variability of both magnetic susceptibility and
NGR as a result of the homogeneous and almost
purely calcareous nature of the sediments below the
boundary.
P-wave velocity measured on the cores shows a relatively stable average value downhole to 1500 m
(within Subunit IIB), where the measured velocity
increases to ~1700 m (at the base of Subunit IIB), and
then remains stable until ~1770 m before increasing
until ~1820 m (just above the Unit II/III boundary;
see “Physical properties”). A slight downward trend
may be present in the bottommost 50 m of core, possibly related to the increasing clay content in the Eocene limestone at the base of the hole. It appears
from these trends that velocity is directly related to
the carbonate cement content of the sediments,
which increases downhole to the Oligocene–Eocene
unconformity within Unit III.
Description of lithologic surfaces and
associated sediment facies
Sedimentary package description
A major objective of Expedition 317 was to understand how eustasy, climate, and tectonics influence
sedimentation in the Canterbury Basin, with a particular emphasis on assessing how these large-scale
forcing mechanisms interacted to create stratal packages and the surfaces that bound them, as resolved
in seismic reflection profiles (Lu and Fulthorpe,
2004). To this end, we developed a very generalized
classification of the different types of lithologic contacts and their associated deposits to aid in core-seismic integration and possible correlation with onshore strata of equivalent age. The classification
scheme used here reflects the general changes in lithology of the different lithologic units and applies
to all drilled sites. In Unit I at Site U1352, the upper-
Proc. IODP | Volume 317
Site U1352
most 50 m of strata (Subunit IA) contains both centimeter-scale and meter-scale interbedding that reflects lithologic changes, whereas the lower 50–
450 m (Subunit IB) generally contains sharp contacts
that separate different lithologies and that are associated with discrete, thicker (1–6 m) beds. Poor recovery within the lower part (450–710 m; Subunit IC)
hampered evaluation of the presence of surfaces and
their associated sediments within that interval. From
710 to 1852 m in Unit II, the sediment is characterized by calcareous cemented lithologies that contain
large-scale (6 m thick) and small-scale (5–30 cm
thick) sedimentary packages bounded by sharp basal
surfaces. Discrete contacts and their associated lithologies are recognized by changes in grain size,
texture, color, bioturbation, and carbonate content.
Based on these large-scale trends and the characteristics of the sediments, distinct contacts and facies associations were defined as sedimentary packages,
which may potentially vary acoustic impedance and
therefore be resolvable in seismic reflection data.
These classifications are greatly simplified for the expedition report, and postcruise analyses will permit
better definition of sedimentary packages and improve their correlation to seismic data. Broadly, contacts and facies associations were classified based on
grain size (muddy sand, sandy mud, sandy marl,
marl, or clay), bed thickness, the nature of the basal
and upper contacts, and the amalgamation of thinner (centimeter thick) beds with discrete contacts
that form meter-thick or thicker packages (Figs. F20,
F21).
Type A contacts and facies associations are characterized by a sharp contact that commonly separates
thick lithologies above and below. The lithology
above the contact can be as thick as 6 m (Fig. F20).
Specific examples include thick, dark greenish gray
calcareous fine to medium muddy sand or sandy
mud in sharp contact with underlying greenish gray
mud. The basal contacts of the overlying greenish
gray sandy lithology are sharp and lightly to heavily
bioturbated, whereas their upper contacts tend to be
gradational. When their lower contacts are heavily
bioturbated, discrete burrows can be tracked 100 cm
beneath the contact. Shells and shell fragments are
present in the greenish gray muddy sand and sandy
mud deposits within the uppermost 150 m of Holes
U1352A and U1352B. These become rare in Hole
U1352B from 150 to 250 m and are generally absent
from the deeper lithologies, except in a bed near 372
m where they are abundant. When shells and shell
fragments are present, they are as long as 3 cm and
are abundant within the greenish gray sands immediately above the basal contact. Shells and shell fragments can be present below the basal contact in the
15
Expedition 317 Scientists
greenish gray mud, but the position of these shells is
most likely the result of burrowing disturbance because they seem to be closely associated with greenish sand-filled burrows. Mass wasting deposits (as
thick as 15 m) found in Unit II from ~1370 to 1550 m
are a second example of Type A contacts and deposits. These deposits have a sharp upper and lower contact, soft-sediment deformation features, and medium to coarse sand layers, with generally lower bulk
density than the adjacent strata (e.g., Sections 317U1352C-94R-6 through 94R-7 and Cores 105R and
106R; Fig. F16).
Type B contacts and facies associations represent
amalgamation packages of thinly bedded but distinctly contrasting lithologies (e.g., sand, mud, and
clay interbedding and sandstone–marlstone and
marlstone–limestone alternations). These contrasting lithologies also form meter-thick packages (Fig.
F20). The individual beds are thin (1–5 cm), have
decimeter-scale spacing, and commonly have sharp
basal contacts. When sand is present, it is normally
graded. Because of the sharp contrasts in lithology
and the thickness of these packages, these types of
deposits may be resolvable in seismic reflection profiles.
Type C contacts and facies associations are always
sharp and separate contrasting lithologies, for example sand and mud or sandstone and marlstone. Because of this sharp contact in lithology, these surfaces have the potential to create reflections in
seismic profiles. However, the thickness of any beds
associated with these contacts could not be clearly
delineated on board ship (e.g., muddy sand layers
fining upward into sandy mud), and additional
shore-based analyses are required to establish the relative importance of each in terms of its potential to
generate a seismic reflection (Fig. F20).
Table T3 lists the occurrence of each of the three
types of contacts and their associated sediments at
Site U1352. Type A contacts and associated facies are
more common in Unit I than they are in Unit II.
Type B contacts and facies associations are more
common in the uppermost 50 m of Unit I and
throughout Unit II, where they consist of more heavily cemented mudstone and marlstone, marlstone
and sandstone, and marlstone and limestone beds
within homogeneous sandy marlstone. Within Unit
I, Type B contacts and facies associations are packages of interbedded sand and mud or mud and clay
(e.g., 26, 34, and 98 m) or several decimeter-thick
muddy sand beds that are closely spaced (e.g., 148,
159, 331, and 338 m). Type C contacts and facies associations are found throughout Holes U1352B and
Proc. IODP | Volume 317
Site U1352
U1352C and represent a range of lithologic contacts.
Within Unit I, these sedimentary packages tend to
comprise fining-upward muddy sand that sharply
overlies mud; within Unit II, they comprise beds
with distinctly different carbonate contents and/or
degrees of cementation. An additional example is
the Unit II/III boundary between sandy limestones
and limestone.
Description of significant surfaces
Because of time restrictions on board ship, surfaces
were only examined close to the predicted depths of
seismic sequence boundaries; therefore, the lithologic surfaces identified here are implicitly linked to
the predicted occurrences of sequence boundaries
identified on the seismic (Lu and Fulthorpe, 2004). A
similar approach was used on board ship during ODP
Legs 150 and 174A, the objectives of which were also
to study sea level changes. Postcruise study will attempt to clarify the exact relationship of all lithologic surfaces and facies associations to sea level
changes and seismic stratigraphy. The numbering
system used in the site chapters, tables, and summary diagrams comprises a hole-specific prefix and a
surface designation (e.g., U1352A-S1) that links each
surface to a seismic sequence boundary; therefore,
these lithologic surfaces and associated sediments
are thought to be correlative between sites across the
transect.
The following section contains a brief description of
each surface identified at Site U1352, as summarized
in Table T4. In contrast to the situation at Site
U1351, the identification of significant surfaces at
this site was complicated by the common presence
of more than one lithologic package per core near
the predicted depths of seismic sequence boundaries
(Lu and Fulthorpe, 2004). In these situations, the significant surface has generally been assigned to the
lithologic package that was thicker and best preserved, but the other surfaces identified nearby are
also mentioned.
Surfaces U1352B-S1 and U1352D-S1
A Type A contact and its associated facies, designated
U1352B-S1, is tentatively placed at Section 317U1352B-7H-6, 96 cm (64.20 m) (Table T4). A second
sharp, angular dipping contact is present above
(62.50 m). This surface is similar to that at 64.20 m,
but the former is less burrowed, contains fewer shells
and shell fragments, and has thinner beds that grade
upward into greenish mud. Note also that from 75 to
93 m there are several intervals of muddy sand beds
25, 86, and 88 cm thick in Cores 317-U1352B-9H,
16
Expedition 317 Scientists
10H, and 11H, respectively. A lithologic surface of
similar character, U1352D-S1, was observed at Section 317-U1352D-8H-3, 88 cm (64.38 m).
Surface U1352B-S2
A Type B contact and its associated facies, designated
U1352B-S2, is tentatively placed at Section 317U1352B-16H-5, 5 cm (147.22 m) (Table T4). The sediments associated with this contact contain a 5 cm
thick cemented layer at Section 317-U1352B-16H-4,
128 cm (146.96 m). A second Type B contact is present at 146 m. The overlying sedimentary package is
30 cm thinner than the sedimentary package beneath, and the contacts above and below are gradational into gray mud. Surface U1352B-S2 is therefore
tentatively positioned in the sharp contact of the
thickest sand bed at 147.22 m.
Surface U1352B-S3
Surface U1352B-S3 is tentatively positioned at a
sharp contact at Section 317-U1352B-23H-1, 130 cm
(200.00 m). A second sharp contact is present at Section 317-U1352B-23H-6, 80 cm (207.00 m). Note
that muddy sand beds 4–27 cm thick are present in
most sections of Cores 317-U1352B-24H and 25H.
Surface U1352B-S4
Surface U1352B-S4 is tentatively placed at Section
317-U1352B-28H-4, 7 cm (250.20 m) and is defined
by a sharp contact that separates gray mud below
from sandy mud above. This Type A contact and facies associations is thin (<1 m) relative to other packages. Two other intervals of sandy mud were also observed between 246 and 249 m. It should also be
noted that sand beds as thick as 88 cm are present in
Sections 317-U1352B-27H-4 through 27H-7, indicating that these sandy intervals extend several meters
below the selected contact.
Site U1352
ond sand bed (4.15 m thick) is present from 478.10
to 482.20 m. This muddy very fine sand is calcareous
and terminates in a sharp basal contact at 482.2 m.
Heavy bioturbation is present for 45 cm beneath the
contact. Surface U1352B-S6 is positioned at the base
of the thickest sand associated with the sharp contact at Section 317-U1352B-56X-5, 70 cm (482.20
m).
Surface U1352C-S9
Surface U1352C-S9 is located within a more calcareous interval within the marlstones defined as chalk.
This interval contains a sandy chalk bed from 990.25
to 990.56 m and several 2–5 cm thick chalk-rich intervals between 1026 and 1034 m. The sandy chalk
bed has a sharp, scoured contact at Section 317U1352C-40R-3, 5 cm (990.56 m), that separates it
from marlstones above and chalky marlstones beneath. This contact is designated as surface U1352CS9.
Surface U1352C-S10
A series of Type B contacts with silty sand beds with
planar bedding, ripple laminations, shell fragments,
and rip-up clasts are present between 1112.50 and
1113.00 m. The contact at Section 317-U1352C-53R1, 90 cm (1113.00 m), is designated as surface
U1352C-S10.
Surface U1352C-S11
An unusually thick (3 m), very calcareous deposit
with sharp basal and upper contacts is present between 1276.50 and 1279.80 m (Fig. F22). This deposit is present within Sections 317-U1352C-73R-2
and 73R-4, where the Pliocene/Miocene boundary
was identified. Surface U1352C-S11 was placed at the
base of the deposit at 1279.8 m.
Surface U1352C-S12
Surfaces U1352B-S5 and U1352B-S5.1
Two potential Type B surfaces and their associated
sediments are designated as surfaces U1352B-S5 (Section 317-U1352B-51X-1, 131 cm [428.81 m]) and
U1352B-S5.1 (Section 53X-5, 30 cm [453.00 m]). The
latter is associated with an unusually thick (6.5 m)
bed of very fine sandy calcareous mud.
Surface U1352B-S6
Lithologically, the depth interval from 475.5 to
484.2 m is very unusual. From the base up, it contains a 15 cm thick bed of very fine sandy mud that
fines upward. The basal contact at 484.2 m that separates the sand above from the mud beneath is heavily bioturbated for 15 cm below the contact. A secProc. IODP | Volume 317
A succession of Type A/B facies deposits found from
1400 to 1438 m documents an increase in laminations and other current structures, such as planar
beds and ripples, as well as an increase in grain size
and sand beds interpreted as stronger current activity (Fig. F15). Additionally, slump deposits are present between 1371 and 1438 m (Fig. F16). The base of
a slump deposit (Section 317-U1352C-94R-7, 60 cm
[1938.20 m]) was selected as surface U1352C-S12.
Surface U1352C-S13
One of the most notable surfaces, U1352C-S13, is
linked to the Marshall Paraconformity (Fig. F18).
This surface is present at Section 317-U1352C-140R2, 48 cm (1852.64 m), and is marked by a limestone
17
Expedition 317 Scientists
gravel bed that separates glauconitic silty limestone
above from limestone beneath. The gravel bed is
composed of ~5–10 cm long subrounded limestone
fragments having weathered surfaces.
Discussion and interpretation
Interpretation of Unit I
Subunit IA
Seismic reflection profiles indicate a series of downlapping reflectors in this interval, as described by
Browne and Naish (2003), who interpreted these as a
series of lowstand delta front deposits. The lithologies and sedimentary structures of Subunit IA are
consistent with this interpretation. We interpret the
sharp-based, dark gray sand to be a product of sediment gravity flows fed by such deltas (gravity flow in
this context implies downslope mass transport, not a
specific type of transport). The slumped and faulted
intervals observed in the uppermost 20 m of both
Holes U1352A and U1352B may be related to high
rates of sediment supply in delta or pro-delta settings.
Subunit IA is characterized by dark gray muds as well
as sharp-based, gray and greenish sand beds. The
greenish sand beds are calcareous, and in part siliceous, with abundant to common macrofossils often
concentrated into layers. This, together with the
sharp-based and fining-upward character of the beds,
suggests deposition from gravity flows. The background mud observed in the cores throughout this
subunit probably formed through hemipelagic deposition in slope settings or areas removed from major
point sources of sediment input from either westerly
or southerly source areas.
The gray sand beds in this subunit have compositions consistent with derivation from low-grade zeolite and prehnite-pumpellyite graywacke rocks (Torlesse Terrane) of the Canterbury region. This subunit
differs from other units, which appear to have a predominantly schist-derived provenance. For example,
XRD data indicate that chlorite is low throughout
Subunit IA but increases markedly at 100 m, which is
consistent with a change from a dominantly Torlesse
provenance in Subunit IA to an increased Otago
Schist provenance in Subunit IB (Fig. F11). Grains
within the subunit are both well rounded and angular—a bimodal transport history suggesting that the
well-rounded grains may be recycled or were
rounded in a high-energy beach setting and mixed
with less rounded, fluvially transported sediment before being transported offshore.
Subunit IA, characterized by an overall high siliciclastic content, corresponds to the past ~0.25 m.y.
Proc. IODP | Volume 317
Site U1352
(see “Biostratigraphy”), a time period of pronounced glaciation in the Southern Alps (Suggate,
1990). The relatively high concentration of quartz
and feldspars within clay-rich intervals might be a
glacial rock-flour signature (Heiden and Holmes,
1998; Peuraniemi et al., 1997), and the input of Torlesse-derived material implies that glaciers that originated from the Southern Alps were a dominant sediment source for this subunit.
Subunit IB
The interbedded greenish sands that occur in dark
gray muds are typically sharp based and fine upward
and are interpreted as sediment gravity flows. Similar
sands were interpreted as contourites in Subunit IIB
at Site 1119, but because we lack evidence of coarsening- and then fining-upward trends, we prefer to
suggest that these sand beds represent either downslope mass transport from shallower areas rich in carbonate or are winnowed condensed intervals. For example, one such sand interval found in Sections 317U1352B-32X-5 through 32X-CC contains abundant
shells that likely represent a sediment gravity flow
originating from a shallow-marine setting. XRD
analyses show that total clays are lowest and calcite
is highest in the calcareous, olive-green muddy
sands, which may indicate that these units formed
during periods of minimal hemipelagic sediment input or during periods of enhanced traction transport
and erosion that winnowed out the clay component.
Lastly, the noted decrease in quartz and mica content at ~250 m corresponds to ~0.8 Ma (see “Biostratigraphy”), which is during the mid-Pleistocene
climatic transition from high- to low-amplitude eustatic sea level change (Clark et al., 2006).
Similarly, we suggest that the homogeneous mud,
which has a relatively high total clay concentration,
represents mostly hemipelagic deposition in upper
slope settings based on micropaleontological evidence for an upper slope depth of deposition (see
“Biostratigraphy”).
Subunit IC
Coulometry and smear slide data indicate higher carbonate content in Subunit IC than in the overlying
Subunit IB interval. The fine-grained sediments are
interpreted to be hemipelagic sediments deposited
with a large amount of pelagic carbonate components. Small to moderate proportions of terrigenous
sand (10%–35%) may have been derived from downslope mixing or along-slope sand supply, but the better sorting and fine-grained nature of the sand fraction, including a high percentage of bioclasts,
suggest considerable transport. Occasional clay-rich
18
Expedition 317 Scientists
muddier units indicate episodic increases in terrigenous supply. Micropaleontological evidence (including the presence of reworked microfauna) supports
an interpretation of drift sedimentation in a slope
setting for this subunit. This interpretation is consistent with the seismic line in a strike perspective between Sites U1352 and 1119 that shows a series of
small drifts at this stratigraphic level (seismic Line
EW00-01-19).
Interpretation of Unit II
Unit II is distinguished from Unit I in that it is dominantly calcareous and is composed of sandy marlstone with minor components of limestone and
sandy mudstone.
Subunit IIA
Subunit IIA is interpreted to have been deposited in
a hemipelagic to pelagic setting largely removed
from clastic sediment sources, although the very fine
to fine sand, mica, and clay were probably derived
from upslope terrigenous sources or along-slope sediment supply, as suggested by the variable mineralogy
within this subunit. The sandy sediments are interpreted as thin traction deposits, formed during the
reworking of some of the sand into planar-laminated
and ripple-laminated beds. We interpret the marlstone as largely drift deposits, consistent with the
seismic profile between Sites U1352 and 1119 (seismic Line EW00-01-19). The more calcareous lithologies may represent condensed intervals formed during periods when sediment was starved on these
drifts. The increased frequency of chalky intervals in
the lower part of Subunit IIA may relate to a slower
rate of sedimentation at the base of a clinoformal
package of sediment.
Site U1352
defined and more pervasive with depth, and lower in
the subunit some of the white calcareous blebs are
dissolving about their margins and forming white
trails into the laminae. Thin section observations indicate that the brown wavy laminae are composed of
organic matter and associated clay-rich material deposited in laminae and reworked in burrows that
have undergone pressure solution, in effect a combination of the proposed mechanisms above.
The sharp-based nature of the darker sandstones and
mudstones suggests increased current activity and
rapid deposition of these beds. It is probable that the
mudstone represents increased deposition of continent-derived material (clay, organic material, etc.).
The horizontally and ripple-laminated sandstone
beds were deposited by traction currents, probably
also sourced from continental areas or, less likely,
from reworking of the sand component of the marlstones. Smear slides indicate a considerable amount
of organic matter (as much as 10%) in the lower portions of the subunit, implying that a landmass was
not too distant from the depocenter. Perhaps the
mudstone intervals represent periods of lower sea
level, when land-derived sediment was prevalent,
whereas the marlstone might represent hemipelagic
sedimentation during periods of reduced sediment
input from the continent, possibly during rising and
high sea level. In general, we interpret these sediments to represent largely hemipelagic to pelagic deposition removed from major sources of sediment
(except during lowstands), consistent with a sediment-drift setting.
The alternation of light-colored marlstone with
darker colored mudstone and sandstone layers indicates an alternation of dominant processes from
more quiescent periods of pelagic deposition to periods of increased sediment supply and current activity. The overall downhole decrease of silicate minerals in this subunit is a continuation of the transition
to a more carbonate rich (pelagic) environment with
less terrigenous sediment input.
Intervals of soft-sediment deformation recorded in
the lower portion of Subunit IIB may reflect instability caused by sediment loading, seismic triggering,
and/or topography within the drift deposits or adjacent submarine highs (the “Endeavour High” of
Field and Browne, 1989). These slumped intervals
have higher quartz, clay, mica, and feldspar contents, supporting the interpretation that these deposits were formed by mass movements originating
closer to shore. The association of a graded sandstone-mudstone turbidite at the top of one of these
mass flow units suggests that subsequent mass flows
were concentrated into bathymetric lows, which
formed by the evacuation of the mass-transport
complex that preceded it.
This subunit contains numerous types of bedding
structures that may have formed from a combination of transport and diagenetic processes. The common thin, brown, wavy lamination that causes the
pin-striped appearance of the lower part of the unit
could be (1) original burrows, (2) primary ripple lamination, (3) stylolitic laminae, or (4) a combination
of the above. These laminae appear to become better
A prominent variance in mineralogy occurs within
this subunit, as indicated by the interval of higher
mica and chlorite content between ~1470 and 1570
m. As opposed to other sections at this site where
micas and quartz/clay positively co-vary, the rise in
mica content does not correspond to a similar rise in
quartz or clay minerals but does increase along with
plagioclase. Smear slides and thin sections from
Subunit IIB
Proc. IODP | Volume 317
19
Expedition 317 Scientists
these same intervals reveal the presence of altered
basaltic fragments, plagioclase, and zeolites, suggesting that chlorite may be authigenic rather than detrital and may have formed from alteration of mafic
volcanic fragments. This depth interval corresponds
to ~10–13 Ma (see “Biostratigraphy”), when the
mafic volcanic fields at Banks and Otago peninsulas
formed (McDougall and Coombs, 1973; Coombs et
al., 1986; Sewell, 1988; Sewell et al., 1992).
Subunit IIC
The marlstones and limestones in this subunit are
predominantly fine grained, suggesting pelagic to
hemipelagic deposition in a deepwater setting with
only a small terrigenous sediment supply. Micropaleontology suggests deposition in a lower bathyal
setting (see “Biostratigraphy”). The horizontally
laminated nature of the intercalated glauconitic
sandstone beds implies some form of traction deposition. In these instances, the sandstone may represent reworking of sediment from areas enriched in
glauconite, such as sediment-starved highs. Although this is probably true for some of the observed
layers, many of the sandstone layers, especially those
in the lower part of the subunit, clearly crosscut primary stratigraphic features and are interpreted as diapiric in nature, both dikes and sills. Examples of
this exist in the core where vertical or near-vertical
glauconitic sandstone bodies fed into or were fed
from subhorizontal glauconitic sandstone. Other
core examples show thinner glauconitic sandstones
branching from thicker equivalents with a dendritictype morphology. Some diapiric sandstones show
fragments of the surrounding carbonate lithology
within them, implying that these fragments were
broken off the surrounding walls during emplacement. Accounts of diapiric sandstones at Oamaru
(Lewis, 1973) and southern North Island (Browne,
1987) indicate that similar diapiric features occur
nearby in similar lithotypes. We suggest that the
source of the glauconitic sandstone is the Oligocene
Kokoamu Greensand, which should occur stratigraphically below these limestones (but was not recovered in the core) and immediately above the
Amuri Limestone (Unit III), based on the nearby exploration well Clipper-1. Sediment loading and/or
mild tectonic warping may have caused the upward
injection of the glauconitic sand into Subunit IIC
(e.g., Jolly and Lonergan, 2002).
Interpretation of Unit III
We interpret Unit III to have accumulated as pelagic
foraminifer-bearing nannofossil ooze. The faint purple bands and rings noted in the cores may be related to disseminated fine iron sulfide within the
Proc. IODP | Volume 317
Site U1352
limestone, as described in similar facies recovered
from the Ontong Java Plateau during ODP Leg 130
(Lind et al., 1993). These color bands are interpreted
as a possible alteration of volcanic ash or other impurities in the limestones. Foraminifer and nannofossil studies reported here (see “Biostratigraphy”)
suggest a lower bathyal depth between 1000 and
1500 m water depth.
This unit is correlative to the regionally extensive
Amuri Limestone, which is widespread throughout
South Island and lower North Island, New Zealand.
Regional studies suggest that deposition may occur
in either shelf (van der Lingen et al., 1978) or
bathyal (Edwards et al., 1979) settings. Seismic reflection profiles suggest that Site U1352 was situated
slightly seaward of the shelf during the early Oligocene but that the margin was a broad ramp at that
time with no clear shelf/slope break. This is consistent with the interpretation made by Field and
Browne (1989) that the Amuri Limestone was deposited in an outer shelf to slope paleoenvironment.
Interpretation of lithologic surfaces and
associated sediment facies
Hole U1352B
The identified lithologic surfaces and their associated sedimentary packages (Types A, B, and C) are in
some instances near the predicted depths of seismic
sequence boundaries (Lu and Fulthorpe, 2004), allowing a tentative correlation of lithology to seismically defined surfaces (Figs. F22, F23; Table T4; see
also Table T4 in the “Site U1351” chapter). The sharp
and often heavily burrowed basal contacts of some
of these packages suggest that erosion and/or sediment bypass are associated with the formation of the
discontinuity. The lithologic surfaces from Site
U1352 differ from those of Site U1351 in that they
are composed of several Type A sedimentary packages per core (~9 m intervals) instead of just one
event, as was observed at Site U1351. Multiple sedimentary packages can potentially provide a stronger
impedance contrast and stronger seismic reflections
on the seismic lines. This should be further investigated during postcruise study.
Surface U1352B-S1 is tentatively correlated to seismic sequence boundary U19, which is predicted at
68 m. Of the two depositional events present, surface
U1352B-S1 is correlated to the thicker sand bed at
64.16 m and to U19. Surface U1352A-S1 has a similar
age range as U1352B-S1 and was identified in interval 317-U1352A-3H-1, 0–70 cm.
Surface U1352B-S2 is a Type B contact and associated
facies and is tentatively placed at 147.3 m and correlated with U18, which has a predicted depth of 142
20
Expedition 317 Scientists
m. Surface U1351B-S2 has a similar age range as
U1352B-S2 and was identified in Holes U1351A and
U1351B at 27.75 and 31.00 m, respectively (see Table
T4 in the “Site U1351” chapter).
Surface U1352B-S3 is tentatively positioned at a
sharp contact at 200 m. The Type A contact and associated facies designated U1352B-S3 correlates to
U17, predicted at a depth of 195 m. Surface U1352BS3 has a similar age range as U1351B-S3 (see Table T4
in the “Site U1351” chapter).
Surface U1352B-S4 is tentatively placed at 250.20 m
and correlated with U16, which has a predicted
depth of 249 m. Again, as in previous surfaces, the
lithology indicates three depositional events associated with this interval. Surface U1352B-S4 has the
same age range as U1351B-S4 (see Table T4 in the
“Site U1351” chapter).
Surface U1352B-S5 at 428.81 m is correlated to U15,
which has a predicted depth of 428 m. Surface
U1351B-S5 is also tentatively correlated to U15.
Surface U1352B-S5.1 at 453.20 m is tentatively correlated to U14, which has a predicted depth of 448 m.
A lithologic surface correlative to U14 was not found
at Site U1351, probably because of a recovery gap between 97 and 113 m over the predicted depth of U14
at Site U1351 (103 m).
Surface U1352B-S6 is tentatively positioned at the
base of the thickest of two sand beds associated with
a sharp contact at 482.20 m. This surface is correlated with U13, which has a predicted depth of 500
m. The sediments from 485.10 to 504.40 m are composed of gray homogeneous mud and had good recovery, except for intervals of cementation between
489 and 492 m. The series of sand beds associated
with this contact provides evidence of sediment
transport. Surface U1352B-S6 has a similar age range
as U1351B-S6, where sand beds and evidence of sediment reworking also make the placement of the
Pleistocene/Pliocene boundary in that core questionable (see Table T4 in the “Site U1351” chapter).
U12 and U10 were not found at Site U1352. A Type A
contact and facies association, U1351B-S7, is present
in Core 317-U1352B-81X at 710.65 m. This lithologic surface and its facies association is tentatively
correlated to a seismic surface that could potentially
be the continuation of U12. U12 could not be traced
from Site U1351 to U1352, so at this point this interpretation is based on the lithology only and not on a
seismic surface.
U11 has a predicted depth of 769 m, but a corresponding lithologic surface was not located near this
depth. U10 is absent. The Type A surface U1351B-S7
is present in Section 317-U1351B-22X-1 (171.4 m)
and tentatively correlated with U12.
Proc. IODP | Volume 317
Site U1352
Hole U1352C
Several cores in this hole have distinctive lithologies
that could be identified as Type A or Type B surfaces.
These surfaces, although rare, are important and
have been tentatively linked to U9 and U8 (U1352CS9 [990.60 m] and U1352C-S10 [1113.00 m], respectively). Sediments from 1113 m to the base of the
hole at 1924 m include notable events associated
with hiatuses and mass flows, also classified as Type
A and Type B contacts and facies associations. These
events have been tentatively linked to U7 and U6
and to the Marshall Paraconformity. These surfaces
and associated sediments are named U1352C-S9 to
U1352C-S13. Sediments that could be interpreted as
representing U5 and U4 were not recovered at Site
U1352.
Surface U1352C-S9 is tentatively linked to U9, which
occurs at a predicted depth of 970 m, based on increased calcareous content and a sandy chalk bed.
Because of its high calcareous content, the deposit
associated with this lithologic surface may represent
late high-stand episodes of condensation. This lithologic surface was not directly identified at Site U1351
because sediments near the predicted depth of U9
(312 m) were not recovered.
Surface U1352C-S10 and its associated facies is a
Type B lithologic contact positioned at 1113 m and
tentatively correlated to U8, which occurs at a predicted depth of 1136 m. The depth at which U8 is
predicted at Site U1351 (394 m) was not recovered
because of recovery gaps between 394 and 400 m.
U1352C-S10 cannot be correlated to the shelf Site
U1351.
Surface U1352C-S11 is a Type A contact and facies
association. As a result of its sediment composition,
sharp boundaries, and age, the base of this deposit is
tentatively correlated to U7, which occurs at a predicted depth of 1251 m. The predicted depth interval
for U7 at Site U1351 is 614 m; however, this surface
was not identified at Site U1351 because of poor recovery from 611 to 620 m.
Surface U1352C-S12 was chosen at the base of a
slump deposit at 1438 m. This lithologic surface
closely corresponds in depth with U6, which occurs
at a predicted depth of 1428 m. The corresponding
U6 interval at Site U1351, at the predicted depth of
895 m, was not recovered.
Based on age and lithologic relations, U1352C-S13 is
correlated to the Marshall Paraconformity (observed
at 1852.6 m), which correlates to the predicted depth
of seismic reflector “Green” at 1824 m. A gravel bed
marks this lithologic surface, which suggests that the
gravels were derived from the limestone beneath and
were reworked prior to deposition of the overlying
21
Expedition 317 Scientists
sediment. It is likely that these reworked limestone
blocks were derived from the weathered top surfaces
of the underlying chalky limestone and that a layer
of unconsolidated glauconitic sand between the rubbly surface of the underlying limestone and the coherent glauconitic limestone above exists in the subsurface but was not recovered by coring. This
interpretation is based on the exposures onshore of a
similar (though shorter duration) unconformity
(e.g., Fulthorpe et al., 1996). The unconsolidated
glauconitic sediments or rubbly top surface of the
limestone would provide a strong impedance contrast, leading to the strong seismic signal of the Marshall Paraconformity sequence bounding unconformity. U1352C-S13 correlates with a hiatus between
the upper Miocene (18–19 Ma) and upper Oligocene
(30–32 Ma), and its recovery was a very important
goal of the expedition. There are no major lithologic
features at 1647 m, the originally predicted depth for
the Marshall Paraconformity.
Biostratigraphy
Core catcher samples from Holes U1352A–U1352D
were examined for calcareous nannofossils, diatoms,
planktonic and benthic foraminifers, and bolboformids to develop a preliminary biostratigraphic
framework for the cored succession (Fig. F24; Table
T5). Calcareous nannofossils and planktonic foraminifers provided the primary means of age control
at Site U1352, but marine diatoms, bolboformids,
and benthic foraminifers also contributed to dating.
Benthic foraminifers were used to determine paleowater depths and interpret depositional environments. All depths in this section are reported in m
CSF-A.
Integrated microfossil records from Site U1352 reveal
a 1924 m thick sedimentary succession spanning the
Holocene to Eocene. Although not identified biostratigraphically, the base of the Holocene was tentatively assigned in Hole U1352B at 1.2 m, the level at
which shell-rich greenish gray marly sands with a
sharp base overlie gray muds. The base of the Pleistocene was loosely constrained between Samples 317U1352B-57X-CC and 61X-CC (491.74–525.34 m).
The base of the Pliocene was constrained with planktonic foraminiferal evidence between Samples 317U1352C-72R-CC and 73R-CC (1266.38–1283.95 m).
The base of the Miocene was picked using calcareous
nannofossil evidence between Samples 317-U1352C139R-CC and 140R-CC (1848.49–1852.71 m), and
the base of the Oligocene was placed between Samples 317-U1352C-146R-CC and 147R-CC (1903.29–
1916.63 m). Calcareous nannofossil and planktonic
Proc. IODP | Volume 317
Site U1352
foraminiferal dating indicate a late Eocene age of
35.2–36.0 Ma at the bottom of the hole (Sample 317U1352C-148R-CC [1924.26 m]).
At least five biostratigraphically defined hiatuses
were recognized. The first hiatus occurs between the
Pliocene and Pleistocene (Samples 317-U1352B-57XCC and 61X-CC [491.74–504.14 m]), where most, if
not all, of the late Pliocene is missing. An intra-late
Miocene unconformity was recognized between the
New Zealand upper Kapitean and lower Tongaporutuan Stages (Samples 317-U1352C-90R-CC and 91RCC [1394.62–1409.66 m]), where at least 5 m.y. is
missing. Another hiatus was identified between the
middle and late Miocene (Samples 317-U1352C101R-CC and 102R-CC [1486.78–1496.50 m]), with
at least 1.3 m.y. missing. A substantial unconformity,
identified as the Marshall Paraconformity, was constrained between Samples 317-U1352C-139R-CC and
140R-CC (1848.49–1852.71 m), where the lower
Miocene unconformably overlies the lower Oligocene. At least 12 m.y. is missing at this level. A hiatus
was also recognized between the early Oligocene and
late Eocene (Samples 317-U1352C-146R-CC and
147R-CC [1903.29–1916.63 m]), where at least 2.3
m.y. is missing.
Calcareous nannofossils
All core catcher samples from Holes U1352A–
U1352D contained calcareous nannofossils, with the
exception of Sample 317-U1352B-71X-CC (614.56
m). Abundances ranged from barren to very abundant, and preservation was generally good in the
Pleistocene–Miocene and moderate to poor in the
Oligocene and Eocene (Table T6). Nannofossil datums used for age determination at this site are summarized in Table T5.
Holocene–Pleistocene
All sediment samples from Hole U1352A were zoned
in NN21 (0–0.29 Ma). Core catcher samples from
Holes U1352A (317-U1352A-1H-CC through 5H-CC
[4.21–43.06 m]) and U1352D (317-U1352D-1H-CC
through 14H-CC [3.55–127.61 m]) contained rare
Emiliania huxleyi, likely placing them below the
acme of this species in Subzone NN21a (0.08–0.29
Ma).
Hole U1352B contained a thick Pleistocene section
from Samples 317-U1352B-1H-CC through 61X-CC
(7.93–525.34 m). Samples 317-U1352B-1H-CC
through 12H-CC (7.93–112.82 m) were zoned in
NN21a. Sediments recovered from Samples 317U1352B-13H-CC through 17H-CC (121.1–155.99 m)
were zoned in NN20 (0.29–0.44 Ma).
22
Expedition 317 Scientists
Site U1352
The highest occurrence (HO) of Pseudoemiliania lacunosa (top of Zone NN19) was recognized between
Samples 317-U1352B-17H-CC and 18H-CC (155.99–
164.18 m). Within the P. lacunosa and small Gephyrocapsa subzones of NN19 (Gartner, 1977), several nannofossil bioevents were observed:
3.7 Ma), defining the base of the middle Pliocene.
The base of the Pliocene was picked between Samples 317-U1352C-72R-CC and 73R-CC (1266.38–
1283.95 m; 5.33 Ma) with planktonic foraminiferal
evidence.
• The highest common occurrence (HCO) of Reticulofenestra asanoi between Samples 317-U1352B29H-CC and 30H-CC (257.09–266.92 m; 0.91 Ma),
Miocene
• The lowest common occurrence (LCO) of R. asanoi
between Samples 317-U1352B-34H-CC and 35HCC (292.54–295.24 m; 1.14 Ma),
• The HO of Gephyrocapsa >6.5 µm between Samples
317-U1352B-42X-CC and 43X-CC (348.59–360.08 m;
1.24 Ma), and
• The HO of Gephyrocapsa >5.5 µm between Samples
317-U1352B-43X-CC and 44X-CC (360.08–369.84 m;
1.26 Ma).
The top of the Helicosphaera sellii NN19 Subzone was
tentatively picked between Samples 317-U1352B46X-CC and 47X-CC (388.95–398.56 m; 1.34 Ma).
The lowest occurrence (LO) of Gephyrocapsa >5.5 µm
was observed within this subzone between Samples
317-U1352B-49X-CC and 50X-CC (412.30–427.34 m;
1.56 Ma).
The primary marker for the Calcidiscus macintyrei
NN19 Subzone (C. macintyrei >11 µm) was not observed; however, secondary markers (LO of Gephyrocapsa >4 µm between Samples 317-U1352B-54X-CC
and 55X-CC [463.67–469.84 m; 1.69 Ma] and LO of
Gephyrocapsa caribbeanica between Samples 317U1352B-55X-CC and 56X-CC [469.84–484.83 m;
1.73 Ma]) support the presence of this subzone.
The Pliocene/Pleistocene boundary was constrained
between Samples 317-U1352B-57X-CC and 61X-CC
(491.74–525.34 m) using nannofossil and foraminiferal markers that approximate the boundary. Below
this boundary, Reticulofenestra ampla was observed in
Sample 317-U1352B-62X-CC (542.58 m). The HO of
this species is dated at 2.78 Ma (Kameo and Bralower,
2000), which suggests a hiatus spanning the late
Pliocene.
Pliocene
Nannofossil biostratigraphy was problematic for
Samples 317-U1352B-60X-CC through 94X-CC
(514.61–821.74 m [total depth]) and 317-U1352C2R-CC through 93R-CC (576.47–1419.06 m) because
almost all standard zonal markers were absent. The
Reticulofenestra lineage, however, was abundant in
this cored succession, making it useful as a secondary proxy for age constraint. The HO of Reticulofenestra pseudoumbilicus was observed between Samples
317-U1352C-29R-CC and 30R-CC (884.91–894.37 m;
Proc. IODP | Volume 317
Standard Miocene nannofossil zonal markers based
on warm-water taxa were sparse at Site U1352, except in the early Miocene, where they were more
common. Planktonic foraminifers and bolboformids
show evidence of a major intra-late Miocene hiatus
between Samples 317-U1352C-90R-CC and 91R-CC
(1394.62–1409.66 m), where at least 5 m.y. is missing. The HO of Coccolithus miopelagicus (11.02 Ma)
was observed between Samples 317-U1352C-94R-CC
and 95R-CC (1438.43–1446.91 m). The first in situ
discoaster was noted in Sample 103R-CC (1515.94 m).
Below this level, discoasters (common in low- to
mid-latitudes) occurred sporadically and were dominated primarily by Discoaster deflandrei; however,
zonal marker species for the late Miocene were still
absent.
The HO of Calcidiscus premacintyrei was observed between Samples 317-U1352C-101R-CC and 102R-CC
(1486.78–1496.50 m), implying a late middle Miocene age (12.45 Ma). In addition, a marked increase
in six-rayed discoasters was observed in Samples 317U1352C-119R-CC and 120R-CC (1670.41–1669.36
m). Some of these were confidently identified as Discoaster deflandrei, but others were too overgrown to
be distinguished at the species level. The interval between these samples was tentatively considered as
the acme event of D. deflandrei (15.80 Ma), even
though the position of this event is uncertain because of caved-in local sediments found between the
two samples.
The lower Miocene HO of Sphenolithus heteromorphus
was picked between Samples 317-U1352C-114R-CC
and 115R-CC (1622.61–1632.61 m). Below Core 317U1352C-115R, S. heteromorphus was consistently seen
in core catcher samples down to Sample 317U1352C-129R-CC (1749.66 m). The LCO of this species, dated at 17.71 Ma, was therefore placed between Samples 129R-CC and 130R-CC (1749.66–
1760.69 m).
Sphenolithus belemnos (HO = 17.95 Ma) occurred in
extreme paucity in cored sediments; however, the
HCO of this species was questionably picked between Samples 317-U1352C-132R-CC and 133R-CC
(1777.59–1789.60 m), and it is certainly common by
Sample 135R-CC (1810.40 m). The LO of this species
was questionably picked between Samples 317U1352C-137R-CC and 138R-CC (1829.75–1841.54
m; 19.03 Ma).
23
Expedition 317 Scientists
A major hiatus, identified as the Marshall Paraconformity, occurred between Samples 317-U1352C139R-CC and 140R-CC (1848.49–1852.71 m). This
hiatus separates the early Miocene (18–19 Ma) from
the early Oligocene (30.1–32.0 Ma), with ~12 m.y.
missing based on nannofossil and foraminiferal biostratigraphy.
Oligocene
Nannofossil abundance was generally low in Oligocene sediments, and preservation was generally poor.
Samples 317-U1352C-140R-CC through 143R-CC
(1852.71–1878.85 m) were constrained between 30.0
and 32.0 Ma (early Oligocene) based on the presence
of Dictyococcites scrippsae, Dictyococcites stavensis, Reticulofenestra filewiczii, and Chiasmolithus altus.
The HO of Reticulofenestra umbilicus (32.0 Ma) was
identified between Samples 317-U1352C-143R-CC
and 144R-CC (1878.85–1885.85 m). The HO of Isthmolithus recurvus was observed between Samples 317U1352R-145R-CC and 146R-CC (1893.50–1903.29 m;
32.5 Ma).
A hiatus spanning at least 2.3 m.y. was recognized
with calcareous nannofossil evidence between the
early Oligocene and late Eocene (Samples 317-U1352C146R-CC and 147R-CC [1903.29–1916.63 m]). Sediments above the hiatus were dated at 32.5–32.9 Ma,
and those below were dated at 35.2–36.6 Ma.
Eocene
Calcareous nannofossil abundances were generally
low in the Eocene, and preservation was better than
in the Oligocene. Samples 317-U1352C-147R-CC
and 148R-CC (1916.63–1924.26 m [total depth])
contained the late Eocene species Reticulofenestra reticulata, Isthmolithus recurvus, and Chiasmolithus
oamaruensis, which constrains the age of the bottom of the hole between 35.2 and 36.6 Ma.
Planktonic foraminifers and bolboformids
Planktonic foraminiferal and bolboformid biostratigraphy at Site U1352, on the upper slope, was based
on the examination of core catcher samples from
Holes U1352A–U1352D (Tables T7, T8, T9, T10, T11,
T12, T13, T14). The absolute ages assigned to biostratigraphic datums follow the references listed in
Table T3 in the “Methods” chapter. The percentage
of planktonic foraminifers in the cored Holocene–
Eocene succession generally increased from ~70% in
the upper part of the succession to ~95% in the
lower part (Fig. F25). Preservation was variable but
was generally good in the Pleistocene, moderate in
the Pliocene, poor in the Miocene and Oligocene,
and moderate to poor in the Eocene. Bolboformids
Proc. IODP | Volume 317
Site U1352
only occurred in the late middle Miocene to early
late Miocene.
Holocene
Although not identified biostratigraphically, the base
of the Holocene was tentatively correlated with a distinct lithologic boundary at Section 317-U1352B-1H1, 120 cm (1.20 m). Mudline samples from Sections
317-U1352A-1H-1, 0 cm (0.00 m), and 317-U1352D1H-1, 0 cm (0.00 m), contained temperate assemblages and were dominated by the eutrophic species
Globigerina bulloides s.l. They also contained abundant Globoconella inflata and Turborotalita quinqueloba and common Truncorotalia truncatulinoides
(sinistral), Neogloboquadrina incompta, and Orbulina
universa.
Pleistocene
The abundance of planktonic foraminifers varied
throughout the Pleistocene sections of Holes
U1352A, U1352B, and U1352D, but planktonic
forms increased on average from 66% of the total
foraminiferal assemblage in the upper part of the
section to 74% in the lower part. This is consistent
with deposition under suboceanic conditions on the
upper slope. Preservation was generally good, and
planktonic assemblages were dominated by small,
thin- to thick-walled specimens. Larger thick-walled
forms were found in some samples, especially in
greenish gray sandy marls, where temperate species
were more common.
The Pleistocene succession included several useful
planktonic foraminiferal bioevents, although primary age control was provided by calcareous nannofossils. Hirsutella hirsuta (0–0.34 Ma) was found in
Samples 317-U1352B-19H-CC (173.71 m) and 317U1352D-14H-CC (130.26 m). Truncorotalia truncatulinoides (0–1.1 Ma) was present in most samples in the
upper part of the succession (Samples 317-U1352A1H-CC through 4H-CC [4.21–33.43 m], 317-U1352B1H-CC through 19H-CC [7.93–173.71 m], and 317U1352D-1H-CC through 14H-CC [3.55–127.61 m]).
The HO of Globoconella puncticuloides was recognized
as a very well defined event in Sample 317-U1352B22H-CC (198.87 m), ~20 m below the LO of the
Haweran marker species Hirsutella hirsuta. It occurred
persistently below this level down to the early Pliocene. The HO of Gc. puncticuloides was dated at 1.5
Ma at ODP Site 1123 (Cooper, 2004), but in Hole
U1352B calcareous nannofossil and diatom dating
suggests it occurs between 0.44 and 0.70 Ma. This
suggests the disappearance of this species occurred
~1 m.y. later in the cold waters of the Canterbury Basin than in the warmer subtropical water surrounding Site 1123. Globoconella cf. puncticuloides, a dis24
Expedition 317 Scientists
tinctly inflated form with a flattened dorsal surface,
occurred in most samples with Gc. puncticuloides between Samples 317-U1352B-24H-CC and 94X
(217.96–821.74 m) and 317-U1352C-7R-CC and
43R-CC (671.20–1019.69 m). This distinctive form
has a pustulate ultrastructure similar to Truncorotalia
crassaformis, but it is distinguished from the latter
species by its low-arched rather than slitlike aperture.
The abundance of the subantarctic species Neogloboquadrina pachyderma was highly variable in the Pleistocene succession. This cold-water species was common and sometimes dominant in gray mud
lithologies and was often associated with high numbers of Gc. puncticuloides. Interbedded greenish gray
sandy marls contained lower abundances of Nq.
pachyderma, and the temperate species Neogloboquadrina incompta and the eutrophic species Globigerina bulloides s.l. and Turborotalita quinqueloba were
generally more common. Subtropical species were
absent except for rare occurrences of Globigerinella
aequilateralis in Samples 317-U1352D-7H-CC (60.32
m) and 317-U1352B-50X-CC (427.34 m) and a single
specimen of Globigerinoides cf. ruber in Sample 317U1352B-21H-CC (189.06 m).
Pliocene
Tentative biostratigraphic evidence indicates that the
late Pliocene is missing at Site U1352 (see discussion
below).
Planktonic foraminifers were abundant and moderately well preserved in the middle Pliocene section
in Samples 317-U1352B-61X-CC through 94X-CC
(525.34–821.74 m) and 317-U1352C-2R-CC through
29R-CC (576.47–884.91 m) and moderate to poorly
preserved in the early Pliocene section in Samples
317-U1352C-30R-CC through 72R-CC (894.37–
1266.38 m). The abundance of planktonic foraminifers varied, but on average abundances increased
down through the Pliocene from ~74% at the top of
the section to 85% at the base. Such abundances are
consistent with deposition under suboceanic conditions on the upper to middle slope.
Pliocene planktonic assemblages were dominated by
Globigerina bulloides and other small species of Globigerina. Globoconella inflata, Gc. puncticuloides, Gc. cf.
puncticulata, Gc. puncticulata s.s. (below Sample 317U1352C-56R-CC [1147.41 m]), Neogloboquadrina
pachyderma, Nq. incompta, Turborotalita quinqueloba,
and related forms were common to abundant in
some samples. Orbulina universa, Truncorotalia crassaformis, and Tr. crassaconica also occurred sporadically
with rare Globigerinita glutinata, Globoconella cf.
pliozea, and Zeaglobigerina woodi.
Biostratigraphically useful middle Pliocene events include the LO of Truncorotalia crassula between SamProc. IODP | Volume 317
Site U1352
ples 317-U1352B-60X-CC and 61X-CC (514.61–
525.34 m; 2.40 Ma) and the poorly constrained HO
of Truncorotalia crassaconica? between Samples 317U1352B-60X-CC and 61X-CC (514.61–525.34 m;
>3.09 Ma). The juxtaposition of these events suggests
the presence of a hiatus at this level with at least 0.7
m.y. missing. The HO of calcareous nannofossil Reticulofenestra ampla (2.78 Ma) between Samples 317U1352B-61X-CC and 62X-CC (525.34–542.58 m)
and the HO of planktonic foraminifer Zeaglobigerina
woodi (>2.7 Ma) between Samples 317-U1352B-67XCC and 68X-CC (573.33–583.56 m) support the presence of a hiatus, but it is uncertain whether the upper Pliocene is represented in part or is completely
missing.
Early Pliocene bioevents include the HO of Globoconella subconomiozea between Samples 317-U1352C26R-CC and 27R-CC (855.29–864.94 m; 3.35 Ma),
the LO of Gc. inflata s.s. and the HO of Gc. puncticulata s.s. between Samples 317-U1352C-56R-CC and
57R-CC (1147.41–1155.03 m; 4.3 Ma), and the HO of
Gc. pliozea between Samples 317-U1352C-61R-CC
and 62R-CC (1197.57–1206.27 m; 4.49 Ma). The LO
of Gc. puncticulata s.s. (5.30 Ma) and the HO of
Globoconella sphericomiozea s.s. (5.30 Ma), which
serve as proxies for the Miocene/Pliocene boundary,
were reliably identified between Samples 317U1352C-72R-CC and 73R-CC (1266.38–1283.95 m).
Miocene
Planktonic foraminifers in Hole U1352C were abundant and moderately to poorly preserved in the late
Miocene (Samples 317-U1352C-73R-CC through
101R-CC [1283.95–1486.78 m]), abundant to common and poorly preserved in the middle Miocene
(Samples 317-U1352C-102R-CC through 121R-CC
[1496.50–1678.60 m]), and common to few and
poorly preserved in the early Miocene (Samples 317U1352C-122R-CC through 139R-CC [1690.44–
1848.49 m]). Foraminiferal tests were often recrystallized and infilled with sparry calcite, and some also
exhibited microcrystalline overgrowths (Fig. F26).
Planktonic foraminiferal abundances generally increased down through the Miocene succession of
Hole U1352C from 85% of the total foraminiferal assemblage at the top of the section to 94% at the base.
Such abundances are consistent with deposition on
the lower slope and base of slope under suboceanic
and oceanic conditions.
Planktonic assemblages in the Miocene succession
were characterized by abundant Globigerina, including Gg. bulloides. Zeaglobigerina woodi and Turborotalita quinqueloba and related forms were also present
along with rare Globigerinopsis obesa and Globigerinita
glutinata. Globoconellids, including Gc. sphericomiozea
25
Expedition 317 Scientists
(upper Kapitean; uppermost Miocene), Gc. miotumida
(Tongaporutuan to upper Waiauan; upper Miocene),
Gc. miozea (Lillburnian to Altonian; middle to upper
lower Miocene), Gc. zealandica (middle Altonian;
middle lower Miocene), and Gc. praescitula (early Altonian; lower Miocene), were present in respective
parts of the Miocene section, but abundances varied.
Neogloboquadrina pachyderma and Nq. incompta were
present in the late Miocene, Paragloborotalia mayeri
s.l. was abundant in the early late Miocene, Orbulina
universa occurred sporadically down into the middle
Miocene, and Orbulina suturalis was present in the
early middle Miocene. Subtropical species were absent, except for single specimens of Globigerinoides
trilobus in the middle Miocene (Sample 317-U1352C105R-CC [1534.20 m]) and lower Miocene (Sample
317-U1352C-128R-CC [1743.51 m]).
The uppermost Miocene (upper Kapitean) species Gc.
sphericomiozea s.s. (5.30–5.60 Ma) was common from
Samples 317-U1352C-73R-CC to 90R-CC (1283.95–
1394.62 m). Its presence immediately above the HOs
of Globoconella miotumida (7.07 Ma), Bolboforma
metzmacheri s.s. (8.85 Ma), and Bolboforma subfragoris
s.l. (10.58 Ma) suggests a hiatus in the late Miocene
section between Samples 317-U1352C-90R-CC and
91R-CC (1394.63–1409.66 m) with at least 3.25 m.y.
missing. The concurrence of these two bolboformid
species is also somewhat enigmatic because they
have range zones that are separated in time, which
suggests there may be another shorter hiatus between Samples 317-U1352C-91R-CC and 93R-CC
(1409.66–1419.06 m).
The HCO of Paragloborotalia mayeri s.l. (10.97 Ma)
occurred between Samples 317-U1352C-95R-CC and
96R-CC (1446.91–1448.00 m) and the LO of Bolboforma subfragoris s.l. (11.64 Ma) occurred between
Samples
317-U1352C-101R-CC
and
102R-CC
(1486.78–1496.50 m). The HO of Globoconella conica
(12.98 Ma) was noted immediately below the LO of
B. subfragoris s.l., which suggests a hiatus between
Samples
317-U1352C-101R-CC
and
102R-CC
(1486.78–1496.50 m). Bolboforma cf. reticulata, a bolboformid that has never before been recognized in
New Zealand, occurred at the same level as the HO of
Gc. conica and was not recorded in any other sample.
Sample 317-U1352C-103R-CC (1515.94 m) is also
significant in that it contained the HO of Gc. miozea,
the LO of Orbulina universa (13.63 Ma), and B. cf. robusta, one sample above the HO of Or. suturalis in
Sample 104R-CC (1525.67 m). The close stratigraphic spacing of these bioevents between Samples
317-U1352C-103R-CC and 104R-CC (1515.94–
1525.67 m) suggests that the section is condensed or
that parts of the section are missing.
Proc. IODP | Volume 317
Site U1352
The LO of Globoconella miozea (16.7 Ma), a bioevent
that marks the base of the upper Altonian, was identified between Samples 317-U1352C-125R-CC and
126R-CC (1714.42–1725.45 m; 16.7 Ma). Other bioevents include the poorly constrained HO of Globoconella praescitula (16.7 Ma) between Samples 317U1352C-123R-CC and 124R-CC (1697.39–1707.63
m), the HO of Gc. zealandica (16.7 Ma) between Samples 317-U1352C-125R-CC and 126R-CC (1714.42–
1725.45 m), the HO of Gc. incognita (18.3 Ma) between Samples 317-U1352C-131R-CC and 132R-CC
(1769.18–1777.59 m), and the poorly constrained HO
of Zeaglobigerina connecta (18.5 Ma) between Samples
317-U1352C-135R-CC and 136R-CC (1810.4–1819.56
m). The early Miocene marker species Catapsydrax
dissimilis was also tentatively identified in Samples
317-U1352C-133R-CC (1789.6 m) and 135R-CC
(1810.40 m). The range of this species is reported by
Cooper (2004) to extend to the top of the lower Miocene Otaian Stage (18.7 Ma), but its HO in Hole
U1352C appeared to be slightly younger. This is consistent with unpublished data from ODP Site 1171
(50°S) on the South Tasman Rise, where Cs. dissimilis
occurs in conjunction with the early Miocene (lower
Altonian) species Gc. praescitula (M.P. Crundwell,
pers. comm., 2009).
Oligocene
The top of the Oligocene was identified using calcareous nannofossil evidence between Samples 317U1352C-139R-CC and 140R-CC (1848.49–1852.71
m). This coincides with a significant change in
planktonic foraminiferal assemblages at the same
level. The HO of the early Oligocene (lower Whaingaroan) marker species Subbotina angiporoides was
identified between Samples 317-U1352C-140R-CC
and 141R-CC (1852.71–1862.57 m). These biostratigraphic events suggest the presence of a major hiatus
(Marshall Paraconformity) between Samples 317U1352C-139R-CC and 140R-CC (1848.49–1852.71
m), where at least 12 m.y. is missing.
Foraminiferal assemblages in the Oligocene (Samples
317-U1352C-140R-CC through 146R-CC [1852.71–
1903.29 m]) are characterized by low abundances
and poor preservation, which is attributed to the cemented nature of the Oligocene limestone. Planktonic foraminifers in the Oligocene composed >95%
of the total foraminiferal assemblage, except in Sample 317-U1352C-144R-CC (1885.85 m). Such abundances are consistent with basin-floor deposition under open oceanic conditions. Planktonic assemblages
include Zeaglobigerina euapertura, Subbotina angiporoides, Globorotaloides suteri, and Zg. brevis?
26
Expedition 317 Scientists
Eocene
The top of the Eocene is marked by the HO of the
late Eocene (Runangan) marker species Globigerinapsis index (34.4–36.0 Ma) between Samples 317U1352C-146R-CC and 147R-CC (1903.29–1916.63
m). The HO of this species coincides with a change
in lithology to less cemented chalky limestone and
more common, better preserved foraminiferal assemblages. Calcareous nannofossil dating (Table T5) suggests the presence of a hiatus between the early Oligocene and the late Eocene with at least 2.3 m.y.
missing.
Eocene planktonic foraminiferal assemblages are
characterized by abundant Globigerinapsis index,
along with Subbotina minima, Tenuitella gemma?, and
Turborotalita ciperoensis. Planktonic foraminifers and
calcareous nannofossils indicate a late Eocene age of
35.2–36.0 Ma at the bottom of the hole.
Benthic foraminifers
Benthic foraminifers from 242 core catcher samples
from Holes U1352A–U1352D were examined (Table
T15). Benthic foraminifer abundances varied
throughout the cored section, ranging from rare to
dominant. Preservation was generally good in the
Pleistocene, poor to moderate in the Pliocene and
Miocene, and poor in the Oligocene and Eocene.
Eleven benthic foraminiferal bioevents were identified at Site U1352 (Table T5). The calibration of benthic bioevents, in general, was poorly constrained,
but these bioevents provided useful age control
where planktonic foraminifers and calcareous nannofossils were poorly represented.
Site U1352
64R-CC and 65R-CC (1217.80–1222.21 m). The stratigraphic suppression of this species relative to
planktonic events suggests that the disappearance of
H. mioindex in Hole U1352B might have been controlled by water depth.
Miocene
The LO of Uvigerina pliozea (5.30 Ma) was observed
between Samples 317-U1352C-84R-CC and 85R-CC
(1342.48–1352.39 m), ~50 m above the Miocene/
Pliocene boundary, located between Samples 317U1352C-72R-CC and 73R-CC (1266.38–1283.95 m;
5.33 Ma) using planktonic foraminiferal evidence.
Other useful benthic foraminifer bioevents include a
single record of Loxostomum truncatum (8.95–12.76
Ma) in Sample 317-U1352C-94R-CC (1438.43 m),
the LO of Notorotalia taranakia (11.01 Ma) between
Samples 317-U1352C-95R-CC and 96R-CC (1446.91–
1448.00 m), occurrences of Notorotalia wilsoni (11.01
Ma) in Samples 317-U1352C-95R-CC and 96R-CC
(1446.91–1448.00 m), and the LO of H. morgani
(~11.01 Ma) between Samples 317-U1352C-95R-CC
and 96R-CC (1446.91–1448.00 m). The LO of H. mioindex (15.10 Ma) was tentatively recognized between
Samples
317-U1352C-104R-CC
and
105R-CC
(1525.67–1534.20 m). These events generally agree
with other microfossil dating.
Oligocene–Eocene
Sample 317-U1352C-147R-CC (1916.63 m) contained the late Eocene to early Oligocene species
Cibicidoides parki.
Paleowater depths
Holocene–Pleistocene
The HO of Proxifrons advena (~0.4 Ma; Table T5) was
identified between Samples 317-U1352B-14H-CC
and 15H-CC (130.26–141.17 m). The HO of Bolivinita pliozea (0.6 Ma) was identified between Samples
317-U1352B-16H-CC and 17H-CC (150.74–155.99
m). The latest Pleistocene–Holocene (Haweran) restricted species, Loxostomum karrerianum (0–0.34
Ma), was recognized in Samples 317-U1352D-1H-CC
(3.55 m) and 8H-CC (70.00 m) and 317-U1352B20H-CC (180.38 m). These ages are generally consistent with those from calcareous nannofossils and
planktonic foraminifers.
Pliocene
The HO of Haeuslerella morgani (3.62–5.30 Ma) was
identified between Samples 317-U1352B-64X-CC
and 65X-CC (557.28–562.35 m). The HO of the deep
middle bathyal species Hopkinsina mioindex (3.62
Ma) was recognized between Samples 317-U1352CProc. IODP | Volume 317
Paleowater depths were estimated by the analysis of
benthic foraminiferal assemblages from Holes
U1352A–U1352D (Table T15; Fig. F27). Terminology
used for paleowater depth estimates is given in Figure F7 in the “Methods” chapter.
Holocene–middle Pliocene sediments were dominated by inner to outer shelf benthic foraminiferal
taxa, although the rare but persistent presence of
bathyal marker species suggests the shelfal taxa composing the bulk of the benthic assemblage were
transported downslope. Paleodepths generally increased downhole to lower bathyal depths in the
lower part of the cored succession. Middle to deep
bathyal taxa occurred in lower Pliocene and older
sediments, which is consistent with a general downhole increase in the abundance of planktonic foraminifers through the progradational foreset sequence into the bottom sets and basin-floor facies. It
also coincides with the change from suboceanic to
open oceanic conditions.
27
Expedition 317 Scientists
Estimates of paleowater depths, based on the general
composition of benthic foraminiferal assemblages,
ranged from outer shelf to upper bathyal during the
middle–late Pleistocene. The estimated upper depth
limits are shallower than seismic data interpretation
at this site suggests, which is consistent with the
bulk of the assemblage being reworked downslope.
The subtidal to inner shelf species Elphidium charlottense and Notorotalia aucklandica were abundant in
upper Pleistocene Samples 317-U1352A-1H-CC
through 5H-CC (4.21–43.06 m) and 317-U1352D1H-CC through 10H-CC (3.55–88.92 m). These species occurred consistently with the outer shelf to uppermost bathyal species Globocassidulina canalisutulata (Eade, 1967) down to Sample 317-U1352D-14HCC (127.61 m). In Samples 317-U1352B-1H-CC
through 31H-CC (7.93–272.44 m), outer shelf species
G. canalisutulata and Nonionella flemingi were dominant and the upper bathyal indicator Trifarina angulosa was abundant, particularly between Samples
317-U1352B-22H-CC and 28H-CC (198.87–250.70
m).
In the lower Pleistocene to upper middle Pliocene
section between Samples 317-U1352B-32H-CC and
65X-CC (280.84–562.35 m), the outer shelf species
Nonionella flemingi and Notorotalia profunda were
dominant. Uvigerinids that characterize deeper water, including Uvigerina peregrina and Trifarina angulosa, were present throughout most of this section
and indicate deposition in upper bathyal water
depths. The middle to lower bathyal species Bulimina
exilis (van Morkhoven et al., 1986) was also present
sporadically between Samples 317-U1352B-38X-CC
and 48X-CC (310.85–408.18 m), suggesting even
deeper bathyal water depths in this interval.
In the lower Pliocene section between Samples 317U1352B-66X-CC and 94X-CC (572.40–821.74 m)
and 317-U1352C-2R-CC and 65R-CC (576.47–
1222.21 m), the outer shelf species Notorotalia profunda was abundant and was associated with upper
bathyal uvigerinids, including Uvigerina peregrina, Astrononion spp., and Bolivinita pliozea. Notorotalia
taranakia was also abundant in this same interval;
however, the ecological preferences of this species
are not fully understood. Melonis pompilioides (an extant deep lower bathyal species) occurred sporadically between Samples 317-U1352C-2R-CC and 65RCC (576.47–1222.21 m), becoming more common in
the lowermost Pliocene between Samples 317U1352C-56R-CC and 64R-CC (1147.41–1217.80 m).
Paleodepth markers that are considered to be in situ
suggest that paleowater depths increased downhole
from outer shelf–uppermost bathyal depths in the
uppermost part of the lower Pliocene section to
Proc. IODP | Volume 317
Site U1352
lower bathyal depths in the lowermost part of the
lower Pliocene section.
Assemblages in the lowermost Pliocene to uppermost Miocene section, between Samples 317U1352C-65R-CC and 88R-CC (1222.21–1379.88 m),
are characterized by a decrease in shelfal species and
a reciprocal increase in upper bathyal marker species,
including Cibicidoides neoperforatus and Uvigerina
spp.
The upper Miocene to upper middle Miocene section
between Samples 317-U1352C-89R-CC and 104R-CC
(1386.04–1525.67 m) was dominated by benthic
taxa indicative of middle bathyal water depths, including Karreriella bradyi and Sigmoilopsis schlumbergeri, and occasional specimens of the deep middle
bathyal species Hopkinsina mioindex.
In the middle Miocene section between Samples
317-U1352C-105R-CC and 112R-CC (1534.20–
1597.08 m), benthic foraminifers were poorly preserved and too sparse for reliable estimates of paleodepths based on benthic assemblages. Deep middle to deep lower bathyal water depths are inferred,
however, based on the presence of Cibicidoides robertsonianus and Melonis dorreeni, which is morphologically similar to the extant deep lower bathyal species
M. pompilioides. Benthic foraminifers were very
sparse to absent in samples from the middle Miocene
to Eocene (Samples 317-U1352C-113R-CC through
148R-CC; 1611.72–1924.26 m).
Diatoms
Samples 317-U1352A-1H-CC through 5H-CC (4.21–
43.06 m) were barren of diatoms (Table T16).
Core catcher samples from Hole U1352B were examined for diatoms (Table T16). In the majority of the
50 samples containing diatoms, abundances were
low, and preservation was moderate to poor. Several
robust diatom species with thickly silicified valves
(e.g., Paralia sulcata, Thalassionema nitzschoides, and
Chaetoceros resting spores) were better preserved
than delicate forms with weakly silicified valves.
The HO of marker species Fragilariopsis fossilis (0.70
Ma) was identified in Sample 317-U1352B-23H-CC
(208.46 m), although this species occurred only sporadically below this level and the reliability of its HO
is uncertain. A single record of Actinocyclus karstenii
(HO 1.73 Ma) was also identified in Sample 317U1352B-55X-CC (469.84 m). The distribution of
these species is consistent with calcareous nannofossil dating.
Most diatom assemblages in Hole U1352B contained
marine resting spores of the genus Chaetoceros,
28
Expedition 317 Scientists
Site U1352
which is indicative of high productivity in nearshore upwelling regions and coastal areas. Also included was a mixture of coastal, brackish, and freshwater species, which suggests marine deposition
with an influence from coastal upwelling and some
river input.
from Hole U1352B (Figs. F28, F29, F30) and are
therefore not shown. The deepest record from Site
U1352 was from Hole U1352C, which overlaps the
record from Hole U1352B for ~200 m (Fig. F31).
Core catcher samples from Hole U1352C were examined for diatoms, and their distribution is reported in
Table T15. Diatoms were only found in six middle
Pliocene samples from this hole, between Samples
317-U1352C-7R-CC and 12R-CC (671.20–719.46 m).
Preservation was moderate to poor. Cyclotella, a
genus diagnostic of brackish and freshwater,
occurred in greater abundances than other taxa in
these samples. Other diatoms included coastal and
brackish water species. The diatom assemblage
shows evidence of some river and/or estuarine input.
Hole U1352B was cored to 246.2 m using the APC
system with both nonmagnetic core barrels (Cores
317-U1352B-1H through 27H [246.2 m]) and magnetic core barrels (Cores 317-U1352B-28H through
36H [246.2–297.0 m]). The remainder of the hole
was cored using the XCB system. Distinct changes in
magnetic behavior were observed where coring systems were changed (Fig. F28). In the nonmagnetic
APC cores, NRM inclinations are usually steeply positive (~70°) with occasional intervals of shallow or
negative inclination. After demagnetization at 20
mT, inclinations are steep and negative (approximately –60°). Declination varies between cores but is
consistent, both before and after demagnetization,
within individual cores.
All core catcher samples from Samples 317-U1352D1H-CC through 14H-CC were examined for diatoms
(3.55–127.61 m) (Table T15). Diatoms were found in
7 of the 14 samples examined and were relatively
rare. The assemblages and preservation were similar
to those of samples from the equivalent section in
Hole U1352B. The presence of a single record of
Hemidiscus karstenii (HO 0.30 Ma) in Sample 317U1352D-12H-CC (108.55 m) is consistent with calcareous nannofossil dating.
Macrofossils
Macrofossils were examined in cored sediments from
all Site U1352 holes. Provisional identification, age,
and habitat preference are provided in Table T17.
Hole U1352B
Declination variability between cores suggests that
this signal was not strongly overprinted by the drilling process (Richter et al., 2007). The first 18 APC
cores (317-U1352B-1H through 18H [0–165.7 m])
were oriented using the Flexit tool, which records
the orientation of the APC core barrel. Data from
this tool were used to orient the cores with respect to
true north (Fig. F30). Orienting core barrels brings
the declination of NRM after 20 mT alternating-field
(AF) demagnetization into good agreement between
cores. The mean orientation is a declination of 27.7°
and an inclination of –60.1°.
Section-half measurements
APC cores were recovered using magnetic core barrels between 246.2 and 297.0 m (Cores 317-U1352B28H through 36H). This depth interval shows an increase in NRM intensity from ~2 × 10–3 to ~9 × 10–3
A/m and a slight increase in susceptibility. Declinations are consistently grouped around north during
this interval, which comprises nine cores. Inclinations are fairly steeply positive throughout (~60°).
Remanence does not change after AF demagnetization and is interpreted to reflect a drilling overprint
that was not removed.
NRM was measured on all archive section halves
from Holes U1352A–U1352D unless the core material was too heavily disturbed by the drilling process.
A single demagnetization step at peak fields of 20 mT
was applied to all measured sections. The NRM intensities of the sediments typically range from 10–4
to 10–2 A/m and tend to decrease with depth. The
short records from Holes U1352A and U1352D are
comparable to and entirely overlapped by the record
The XCB system was used to core from 297 m to the
total depth of Hole U1352B (Cores 317-U1352B-37X
through 94X [830.9 m]). Intensity varies throughout
this interval. Declinations cluster close to north both
before and after AF demagnetization at 20 mT, suggesting unsuccessful removal of the drilling overprint. Inclinations are typically fairly steeply positive
(~55°) but do vary, with some cores showing shallower or negative inclinations.
Paleomagnetism
Paleomagnetic analyses at Site U1352 included routine measurement and partial demagnetization of
natural remanent magnetization (NRM) of archive
section halves and some discrete samples from the
working halves of cores. Rock magnetic experiments
were also performed on these discrete samples. All
depths in this section are reported in m CSF-A.
Proc. IODP | Volume 317
29
Expedition 317 Scientists
Hole U1352C
Hole U1352C was cored with the RCB system from
575 m to a total depth of 1928 m (Fig. F31). Both recovery and intensity are low downcore to ~900 m,
and a low signal to noise ratio is apparent downcore
to ~1200 m. Below this interval, NRM has a consistently steep positive inclination (~70°) and a northward declination before and after demagnetization.
This indicates a drilling overprint that was not removed. Intensity peaks at 1 × 10–2 A/m at ~1700 m
and drops significantly between 1850 and 1870 m
from ~1 × 10–3 A/m above to 1 × 10–4 A/m below. At
this point, data again show a wide spread as the
noise level of the SRM was approached. Demagnetization at 20 mT removed up to 80% of the intensity
of samples from Hole U1352C.
The boundaries between low-intensity intervals,
noisy intervals, and higher intensity intervals with
tightly grouped orientations correlate to lithologic
unit and subunit Boundaries IIA/IIB (1189 m) and
IIC/III (1852.64 m; Marshall Paraconformity), as described in “Lithostratigraphy.”
Discrete measurements
Pairs of cubes were extracted from the working half
of each core for shipboard analyses. Ten of these
pairs were selected to characterize the sediments in
Hole U1352B. Seven samples from lithologic Subunit
IIC, which was found only in Hole U1352C, were
also demagnetized with alternating fields.
NRM demagnetization
Seventeen samples were AF demagnetized up to 80
mT. The signal to noise ratio is poor, but drilling
overprints persist to ~30 mT. Some samples (e.g.,
70.33 m; Fig. F32A) show a characteristic component
that demagnetizes toward the origin. Others (e.g.,
255.00 m; Fig. F32C) appear to acquire a gyro-remanent magnetization (GRM) after 40 mT.
Ten samples were thermally demagnetized at 30°–40°
intervals up to 460°C (Fig. F32B, F32D). Susceptibility was measured after each heating step to monitor
any chemical alteration resulting in new magnetic
minerals (Fig. F32G). All samples increase in susceptibility by 460°C, and four samples (at 70.32, 105.84,
230.99, and 255.03 m) show a slight (5%–10%) decrease in susceptibility between 310° and 430°C. The
drilling overprint is removed by ~150°–200°C, and a
characteristic component can be resolved before
thermal alteration occurs.
Demagnetization of discrete samples confirmed that
a subvertical overprint persists in the samples be-
Proc. IODP | Volume 317
Site U1352
yond the 20 mT demagnetization step applied to section-half samples. This overprint was removed in
most cases by further demagnetization. Magnetic
mineralogy is not fully constrained, but the acquisition of a GRM in some samples at 40 mT and the
thermal alteration above 310°C suggest the presence
of some iron sulfides in the sediment.
IRM acquisition and demagnetization
An isothermal remanent magnetization (IRM) was
imparted to 10 samples from Hole U1352B in a stepwise manner up to 1 T (Fig. F33A) followed by backfield acquisition, which revealed fairly uniform behavior. All samples appear to saturate between 400
and 600 mT with coercivities of remanence between
40 and 75 mT.
The same 10 samples then had a 1 T IRM imparted
and were stepwise AF demagnetized up to 80 mT
(Fig. F33B–F33C). A clear grouping of significantly
higher saturation remanence is evident in four samples (at 70.33, 202.58, 221.68, and 326.96 m) (Fig.
F32B). This group also shows a softer magnetization
and loses a higher proportion of this magnetization
at 80 mT peak fields (Fig. F32C). The presence of
these two behaviors upon AF demagnetization of
IRM may suggest the presence of two low-coercivity
mineralogies and/or grain-size fractions. Median destructive fields are between 40 and 50 mT for all samples.
Magnetostratigraphy
A pervasive drilling overprint prevented magnetostratigraphic interpretation of much of Site U1352.
Biostratigraphic evidence suggests that the Brunhes/
Matuyama boundary lies between Samples 317U1352B-24H-CC and 30H-CC (217.9–266.9 m), constrained between the presence of Fragilariopsis fossilis
(diatom; HO = 0.70 Ma) and the HCO of Reticulofenestra asanoi (nannofossil; 0.91 Ma) in these two
samples, respectively. Some noisy intervals were encountered within this range, where inclinations are
not clearly negative (normal). From Core 317U1352B-28H (246.2 m) downward, magnetic core
barrels were used, and characteristic magnetizations
could not be isolated. Therefore, it was not possible
to identify the Brunhes/Matuyama boundary, despite a promising trend toward positive inclinations
throughout Core 317-U1352B-27H.
No further magnetostratigraphic constraints could
be given at this site because of the pervasive drilling
overprint. This overprint may eventually be removed
by the full demagnetization of working-half cube
samples onshore.
30
Expedition 317 Scientists
Physical properties
Gamma ray attenuation (GRA) densitometer bulk
density, magnetic susceptibility (loop sensor; MSL),
NGR, and P-wave velocity measured with the P-wave
logger (PWL) were measured on whole-round core
sections from Holes U1352A–U1352C. Additionally,
magnetic susceptibility (point sensor; MSP) and
spectrophotometry and colorimetry were measured
on section halves from the same holes. Whole-round
measurements on XCB and RCB cores were degraded
in quality because of drilling disturbance associated
with coring and the slightly smaller diameters of
XCB and RCB cores relative to the core liner. Discrete
P-wave velocity (measured using the P-wave caliper
[PWC] and P-wave bayonets [PWB]) and moisture
and density (MAD) were measured on section halves
from Holes U1352B and U1352C. Sediment strength
was determined only on section halves from Hole
U1352B. Unless otherwise noted, all depths in this
section are reported in m CSF-A.
Gamma ray attenuation bulk density
GRA bulk density was measured with the densitometer at 5 cm intervals (measurement time = 5 s). Below
1361.4 m, the sampling resolution was increased to
2.5 cm (Section 317-U1352C-87R-1). The raw data
range from –0.42 to 2.467 g/cm3 in Hole U1352B and
from –0.417 to 2.606 in Hole U1352C (Fig. F34). Low
values are assumed to represent cracks in the cores,
gas, and/or incomplete filling of the cores with sediment. Thus, only bulk density values >1.4 g/cm3
were compared with MAD-derived bulk density and
downhole GRA density estimates (Figs. F35, F36).
GRA bulk density records from Holes U1352B and
U1352C show the same long-term trends even where
XCB cores from Hole U1352B overlap RCB cores
from Hole U1325C.
A comparison of GRA densitometer and MAD data
from samples in Hole U1352B (Fig. F35) yields results similar to those from Hole U1351B. That is,
GRA data tend to be lower than, but roughly parallel
to, MAD bulk density estimates. Lower bulk density
estimates mainly reflect incomplete filling of the
core liner as well as cracks in the core caused by gas
expansion. These features do not affect MAD results.
In general, higher GRA bulk density values are consistent with MAD estimates, and both methods reveal a trend toward higher bulk densities with depth.
The increase and broadening of bulk density estimates across the lithologic Subunit IB/IC boundary
(Fig. F35A–F35B) may reflect the sampling of increasingly calcite-cemented soft sediments.
In Hole U1352C, both MAD and GRA bulk density
estimates generally increase with depth (Fig. F36). As
Proc. IODP | Volume 317
Site U1352
the sediments become increasingly cemented, filled
core liners become increasingly rare. In a few locations, GRA estimates reach those obtained by MAD
analyses. The two trends of MAD data between 550
and 1200 m (Fig. F36A–F36B) are a result of cementation, with the higher MAD bulk densities reflecting
more highly indurated strata (see “Moisture and
density”). GRA data in this region (lithologic Subunit IIA; see “Lithostratigraphy”) include measurements of both denser cemented sediments and less
dense uncemented sediments. Lower GRA values
(Fig. F36B) suggest that parts of the core liner were
not filled. Below 1200 m (lithologic Subunit IIB), unlithified strata are rare and GRA results are consistently lower than MAD results by ~0.2 g/cm3 (Fig.
F36A, F36C), which can be attributed to the ~10%
smaller diameter of the RCB cores relative to the APC
core diameter used for calibration.
A comparison of bulk density generated by the different laboratory methods and expanded to include
downhole logging results in Hole U1352B is shown
in Figure F35. For this comparison, only results from
100 to 500 m are considered because downhole density logs were acquired only in this interval. Very low
densities measured by the downhole tool are spurious and may reflect hole enlargement generated by
drilling (see “Downhole logging”). Beginning at
~280 m, density log values that exceed MAD values
may be the result of the downhole tool sensing the
more indurated beds not sampled via APC and XCB
coring.
Magnetic susceptibility
Magnetic susceptibility (MSL and MSP) was measured at 5 cm intervals (measurement time = 2 s). Because of the excellent core quality and time-consuming RCB coring at greater depths, the sampling
resolution for MSL data was increased to 2.5 cm below 1361.4 m (starting with Section 317-U1352C87R-1) in order to provide a more detailed magnetic
susceptibility record of the lowermost cores from
Hole U1352C.
Raw MSL values are 3.7–39.0 instrument units for
Hole U1352A, 0–97.2 instrument units for Hole
U1352B, and 0–28.1 instrument units for Hole
U1352C. The data were filtered using a Gaussian
low-pass filter (30 passes; Fig. F34). The overall less
noisy loop sensor data, lower amplitudes, and a shift
to lower values in Hole U1352C compared to Hole
U1352B may have resulted from the mainly cemented sediment and the smaller diameter of RCB
cores relative to the core liner.
Magnetic susceptibility was measured at 5 cm intervals on the Section Half Multisensor Logger (SHMSL)
using the MSP (Fig. F34). These measurements were
31
Expedition 317 Scientists
made on all sections unless drilling disruption or surface disruption precluded the collection of meaningful results. MSP measurements made on section
halves from Holes U1352A and U1352B are noisier
than MSL data; however, the key trends are still resolvable after 30-pass Gaussian filtering of the data,
and these trends correlate well with MSL data. The
noise level in the MSP data increases markedly at
~520 m in Hole U1352B. A faulty MS2F sensor was
suspected to cause this deterioration in data quality,
and the unit was replaced at Core 317-U1352C-4R.
The replacement MS2F sensor measured slightly
higher instrument unit values than the older unit,
and data quality improved using this new sensor.
In general, higher magnetic susceptibility values
from both instruments correspond to darker sediments, and lower values correspond to lighter sediments (see “Spectrophotometry and colorimetry”
and Fig. F37). In the uppermost 275 m, both MSP
and MSL data display a cyclicity, with peaks and lows
(Fig. F34) that are broadly similar to the magnetic
susceptibility record from Site 1119 (Shipboard Scientific Party, 1999b) for the last glacial cycle. For example, three peaks between 50 and 70 m in Hole
U1352B (Fig. F38) correlate with similar peaks from
Holes 1119B and 1119C (between 36 and 46 m) that
were interpreted as representing marine isotope
Stage (MIS) 5 (Shipboard Scientific Party, 1999b). Between 555 and 630 m, magnetic susceptibility, NGR,
and GRA density show a conspicuous negative peak
followed by a corresponding positive peak between
630 and 655 m (Fig. F34). Below ~1500 m, magnetic
susceptibility and NGR signals decrease overall,
whereas color reflectance increases, likely reflecting
increasing carbonate content (see “Spectrophotometry and colorimetry”).
Below the Marshall Paraconformity at ~1852.6 m
(see “Lithostratigraphy”), magnetic susceptibility
measured with both the MSP and MSL becomes negative, indicating the predominance of diamagnetic
limestone.
Natural gamma radiation
NGR was measured on all core sections at 10 cm intervals down to 1333 m (Section 317-U1352C-82X1). The values measured range from near zero to >60
counts per second (cps), with higher values associated with muddy lithologies and lower values associated with sands (Fig. F34).
A sustained downhole decrease in NGR from ~55 to
~40 cps occurs between the seafloor and ~1400 m,
consistent with decreasing radiogenic terrigenous
material and increasing carbonate content. This decrease is punctuated by a major NGR perturbation
Proc. IODP | Volume 317
Site U1352
(with matching displacements in the magnetic susceptibility records) between 555 and 655 m, which
comprises a negative excursion to <20 cps, followed
by a positive excursion to >50 cps. No obvious lithologic explanation exists for this feature.
A further rapid decrease in NGR to ~15 cps occurs
just above the Marshall Paraconformity. Below the
unconformity, in the Amuri Limestone facies, NGR
reaches its lowest levels of 3–5 cps, indicating very
low terrigenous content.
A marked NGR cyclicity is apparent in the uppermost 275 m (Fig. F38), similar to that observed at
nearby Site 1119. Postcruise analysis will likely permit the development of a timescale for this interval
by comparing it with the Site 1119 age model, which
Carter and Gammon (2004) constructed using an inverted NGR record that apparently forms a close
New Zealand proxy for global MIS stages to at least
2.4 Ma. Assigning a provisional age of 127 ka to the
inferred MIS 5e peak at ~63 m at Site U1352 gives an
average sedimentation rate of ~50 cm/k.y. for the last
glacial–interglacial cycle, compared with 37 cm/k.y.
at Site U1351 and 20 cm/k.y. at Site 1119 over the
same interval.
P-wave velocities
P-wave velocity measurements were recorded with
different coverage on sections from Holes U1352A,
U1352B, and U1352C using the PWL, PWC, and
PWB (Figs. F39, F40). The differences in coverage depend on core condition. Poor results were caused by
signal attenuation and sediment cracking resulting
from high gas content in APC and XCB cores and
sediment disturbance related to XCB coring. PWL
measurements on whole-round sections from Holes
U1352A and U1352B at 4 cm intervals yielded poor
results below ~20.0 and 14.0 m, respectively (Fig.
F39A–F39B). PWB P-wave velocities were measured
only on the uppermost 20 m (y-axis) and 10 m (zaxis) of core, respectively, from Hole U1352B (Fig.
F39C). Excellent PWC results were obtained from
mainly cemented sediments in RCB cores (Fig.
F40A). PWC measurements were conducted on the
uppermost 20 m of core from Hole U1352B (Fig.
F39C) and continuously on nearly every section half
from Hole U1352C (Fig. F40A).
P-wave velocities (PWL, PWB, and PWC) from Holes
U1352A and U1352B range from ~1400 to ~1800
m/s, whereas P-wave velocities (PWC) from Hole
U1352C range from ~1600 to ~5900 m/s. The significantly higher velocities in XCB cores are probably
caused by good conductive contact with the mainly
cemented material. All velocity measurements from
Holes U1352A and U1352B show an abrupt change
32
Expedition 317 Scientists
from lower to higher values between 1.5 and 2 m. A
similar change was observed in the uppermost ~2.0 m
of Hole U1351B. Scattered values in the lowermost
part of the PWL curves are attributed to decreased
core quality (Fig. F39). Whole-round and discrete Pwave velocity measurements from Hole U1352B are
positively correlated. PWC and PWB velocity measurements abruptly decrease between 12 and 14 m,
approximately the same depth at which the PWL
measurements terminate. At greater depth, PWC values from Hole U1352C are inversely correlated with
porosity (Fig. F40C), showing a slight increase below
1255 m (average = ~2500–3500 m/s) and a large increase between 1500 and 1670 m (average = ~3500–
4000 m/s) (Fig. F40B). Between 1670 and ~1795 m,
velocity is continuously high (average = ~3900 m/s).
Below ~1795 m, P-wave velocity increases again to
~5900 m/s. A slight increase in P-wave velocity was
observed below the Marshall Paraconformity at
~1852.6 m (see “Lithostratigraphy”). The unexpectedly high velocities below 1255 m require a revision
of the present traveltime/depth interpretation of
seismic records in this area.
Spectrophotometry and colorimetry
Spectrophotometric measurements and associated
colorimetric calculations were made on section
halves at 5 cm intervals at the same positions used
for MSP measurements. Color data were recorded as
L*, a*, and b* variations. L* values are relatively constant in Holes U1352A and U1352B. Where Holes
U1352B and U1352C overlap between 600 and 800
m, L* values are slightly but noticeably lower in Hole
U1352C (Fig. F37). A clear increase in L* values occurs at ~1576 m. This increase can be attributed to a
transition to paler, more carbonate-rich lithologies
over this interval (see “Lithostratigraphy”).
Notably, the degree of scatter in L* data also increases from Hole U1352B to Hole U1352C, and a
further, more pronounced increase in scatter occurs
downhole in Hole U1352C below ~1600 m. This
change in L* between Holes U1352B and U1352C is
due to the change from homogeneous, unlithified
sediment in Hole U1352B to the more heterogeneous, lithified rock that characterizes Hole U1352C.
The additional increase in L* variance downhole in
Hole U1352C at ~1600 m is partly due to an increase
in reflectivity difference (i.e., contrast) between the
dominant pale marlstone and carbonate lithologies
and the minor interbedded dark lithologies (e.g.,
glauconitic muddy sandstone layers; see “Lithostratigraphy”). Furthermore, the change to lithified rock in Hole U1352C is also accompanied by an
increase in the occurrence of fractures within the
split sections. Spectrophotometer measurements of
Proc. IODP | Volume 317
Site U1352
these fractures resulted in anomalously low L* values.
The a* and b* values are less scattered than L* values,
and trends are discernible (Fig. F37). b* values in
Holes U1352B and U1352C increase over the uppermost ~1000 m before falling slightly downhole to
the bottom of Hole U1352C. Several more abrupt
changes also occur, and these are readily discernible
in the filtered b* data. The position of abrupt
changes in b* values matches the position of similarly abrupt shifts in a*. However, a* values do not
show the same long-term trends as b*, and, as demonstrated at Site U1351, the a* record is not as variable as the b* record. In general, a* is negatively correlated with b*, and changes in these two parameters
reflect changes from grayer (low b* and high a*) to
more greenish lithologies (high b* and low a*).
a* and b* values for Hole U1352A do not correlate
well with those for Hole U1352B between ~18 and
~54 m (i.e., the bottom of Hole U1352A). This difference may reflect a problem with instrument calibration in Hole U1352A for this depth interval because
the minimum and maximum values of b* and a* in
Hole U1352A are anomalous compared to values
from the rest of the site. In addition, as observed in
L* data, a* and b* data from Holes U1352B and
U1352C where the two holes overlap have different
absolute values. In particular, b* values at the top of
Hole U1352C (i.e., between ~574 and 670 m) are
anomalously high compared to b* values from the
same depth interval in Hole U1352B (Fig. F37). As
inferred for L* data, these differences reflect the fact
that different coring techniques preferentially recover different lithologies that may have different
color characteristics.
Moisture and density
Throughout Holes U1352B and U1352C, MAD measurements were made on approximately one sample
per core for low-recovery cores and on 3–5 samples
per core for complete cores with 6–8 sections. Samples were coordinated with thermal conductivity
measurements, discrete P-wave (PWC) measurements, chemical analyses, and smear slide samples.
Whenever possible, both soft-sediment samples and
indurated sediment samples were taken (using syringe or minicorer, respectively). When soft sediments were sampled, the standard Method C analysis was employed. When the minicorer was used and
the sample formed a measurable cylinder, the
Method D calculation was used in addition to the
Method C calculation (see “Physical properties” in
the “Methods” chapter).
The first three cemented samples were tested to evaluate the need to soak samples in seawater for 24 h
33
Expedition 317 Scientists
before obtaining reliable wet sample measurements.
These test samples were weighed immediately,
soaked in seawater for 24 h, and then weighed again.
In all cases, sample weight after soaking was greater
than that before soaking. However, the difference
was never >0.55% of sample weight. Because this is
well below other measurement errors, subsequent
rock samples were not soaked in seawater for 24 h
before their wet mass was determined.
MAD porosity results were calculated using Methods
C and D (Fig. F41) with a correlation coefficient of
0.98. Method D, in which wet volume is measured
and wet mass is calculated, is best used when pore
water may have been lost during sampling, because
it tends to slightly overestimate porosity (Fig. F41A).
Although occasional chips caused cylinder volume
to be overestimated, more often the top and base or
sides of the cylinder were not parallel. In either case,
the calipers tend to measure the largest diameter or
height, resulting in a tendency to overestimate wet
volume.
The porosity trend for sediments that were hard
enough to be sampled with the minicorer is readily
distinguished from that of sediments that were soft
enough to be sampled using the syringe method
(Figs. F41B, F42) between ~600–1750 m. Indurated
sediments within that interval have a relatively constant porosity downhole, whereas the softer sediments follow a decreasing porosity trend downcore
until their trend merges with that of the indurated
samples at ~1750 m.
Overall, density and porosity trends are clear in
Holes U1352B and U1352C (Fig. F42), and the data
are consistent over the interval from 575 to 821 m
where the two holes overlap. A few semi-indurated
samples from Hole U1352B show cementation beginning at ~500 m, whereas softer sediments tend to
more gradually decrease in porosity and void ratio
and increase in bulk density. Grain density slightly
decreases with depth and shows no relationship with
cementation, which supports the interpretation that
the lithologic makeup of these rocks is consistent
throughout.
Downhole logging of Hole U1352B was successful
between ~100 and 500 m (Fig. F43). An overall decrease in porosity with depth is overshadowed by
very large excursions that are indicated by downhole
logging tools and missing from MAD porosity estimates. Despite a relatively wide hole near the top,
porosity values derived from density logs are consistent with those obtained by MAD analyses, whereas
neutron porosity log values follow similar, though
consistently higher, trends (Fig. F43B). Slight offsets
between MAD results and downhole log measure-
Proc. IODP | Volume 317
Site U1352
ments may be related to differences in depth measurement between CSF-A and wireline log matched
depth below seafloor (WMSF). Deeper in the hole
(Fig. F43C), very high downhole log porosity values
(70%–100%) occur where the hole diameter is ≥20
inches, suggesting that the sediments surrounding
the hole may have been washed out at that level,
generating unrealistically high porosities. In general,
where this is not the case, porosities from downhole
logging are near or in some cases lower than those
obtained from MAD analyses. Again, this could reflect the presence of indurated horizons that were
not sampled by the XCB system, despite the apparent near-100% recovery at these depths. The very
high porosity excursion seen between ~240 and 300
m in the downhole logs (Fig. F43A) is likely a result
of drilling disturbance in the surrounding sediments.
Sediment strength
Sediment strength measurements were conducted on
working section halves from Hole U1352B using automated vane shear (AVS) and fall cone penetrometer (FCP) testing systems (Fig. F44). A comparison of
both measurement methods is shown in the crossplot in Figure F44D. Shear strength indicates that
sediments range from very soft (0–20 kN/m2) to very
stiff (150–300 kN/m2). Vane shear and fall cone shear
strength correlate well in very soft and soft sediments, but AVS values are about three times lower in
firm to very stiff sediments (standard deviation =
24.5 kN/m2) than FCP values (standard deviation =
81.8 kN/m2). These findings suggest that the applicability of vane shear in firm to very stiff sediments is
limited and that sediment strength from vane shear
tests is underestimated. Overall, vane shear and fall
cone strength data from Hole U1352B are positively
correlated (Fig. F44). Between 0 and ~295 m, shear
strength increases continuously, indicating a change
from very soft to firm sediments. The increase and
range of values seen in vane shear tests in the uppermost ~170 m of Hole U1352B are comparable to
those from Hole 1119B (Fig. F44C). A pronounced
cyclicity like that seen in holes at Site U1351 was not
observed. The generally lower sediment strength below ~295 m in cores from Hole U1352B coincides
with the change from APC to XCB drilling at 297 m.
Below ~295 m, sediment cracking resulting from
high gas content and sediment disturbance associated with XCB coring often prevented optimal insertion of the vane blade. Repeated downcore increases
in sediment strength were also observed within longer cores of uniform lithology by both AVS and FCP
tests. These downhole increases were interpreted as
an effect of secondary sediment compaction caused
34
Expedition 317 Scientists
by drilling stress. Similar drilling effects were observed in magnetic susceptibility data. We caution,
therefore, that interpretation of sediment strength
using raw data from XCB cores from Hole U1352B
requires careful analysis of both the drilling technique used and also of the in situ sedimentology.
Geochemistry and microbiology
Organic geochemistry
Shipboard organic geochemical studies of cores from
Site U1352 included monitoring hydrocarbon gases,
carbonate carbon, total organic carbon (TOC), total
sulfur (TS), and total nitrogen (TN) and characterizing organic matter by pyrolysis assay. The procedures
used in these studies are summarized in “Geochemistry and microbiology” in the “Methods” chapter.
All depths in this section are reported in CSF-A.
Volatile gases
All cores recovered in sufficient quantity at Site
U1352 were monitored for gaseous hydrocarbons using the headspace (HS) gas technique, and, where
possible, core gas voids were analyzed using the vacuum syringe (VAC) technique (Tables T18, T19, T20;
Figs. F45, F46).
Hole U1352A was sampled at relatively high frequency (every other section or at ~1.5–3.5 m intervals) with the headspace gas technique to estimate
the depth of the sulfate–methane transition (SMT)
and the dissolved methane versus depth gradient.
Sediment gas was below detection levels in the uppermost five samples collected (1.5–8.7 m), with the
first trace appearance of methane (~3 µM) occurring
at 11.7 m. Comparison with dissolved sulfate measurements (see “Inorganic geochemistry”) indicates
that the SMT is between 15.2 and 16.6 m.
Headspace methane contents in Hole U1352A are
shown in Table T18 as dissolved methane in sediment pore space (millimolar) and headspace gas
(parts per million by volume). At shallow depths beneath the sulfate reduction zone, methane concentrations are initially greater than saturation at surface conditions (~2 mM), having apparently
maintained some degree of supersaturation during
core retrieval and sampling. Estimated methane concentrations in sediment pore space were not calculated for Holes U1352B and U1352C because of obvious gas loss prior to sampling.
Detectable ethane is present in all cores from 18.2 m
and below. The composition of the gas, as expressed
by the C1/C2 ratio (Fig. F45C), shows the expected
gradual increase in relative ethane content with in-
Proc. IODP | Volume 317
Site U1352
creasing depth and temperature. The C1/C2 values of
the core void gases (Fig. F45C) generally parallel
those of headspace gas but are offset to higher values
because of greater retention of C1 during core retrieval and sampling. The ratio of C1/C2 in headspace
gas decreases by three orders of magnitude (from
>10,000 to 60) over the total depth interval (1927
m). The decrease in C1/C2 is due to a gradual increase
in ethane content (Fig. F45A); below ~650 m, this
decrease is also due to a decrease in methane and total gas content associated with a decrease in core porosity and fluid content. However, C1/C2 reaches
very low values (<10) from 1385 to 1393 m, with
headspace methane contents of <50 ppmv. This
depth also corresponds to a major discontinuity
identified between 1394 and 1410 m, with an apparent hiatus of >5 m.y. (see “Lithostratigraphy” and
“Biostratigraphy”). In addition, the deepest interstitial water sample (collected at 1385 m) contains 3
mM of sulfate, somewhat more than is usually observed for cores contaminated with seawater. The
very low methane and C1/C2 values of these cores
must represent either an interval in which all methane was lost or an interval in which sulfate was never
depleted and methane was never generated. In a few
cores below this depth, gas returns to normal, with
methane contents of 3,000–22,000 ppmv and most
C1/C2 ratios of 35–75 from 1512 to 1922 m. There is
no evidence for migrated thermogenic hydrocarbons
in the gas profiles.
Some headspace and all core void gas samples were
analyzed for the presence of C4+ hydrocarbons (Tables T19, T20; Fig. F46). Headspace samples were analyzed mainly below 650 m, when the character of
the cores made core void gas samples less frequent.
Butanes, pentanes, and occasionally hexanes are
present in most of the samples analyzed but are generally at low levels (1–100 ppmv) (Fig. F46A–F46C).
C4–C6 hydrocarbons become more abundant with
increasing burial depth (excluding the ~1300–1450
m interval that spans the hiatus), probably because
of the same low-temperature thermogenic alteration
of the indigenous organic matter that produces increased ethane and propane. A distinct preference
for branched C4 and C5 isomers is present in shallow
headspace samples, and this tendency diminishes
somewhat with increasing depth to ~1750 m (Fig.
F46F). The C4 hydrocarbons in the deeper parts of
the core have subequal amounts of normal and
branched isomers (e.g., below 1700 m), but C5 hydrocarbons continue to be dominated by i-C5, and
only the branched C6 hydrocarbons and no n-C6
could be detected in this depth range (Table T19; Fig.
F46F). A predominance of branched alkanes was also
35
Expedition 317 Scientists
observed in some core void gas samples from Site
U1351 (at 811 and 961 m) and may be a result of organic matter being derived mainly from land plants.
Core void gas samples have less of an iso-predominance (60%–80%) than headspace gases from similar
depth intervals and have no depth trend.
The plot of the ratio of n-butane to the sum of nbutane and i-butane (n-C4/[n-C4 + i-C4]) follows a
wormlike trend with increasing depth. This trend
has remarkable detail that may be related to lithostratigraphy and more precisely to formation porosity and permeability. Below 1400 m, where gas content is very low, this ratio first increases and then
decreases below 1500 m from a high of ~60% n-C4 to
~15% at 1600 m (Fig. F46F). With increasing depth,
this ratio increases again to ~45% at 1700 m, below
which it remains roughly constant. The n-C5/(n-C5 +
i-C5) ratio shows a similar but less distinct trend,
with a low between 1600 and 1700 m (0%–9%) and a
high of ~20% n-C5 in the deepest sample. Explanations for these trends remain uncertain, but they
most likely reflect the mixing of thermogenic equilibrium mixtures of normal- and iso-isomers (e.g., below 1800 m) with variably retained original hydrocarbons dominated by iso-isomers. Gases below 1650
m that are richer in C4 and C5 hydrocarbons (Fig.
F46A–F46B) form a cluster at the lowest part of the
n-C4/(n-C4 + i-C4) ratio trend, implying that original
hydrocarbons were retained in this zone and were
not substantially diluted by more recently generated
thermogenic gases. In contrast, the zone immediately underlying the discontinuity mentioned previously (below ~1450 m) is strongly depleted in original hydrocarbons, and gases in these sediments are
diluted by small amounts of more recently generated
thermogenic gas.
Table T20 and Figure F46D show ethene (C2=) contents measured in some void gas samples. Ethene
was also detected in some headspace gases (Table
T18). Ethene is common in near-surface sediments,
where it is thought to be a product of biological activity (Claypool and Kvenvolden, 1983). However,
short-chain unsaturated hydrocarbons are unstable
under the reducing conditions of deeper sediments,
so the presence of ethene and other alkenes in some
of the deeper (250–690 m; especially 422–467 m)
cores is somewhat unusual. Especially noteworthy is
the composition of the gas sample from Section 317U1352B-55X-1 (467 m), in which the amounts of
some of the alkenes are greater than the saturated
homologs (Fig. F47). Ethene and propene were quantified, but the peaks attributed to butenes and
pentenes could be only tentatively identified. These
samples are from depth intervals where XCB cores
that jammed in the core catcher gave off a strong
Proc. IODP | Volume 317
Site U1352
odor (422–467 m) and from the interval just above,
where cores became continually jammed in the core
barrel, requiring a switch from XCB to RCB coring.
Alkenes observed in some void gas and headspace
gas samples, especially the high quantities at 467 m,
may be the product of high-temperature alteration
of organic matter due to frictional heating during
drilling, similar to that observed in previous ODP
cores from Site 682 (Shipboard Scientific Party,
1988).
CO2 concentrations in headspace and void gas samples are given in Tables T19 and T20 and illustrated
in Figure F46E. Atmospheric air contains 387 ppm
CO2, but the majority of measured values exceed this
significantly; furthermore, some samples contain
less CO2 than is found in air. Values obtained from
core void gas samples show no clear difference from
those of headspace samples. Therefore, these values
likely reflect the true composition of CO2 in the subsurface, which was formed by a mixture of microbial,
diagenetic, and early thermogenic processes.
Carbon and elemental analyses
Inorganic carbon (IC), total carbon (TC), total organic carbon by difference (TOCDIFF), TN, and TS
were analyzed in 323 samples from 0 to 1924 m at
Site U1352 (Table T21; Fig. F48). Carbonate fluctuates between 0.14 and 96 wt%, with lower averages
of ~5 wt% in the uppermost 500 m and scattered
high-carbonate (>20 wt%) samples throughout the
cored sediments below ~600–700 m (Fig. F48A). The
deepest samples (1700–1927 m) are characterized by
very high carbonate contents (up to 96 wt%). Four
samples (317-U1352C-129R-2, 104 cm; 130R-4, 103
cm; 131R-2, 96 cm; and 147R-6, 71 cm; Table T21)
are carbonate-calcareous mudstone/marlstone pairs
from closely interbedded strata. TC values cluster at
~0.5–1 wt%, with frequent scatter as high as 9 wt%
in the uppermost 700 m. Below this depth, average
values increase slightly and fluctuate between 1 and
7 wt% from 700 to 1600 m before finally reaching 12
wt% in the deepest part of the hole below 1600 m
(Fig. F48B).
TOCDIFF fluctuates between 0.1 and 1.5 wt% but is
mostly <0.5 wt% and averages 0.4% (Fig. F48C).
TOCDIFF is systematically lower than the organic carbon determination given by the source rock analyzer
(TOCSRA) (Fig. F48D). A cross-plot of TC from the elemental analyzer and TOCSRA plus IC from the coulometer shows a good correlation and indicates that
the elemental analyzer gives consistently lower values by 0.3–0.5 wt% (Fig. F49), similar to Site U1351.
TN and TS contents are scattered in the range of
0.02–0.07 wt% for TN and 0–0.5 wt% for TS in the
36
Expedition 317 Scientists
uppermost 500 m. They slightly decrease downhole
to averages of ~0.01–0.02 wt% and 0.02 wt% for TN
and TS, respectively (Fig. F50C–F50D). No TS content was analyzed below 1526 m because of an onboard shortage of elemental analyzer reactors, which
had to be replaced every 100 samples because the vanadium pentoxide catalyst used for sulfur determination deteriorated the activated copper package in
the reactor. TOCDIFF/TN ratios generally range from 5
to 50 and tend to increase with depth to ~1600 m
before slightly decreasing near the bottom of the
hole. Occasional values as high as 304 were observed, reflecting very low TN contents, mostly in
carbonates (Fig. F50A). TOCDIFF/TS ratios are mostly
low, ranging from 0.1 to 10, with scattered individual samples as high as 80 (Fig. F50B). TOCDIFF/TN and
TOCDIFF/TS ratios are much higher at Site U1352 than
at Site U1351.
Organic matter pyrolysis
Most of the samples used for carbon-nitrogen-sulfur
analysis were also characterized by source rock analyzer (SRA) pyrolysis (Table T22; Figs. F51, F52). S1
and S2 slightly increase with depth from 0 to 400 m,
with ranges of ~0.0–0.3 and 0.1–1.4 mg/g, respectively. Values range widely at these shallow depths,
especially for S1. Below 600 m, values cluster more
tightly near 0.04 mg/g for S1 and 0.1–0.5 mg/g for S2
(Fig. F51A–F51B), and average values decrease
slightly downhole. S2 values at 1500–1600 m are occasionally as high as 2.4 mg/g, representing sediment layers with higher organic carbon contents. S3
has no trend and a range of 0.1–0.9 mg/g, with occasional values as high as 1.2 mg/g (Fig. F51C). Pyrolysis carbon necessarily mirrors S1 and S2 (Fig. F51D).
The hydrogen index (HI) generally ranges from 10 to
100 mg/g C, and values increase with depth, reaching a maximum at ~1000 m before decreasing
slightly toward the bottom of the hole. Values as
high as 133 mg/g C were observed at ~1500–1600 m
(Fig. F52A). The oxygen index is scattered between
10 and 100 mg/g C, and values slightly decrease with
depth (Fig. F52B). Tmax values (Fig. F52C) range from
~370° to 440°C, with more scatter in the uppermost
500 m of the sediment column. Values increase
downhole from ~400°C at the seafloor (with much
scatter) to 430°C at the bottom of the hole. The
cloud of data in the uppermost 500 m tightens into a
more consistent trend in deeper sediments. The production index decreases in the uppermost 700 m
from an average of 30% at the seafloor to ~10% at
~1000 m and then stays constant to ~1500 m (Fig.
F52D). Below 1500 m, the production index increases significantly in some samples to 20%–30%.
Proc. IODP | Volume 317
Site U1352
A modified van Krevelen diagram (Fig. F53) indicates
that organic matter is of slightly better quality at Site
U1352 than at Site U1351, with most samples clustering near the Type III line. The highest HI (133
mg/g C) was found in a thin Miocene calcareous
mudstone from 1575 m that was interbedded with
carbonates.
Preliminary interpretation of organic matter
Variations in organic matter composition are subtle
and largely correlate with lithologic Units I (0–710
m) and II (710–1853 m). Unit I is interbedded mud
and sand with scattered high carbonate intervals,
Unit II is a marlstone with mud and sand, and Unit
III is nannofossil-rich limestone. In Unit I, samples
with high carbonate contents tend to have higher
organic carbon contents and higher S1 and S2 pyrolysis yields, whereas background mud/sand sediments
have relatively low organic carbon contents (TOCDIFF
= 0.4–0.5 wt%). Unit II marls generally have low organic carbon contents (TOCDIFF = 0.5–0.6 wt%), and
organic carbon does not correlate with carbonate
content except at the bottom of the unit below 1700
m, where an inverse correlation is apparent (Fig.
F48A, F48C). Unit III limestones contain the highest
amounts of carbonate and the lowest TN at Site
U1352 (Figs. F48A, F50C). Unit I probably contains
organic matter that is less altered diagenetically and
that volatilizes more readily during pyrolysis. Organic matter in Units II and III is more diagenetically
stabilized as protokerogen, as shown in the lower
part of Unit II by the elimination of some scatter in
the pyrolysis response, especially in S3 and Tmax (Figs.
F51C, F52C). Tmax values, especially those below
~1200 m, have less scatter and are shifted to higher
temperatures, consistent with the progressive elimination of the more reactive protokerogen components of organic matter (Fig. F52C).
SRA data from Site U1352 can be compared to existing source rock quality and thermal maturity data
from the Canterbury Basin (Newman et al., 2000;
Sykes and Funnel, 2002; Sykes, 2004; Sykes and Johansen, 2009). Source rock quality at Site U1352 is
rather low, with most HI values <100 mg/g (Fig.
F52A), so the organic matter is largely interpreted to
be land plant or degraded marine in origin. This contrasts with deeper Pukeiwitahi Formation coals (late
Campanian–early Maastrichtian) in offshore petroleum exploration wells that have considerably
higher HI values (Endeavour-1: 2094–2353 m, mean
HI = 210; Galleon-1: 2822–2885 m, mean HI = 250)
(Sykes, 2004). It is possible that the reported HI values at Site U1352 are underestimates because it was
demonstrated that the TOCSRA values from the SRA
37
Expedition 317 Scientists
are high relative to TOCDIFF calculated by the difference method (see “Geochemistry and microbiology” in the “Methods” chapter).
The thermal maturity gradient defined by Tmax variation at Site U1352 ranges from ~380° to 400°C in the
shallowest samples to an average of ~430°C at ~1900
m (Fig. F52C). This trend is quite steep and suggests
a rather high geothermal gradient. The deepest values are within the early oil-generative window, according to published Tmax data that show that the
onset of oil generation in the Canterbury Basin occurs at Tmax = 425°C (0.65% vitrinite reflectance; coal
band rank Sr = 10) (Sykes and Funnel, 2002; Sykes
and Johansen, 2009). Note that this conclusion is inconsistent with the interpreted bottom-hole temperature of ~60°C at Site U1352 that was calculated using a variable heat flow determined by thermal
conductivity (see “Heat flow”). However, the geochemical results are more consistent with a bottomhole temperature of ~100°C that is achieved if a constant geothermal gradient is assumed. Supporting evidence for an early oil-generative window thermal
maturity at the base of Site U1352 is that the production index increases above 20% below 1700 m (Fig.
F52D), showing greater free hydrocarbons that are
perhaps generated by thermal processes. However,
similar Tmax values are generally only reached deeper
in the Canterbury Basin (>2100–2800 m in Endeavour-1 and Galleon-1; Sykes, 2004; Sykes and Johansen, 2009). One possible explanation is the known
loss of section at erosional unconformities at Site
U1352 (see “Lithostratigraphy”). A second possible
explanation is that heat flow at Site U1352 is high
because of a deeper igneous intrusion, although
there is no evidence on seismic profiles at Site U1352
for igneous intrusions at depths that we believe
could have had a thermal effect in the drilled hole.
Newman et al. (2000) have shown evidence for
anomalously high thermal maturities in the Canterbury Basin at Clipper-1 below 4000 m, which is inferred to have been the result of a thermal intrusion
at depth.
SRA data obtained on board ship, especially HI and
Tmax, will need to be confirmed and calibrated using
a second instrument before initial interpretations
about source rock quality and thermal maturity at
Site U1352 can be confirmed. These analyses are
scheduled for early postcruise research.
Inorganic geochemistry
Interstitial water
A total of 112 interstitial water samples (Tables T23,
T24, T25) were collected and analyzed at Site U1352.
Thirty-one samples were taken from Hole U1352A,
Proc. IODP | Volume 317
Site U1352
which was dedicated mainly to whole-round sampling for geochemistry and microbiology. Three samples were taken from Core 317-U1352B-6H, two samples were taken from each of Cores 7H through 10H,
and one sample was taken per core (where recovery
was sufficient) to Core 35H (294 m). Samples were
taken from approximately every other core for Cores
317-U1352B-37X through 90X (783 m). Cores 317U1352C-2R, 3R, 6R, 11R, 14R, and 18R were spot
sampled from 576 to 780 m, and every second or
third core was sampled to Core 40R (989 m). From
Cores 317-U1352C-45R through 58R (1034–1164 m),
sampling was irregular and some whole-round samples as long as 30 cm failed to yield any water for
analysis. Thereafter, only a few interstitial water samples were selected, based on appearance and apparent degree of cementation. The amount of interstitial water extracted from whole-round samples
decreases with depth from ~2 to 5 mL/cm in the uppermost 700 m to <1 mL/cm below 900 m (Table
T23; Fig. F54). The deepest whole-round sample successfully sampled for interstitial water was from 1386
m and yielded 0.5 mL/cm.
Salinity, chloride, and pH
Salinities in samples near the seafloor are slightly
lower than normal seawater at 3.3, rapidly decline to
3.0 at 28 m, and remain relatively constant at 2.9–
3.1 in the rest of the section analyzed (to 1400 m)
(Fig. F55B, F55D). Chloride parallels salinity measurements, with the shallowest samples having
slightly lower concentrations than seawater (540
mM) and most other samples having relatively constant but scattered chloride concentrations of 520–
550 mM (Fig. F55A, F55C). The isolated deepest sample has a chloride concentration of 453 mM. Measured pH values seem to vary with the dominant diagenetic process—decreasing during sulfate reduction,
increasing during methanogenesis, and decreasing
again, possibly because of authigenic carbonate precipitation (Fig. F56C).
Alkalinity, sulfate, ammonium, phosphate, and
dissolved silica
Alkalinity increases relatively slowly from 2.8 mM at
the seafloor to 9 mM at 10.1 m and then increases
rapidly to a maximum of 24.2 mM at 16.6 m (Fig.
F57A, F57C). Alkalinity then decreases to ~15 mM at
100 m and remains relatively constant to ~400 m.
From 400 to 600 m, alkalinity declines steadily to
~2.3 and then remains in the range of 1.4–3.0 mM to
the base of sampling. The sulfate decline is almost
exactly the inverse of the alkalinity increase, with
sulfate declining slowly over the 0–10 m interval and
rapidly from 10 to 24 m (Fig. F57B, F57D). Below
38
Expedition 317 Scientists
this depth, sulfate remains essentially at zero, except
in cores contaminated with small amounts of the
seawater used as a drilling fluid.
Ammonium is zero to ~7 m and then increases gradually to a shallow maximum of ~2.3 mM between 40
and 50 m (Fig. F58C). After decreasing to 1.3 mM between 50 and 80 m, ammonium increases again to a
maximum of ~7 mM at 470 m before decreasing
steadily to ~2 mM in the deepest samples analyzed at
1000 m. Phosphate is low at 1–4 µM from the seafloor to 8.5 m. It then increases rapidly and spikes at
92 µM at 14.8 m (Fig. F58A). After dropping back to
41 µM at 16.6 m, phosphate varies within a range of
23–59 µM to 55 m before dropping to ~7 µM at 60 m
and declining steadily to essentially zero by 400 m.
Dissolved silica is present at 272 µM at 1.1 m, increases to 645 µM at 27.5 m, remains relatively constant at a range of 380–660 µM to 200 m, and then
increases to a maximum of 1066 µM at 480 m (Fig.
F58B). Silica declines to 215 µM at 653 m and then
increases again to 831 µM at 699 m before dropping
back to 258 µM in the deepest samples analyzed at
765 m.
Calcium, magnesium, and strontium
Calcium and magnesium both decrease during sulfate reduction and then continue to decrease to minimum values (1.4 mM for Ca2+ and 7 mM for Mg2+)
from 300 to 400 m (Fig. F56D–F56E). Below 400 m,
both major cations increase to ~20–23 mM at ~600
m, below which magnesium remains relatively constant to ~1200 m. Calcium continues to increase to
just above 30 mM in the deepest samples at >1100
m. The ratio of magnesium to calcium increases
from 5 to >9 in surface sediments as deep as 18 m. It
then decreases to ~1.6 at 500 m before slowly declining to 0.5 at 1386 m (Fig. F56F). Strontium is initially at seawater values of ~0.1 mM and then increases slightly to 0.3 mM at 400 m before quickly
increasing downhole to 2 mM at ~800 m. Strontium
then decreases to ~1.7 mM at maximum depth (1386
m) (Fig. F56A). The Sr/Ca ratio increases steadily
from 0.01 to a maximum of 0.15 at 480 m and then
decreases again to 0.06 in the deepest sample (Fig.
F56B).
Sodium, potassium, lithium, barium, silicon,
boron, iron, and manganese
Sodium increases from near-seawater values of 466
mM to 515 mM at 286 m, decreases to ~440 mM at
600 m, drops to ~410 mM at 1184 m, and then decreases to 348 mM at 1386 m (Fig. F59C). Potassium
decreases during sulfate reduction and then increases
during the initial stages of methanogenesis, reaching
Proc. IODP | Volume 317
Site U1352
a maximum of 11 mM at 144 m. Potassium then declines to ~6 mM at 550 m, declines further to ~3 mM
at 600 m, spikes back up to 5.7 mM at 709 m, and finally declines to ~2 mM in the deepest samples below 900 m (Fig. F59B).
Lithium increases steadily from 25 µM at the sediment/water interface to ~50 µM at 450 m. Lithium
then increases more rapidly to a maximum of 166
µM at 700 m before declining to ~100 µM at 800–
900 m (Fig. F59A). Below 900 m, lithium increases
again to ~130 µM between 1000 and 1200 m before
dropping to 76 µM in the deepest sample. Barium increases from ~2 to ~7 µM at the base of the SMT before gradually increasing to 19 µM at 500 m (Fig.
F59D). The profile becomes more scattered below
500 m, with concentrations ranging from 7 to 30
µM. A pronounced barium maximum of 84 µM is evident at 1091 m.
Silicon has no obvious trend in the uppermost 300 m
of sediment, and values are scattered between 300
and 600 µM (Fig. F58D). Below 300 m, silicon increases to a maximum of ~900 µM at 524 m, drops
significantly to ~300 µM at 600 m, increases again to
1013 µM at 700 m, and finally averages ~400 µM below 800 m.
The boron profile shows a remarkable increase from
seawater values of ~0.4 mM to a maximum of 5.4
mM at 1113 m before a slight decrease in the subsequent samples and a sharp drop to 1.5 mM in the
deepest sample (Fig. F60A).
Manganese ranges from 3 to 9 µM, with a maximum
value of 13 µM between 0 and 50 m. Below this
depth, manganese declines to 2–4 µM at 100–300 m
and then increases downhole to average values of 5–
15 µM (Fig. F60C). Iron shows a similar trend, increasing rapidly from ~10 µM to a maximum of 34
µM at 26 m, decreasing again to 4 µM at ~300 m,
and finally increasing to 20–36 µM between 500 and
1386 m (Fig. F60B).
Preliminary interpretation of diagenesis
Interstitial water geochemistry in the uppermost 20
m at Site U1352 is dominated by the two main microbially mediated diagenetic processes, sulfate reduction and methanogenesis (Fig. F61). A zone of
very gradual sulfate depletion and alkalinity increase
occurs in the 0–8.5 m depth interval and represents
either very slow organic matter oxidation or a zone
of intense bioturbation. The very low phosphate and
the absence of ammonium in this interval are more
consistent with bioturbation or other physical mixing of seawater than with organic matter oxidation.
From 8.5 to 16.6 m, sulfate declines rapidly from
25.1 to 0 mM, whereas alkalinity (dominantly bicar-
39
Expedition 317 Scientists
bonate ions) increases to a maximum of 24.2 mM.
Over this same 8.5–16.6 m depth interval, calcium
and magnesium decline by a combined 15 mM, presumably because of authigenic carbonate precipitation. The amount of carbon oxidized during sulfate
reduction is equivalent to the alkalinity increase plus
the major cation decrease, or ~38 mM, which is 1.5
times the amount of sulfate removed (Fig. F62). This
ratio of carbon oxidized to sulfate removed implies
that half of the sulfate reduction at Site U1352 is the
result of organic matter oxidation and the other half
is the result of methane oxidation. At this 1.5 proportion of carbon added to sulfate removed, twothirds of the carbon added would be from the oxidation of organic matter and one-third would be from
the oxidation of methane (see equations in Fig. F62).
These interpretations based on interstitial water
chemistry could be confirmed by postcruise analysis
of stable carbon isotope ratios of dissolved inorganic
carbon and diagenetic carbonates. The moderate
amount of phosphate in the sulfate reduction zone,
especially in the sample at 14.8 m (92 µM), is consistent with some oxidation of marine organic matter
because methane oxidation would generate no phosphate.
After an initial drop of 2 mM beneath the sulfate reduction zone, alkalinity remains relatively constant
to ~50 m, indicating that organic matter oxidation is
replenishing bicarbonate as rapidly as it is removed
by methane generation and carbonate precipitation.
Below 50 m, alkalinity drops in stages to ~14 mM
over the 100–350 m depth interval and then drops
steadily to ~3–5 mM at 500–600 m, below which it
ranges from 1.4 to 3.0 mM in the rest of the sampled
interval. This gradual decline in alkalinity probably
represents the final stages of biological activity resulting in the oxidation of organic matter and the reduction of dissolved CO2 to produce methane. The
alkalinity decrease and the major calcium increase
occur over the same interval (350–600 m), which
also corresponds to a decline in the degree of preservation of calcareous microfossils (see “Biostratigraphy”).
Calcium at Site U1352 generally increases to 400 m
and then reaches a steady state at ~800 m. This is
common for pore fluids from carbonate-dominated
sections with little influence from diffusive flux
below (McDuff and Gieskes, 1976) and is consistent
with high carbonate throughout the cored sediments
below ~600–700 m (Fig. F48A). Below 400 m,
strontium and magnesium also have profiles similar
to that of calcium. Upward diffusive flow from
carbonate-dominated sections may cause the abrupt
gradual increases in divalent cation concentrations
from 400 to 800 m. Strontium reaches values as high
Proc. IODP | Volume 317
Site U1352
as >20 times seawater. The dissolution of Sr-rich
aragonite in the sediments is one way of explaining
the increasing flux of strontium to pore fluids.
Particles of barite are one of the main carriers of barium in sediments (Dehairs et al., 1980). Barium initially shows a clear rise below the depth of the SMT.
The removal of SO42– increases the solubility of barite, and dissolved Ba2+ concentrations increase from
2 to 9 µM (Fig. F59D). The scattered profiles of silica
and silicon below 400 m are possibly associated with
the scatter of barium concentrations and may reflect
diagenetic dissolution/transition and/or changes in
paleoproductivity (Paytan et al., 1996).
Boron steadily increases in the deepest samples,
where values are as high as 13 times that of seawater
(Fig. F60A). The maximum boron value of 5.4 mM is
very similar to that at Site U1351, although at Site
U1351 the highest value occurs much shallower (at
300 m). The boron increase is possibly related to the
diagenetic opal-A/opal-CT transition and microbial
degradation of organic matter.
Biogenic opal is assumed to be a major source of
lithium in sediments. The rapid increase and
decrease of lithium concentrations from 450 to 800
m (Fig. F59A) is within a zone of highly variable
silica and silicon concentrations (Fig. F58B, F58D). A
fraction of lithium enrichment also may be associated with lithium in sediments and clay minerals.
Lithium is easily removed from clay interlayer
exchange sites because of its high hydration energy,
which may account for the observed steady increase.
Microbiology
The principal shipboard microbiological objectives
at Site U1352 included testing samples for contamination using PFT and a particulate tracer and starting
enrichment cultures for different types of metabolisms.
Sample collection
At this site, 107 whole-round samples were collected
for microbiological investigations (51 for microbial
characterization, 51 for intact polar lipid analysis,
and 5 for incubation tests).
Contamination tracer tests
Contamination tests were continuously conducted
using water-soluble tracers (PFT) or bacteria-sized
particles (fluorescent microspheres) in order to confirm the suitability of sediment samples for microbiological research. The chemical and particle tracer
techniques used are described in “Geochemistry
and microbiology” in the “Methods” chapter.
40
Expedition 317 Scientists
Site U1352
Water-soluble tracer
Cultivations
At Site U1352, PFT was continuously delivered at its
maximum solubility (2 mg/mL in seawater) into
cores from which microbiology whole-round samples were taken. To maximize the detection of seawater contamination from drilling, 5 cm3 sediment
samples were taken. No results could be achieved because of analytical problems. In particular, gas chromatograph traces had small peaks (other than PFT)
of similar retention time to that of PFT, even after
cleaning or changing the column. Consequently, we
decided to preserve the samples for onshore analysis
by injecting 2 mL of an autoclaved 3% NaCl solution
into the vials containing the sediment. The samples
were then immediately frozen upside-down at –80°C
until they could be shipped on dry ice.
In Hole U1352C, 11 whole-round samples were collected below the deepest microbiology sample analyzed to date (1626 mbsf; Roussel et al., 2008). These
samples are characterized by the presence of lithified
layers rich in carbonate (up to 70%; see “Lithostratigraphy”) alternating with dark glauconitic
layers. When a glauconitic layer was present in the
10 cm long microbial characterization samples, it
was sampled separately in order to determine any
differences in microbial diversity. Three samples
(Sections 317-U1352C-137R-3, 137R-3 [glauconite],
and 148R-3) (Table T26) were inoculated on several
enrichment media, as described in “Geochemistry
and microbiology” in the “Methods” chapter (see
Table T16 in the “Methods” chapter). All onboard
enrichments were incubated near the in situ temperature (70°C) (the temperature at the bottom of the
hole was estimated to be 60°C; see “Heat flow”) under 5% H2, 5% CO2, and 90% N2 (biogas) headspace.
Particulate tracer
Fluorescent microspheres were used as a particulate
tracer on all cores from which whole-round samples
were subsequently taken. A 1 cm3 sample of sediment was diluted 10× by a 3% NaCl/3% formalin solution for microsphere detection. Microspheres were
detected in 49 of the 51 outer sediment samples (Table T26), indicating a heterogeneous distribution of
microspheres along the core liner. In the inner part
of the core, 40 sediment samples did not contain any
microspheres, indicating that no contamination by
micron-sized particles took place. In 11 sediment
samples (Table T26), between 2 × 102 and 2 × 105 microspheres/cm3 were counted, indicating that potential contamination from drilling fluid may have occurred in the inner part of the core during the
drilling process. The bead-delivery method used at
this site is the one described by Smith et al. (2000).
In samples where microspheres were not detected
(Sections 317-U1352A-15H-4 and 317-U1352C-87R4), prokaryotic molecular diversity was compared to
determine if contamination occurred. No difference
was found in the deployment of microspheres between APC and XCB coring, but, on average, fewer
microspheres were observed at the periphery of the
core when RCB coring was used (Fig. F63). This can
be explained by the fact that beads can potentially
become diluted in the drilling fluid. Therefore, we
suggest using 40 mL instead of 20 mL microsphere
bags for cores retrieved with the XCB or RCB systems.
Total cell counts
Sediment samples of 1 cm3 were taken from all
whole-round samples for microbial characterization
and stored at 4°C in a 3% NaCl/3% formalin solution
for onshore prokaryotic cell counting.
Proc. IODP | Volume 317
Heat flow
Geothermal gradient
Temperature measurements were conducted using
the APCT-3 during APC coring in Holes U1352A and
U1352B and the Sediment Temperature tool during
XCB coring in Hole U1352B. Six temperature measurements were taken in total (Fig. F64; Table T27),
and the geothermal gradient was successfully obtained from four of these (Cores 317-U1352B-10H,
15H, 20H, and 38X) within the depth interval of
93.7–313.2 m CSF-A. Unless otherwise noted, all
depths in this section are reported in m CSF-A. The
other two measurements in Cores 317-U1352A-4H
and 317-U1352B-6H were not used because the conductive cooling time after sediment penetration was
too short (<300 s) to generate reliable fitting curves
(Fig. F64A), which could be the result of tool movement caused by ship heave. The fitting line to
temperature versus depth data was derived from the
four good results (Fig. F64B):
T(z) = 0.0460 × z + 8.2325
(R2 = 0.9991),
where T(z) is in situ temperature at depth z (m CSFA). The estimated geothermal gradient is therefore
46.0°C/km. Note that this geothermal gradient was
established for the depth interval above ~310 m,
which consists of soft sediments, and might
significantly decrease with depth in accordance with
a rapid increase in thermal conductivity, particularly
for the interval below 575 m, where rock first occurs
(see “Thermal conductivity,” below). For reference,
41
Expedition 317 Scientists
the geothermal gradients at nearby ODP Sites 1120,
1124, and 1125 are 57.4°, 52.1°, and 64.9°C/km,
respectively (Carter, McCave, Richter, Carter, et al.,
1999). These gradients were established in soft
sediments above ~130 m.
Thermal conductivity
Thermal conductivity was measured preferentially in
available whole-round core sections from Holes
U1352A and U1352B using the full-space needle
probe method and in section halves from Hole
U1352C using the puck probe method. The puck
probe method was employed because the degree of
induration of sediments increases with depth in
Hole U1352C. Three full-space needle probe
measurements were conducted in indurated
sediment drilled with the RCB system: two in rather
soft sediments in Cores 317-U1352C-6R and 18R and
one in rock to compare with puck probe
measurements
in
Core
317-U1352C-76R.
Measurement frequency was usually once per section
with one measuring cycle at each point for the fullspace needle probe method and once per core with
five measuring cycles at each point for the puck
probe method. This included 32 points in Hole
U1352A (0.2–42.1 m), 443 points in Hole U1352B
(0.7–821.7 m), and 155 points in Hole U1352C
(575.1–1920.6 m) (Table T28). The middle of each
section was chosen as the measurement point unless
a void or crack was observed (see “Heat flow” in the
“Methods” chapter). Few lithologic variations occur
in each section at Site U1352, so this sampling
procedure was appropriate. Probes V10701 and
V10819 were used, and heating power was kept to ~3
and 2 W for full-space needle probe and puck probe
methods, respectively.
Thermal conductivity data were screened when (1)
contact between the probe and sediment was poor,
(2) thermal conductivity values were close to that of
water (0.6 W/[m·K]) because of sediment dilution
during coring, and (3) measurements were taken in
caved-in layers such as shell hash. In most cases, the
first two criteria were controlling parameters for
deciding the quality of measurements. Good results
were obtained at 13, 214, and 149 points in Holes
U1352A, U1352B, and U1352C, respectively,
covering depths of 8–42, 1–793, and 575–1921 m,
respectively (Table T28). The ratio of reliable
measurements to total measurements is larger for
Hole U1352C, where five measuring cycles were run,
than for Holes U1352A and U1352B, where only one
measuring cycle was run. Therefore, we recommend
increasing the number of measuring cycles at each
Proc. IODP | Volume 317
Site U1352
point from one to at least three, even at the expense
of decreasing measuring frequency from once per
section to once per core.
Thermal conductivity measurements at Site U1352
range from 0.849 to 3.440 W/(m·K): 0.849–1.696
W/(m·K) (average = 1.305 W/[m·K]) for sediments in
Holes U1352A and U1352B in the depth interval of
1–793 m and 1.572–3.440 W/(m·K) (average = 2.360
W/[m·K]) for rocks in Hole U1352C in the depth interval of 575–1921 m (Table T28). For reference, the
two lowest values in Hole U1352C were measured
with the full-space needle probe in sediments in
Cores 317-U1352C-6R and 18R. For Core 317U1352C-76R we compared thermal conductivity
measured by the puck probe with that measured by
the full-space needle probe within a drilled hole
filled with thermal compound. Results were similar
(within 0.6%), indicating that the difference between methods is negligible. Thermal conductivity
for rocks is ~1.8 times greater than that for sediments. In the uppermost 130 m, thermal conductivity values are higher at Site U1352 than in the same
interval at nearby Site 1119 (Shipboard Scientific
Party, 1999b). The high conductivities at Site U1352
may be due to high concentrations of quartz (6.5–
12.5 W/[m·K]) in fine-grained sediment, including
the clay-sized fraction (see “Lithostratigraphy”)
and/or carbonate cementation (0.5–4.4 W/[m·K]).
Thermal conductivity versus depth data for Holes
U1352A and U1352B are consistent (Fig. F65A). In
the overlapped depth interval between Holes
U1352B and U1352C (575–793 m), a gap exists
between trends of values that increase with depth.
This gap can be explained by sampling bias with
respect to the lithologies measured in each hole: soft
sediment was preferentially recovered in Holes
U1352A and U1352B and hard rock was
preferentially recovered in Hole U1352C. Although
marlstone first occurs at 575 m (Core 317-U1352C2H), the rock is still porous and firm at ~900 m,
which is manifested in the larger amount of scatter
above 900 m compared to below. Three linear trends
can be recognized at other depth intervals, including
a downhole decrease from 0 to 90 m and increasing
trends from 90 to 661 m and 826 to 1921 m. It is
unclear why values in the uppermost 90 m interval
decrease with depth, because bulk density and
porosity are relatively constant in the same interval.
However, thermal conductivity, in general, correlates
negatively with the porosity profile (see “Physical
properties”), particularly for hard rock (Fig. F65B). A
good positive correlation with bulk density obtained
from moisture and density, Method C (see “Physical
42
Expedition 317 Scientists
Site U1352
properties” in the “Methods” chapter), is shown in
Figure F65C. The linear fits between thermal conductivity and depth for sediments are
λ0–90(z) = 1.3484–0.0015 × z
(R2 = 0.0954)
λ90–661(z) = 1.1878 + 0.0003 × z
(R2 = 0.1742),
and
and the linear fit for hard rock is
λ826–1921(z) = 1.7819 + 0.4101 × z (R2 = 0.4101),
where z is depth (m CSF-A). Thermal conductivity in
the 661–826 m depth interval was calculated using
the harmonic mean for cores consisting of
alternating sediment–rock layers, based on two
regressions, λ90–661(z) and λ826–1921(z), and the ratio of
sediments/rocks in each core because (1) the
sediment portion of each core from Hole U1352C
steadily decreases below ~661 m (Core 317-U1352C6R) and almost disappears at 826 m (Core 317U1352C-23R) and (2) the measured thermal
conductivities in Holes U1352B and U1352C
represent either sediment or rock among alternating
layers of sediment or rock. The resulting linear fit for
sediment/rock mixture is
λ661–826(z) = –1.1457 + 0.004 × z
(R2
= 0.6376),
where z is depth (m CSF-A). In addition to this interval, Cores 317-U1352C-2R, 27R, 53R, and 77R contain alternating layers of sediment and rock, but
these were not taken into account in terms of trends.
Bullard plot
Because the thermal conductivity profile in the 94–
313 m depth interval, where the geothermal
gradient was established, is represented as a linear fit,
λ90–661(z), thermal resistance for the interval is derived as
Ω90–661(z) = [ln(1.1878 + 0.0003 × z) – ln(1.1878)]/
0.0003,
where z is depth (m CSF-A).
Following the Bullard approach and assuming
conductive heat flow, a linear fit of temperature
versus thermal resistance is expected (Fig. F66):
T90–661(z) = 7.8814 + 0.0578 × Ω(z) (R2 = 0.9985),
where z is depth (m CSF-A).
This yields a heat flow of 57.8 mW/m2 for the 94–
313 m depth interval, which can be applied to the
Proc. IODP | Volume 317
entire cored depth interval if steady state heat flow is
assumed. The resulting heat flow is comparable to
the regional heat flow distribution, which decreases
from 100–120 mW/m2 in the mountainous area to
the southwest to <60 mW/m2 on the west coast of
New Zealand (Reyes, 2007).
Temperature profile
The temperature profile at Site U1352 was predicted
using the estimated heat flow of 57.8 mW/m2 from
the 94–313 m depth interval and estimated thermal
conductivity trends under the assumption of steady
state (Fig. F67). The temperature profile based on
thermal conductivity shows a large inflection at
~575 m because of a rapid increase in thermal conductivity in the lithified material. At the bottom of
Hole U1352C (1927 m), the predicted temperature
based on this method is ~60°C, which is ~40°C lower
than that obtained by assuming a constant
geothermal gradient. However, SRA data suggest a
higher thermal maturity at the bottom of Hole
U1352C than would be consistent with 60°C (see
“Geochemistry and microbiology”).
Downhole logging
Operations
Two holes were logged at Site U1352. Hole U1352B
was logged after reaching refusal with the XCB coring system, and Hole U1352C was logged after coring deep enough to be able to log and fully characterize the Marshall Paraconformity.
Hole U1352B
After an extended period of low core recovery, slow
rate of penetration, and hot conditions at the bit,
APC/XCB coring operations were terminated, with
the last core on deck at ~1900 h on 4 December 2009
(all times are ship local time, UTC + 13 h). Logging
operations began by conditioning the hole immediately after the last core was recovered from a total
depth of 1185.5 m DRF (831 m DSF). Because of the
tool recovery operations in Hole U1351C (see
“Downhole logging” in the “Site U1351” chapter),
mud supplies on the ship were limited for the remaining sites; however, the drilling process and lithology indicated that this hole could be displaced
with seawater rather than logging gel. After the hole
was swept, circulated with 50 bbl of high-viscosity
mud, and displaced with seawater, the bit was raised
to the logging depth of 436.6 m DRF (82 m DSF). Rig
up of the triple combo tool string (natural gamma
ray, bulk density, electrical resistivity, and porosity)
began at 2315 h and was complete at 0025 h on
43
Expedition 317 Scientists
5 December, and the tool string was RIH at 0045 h.
While the tool string was being lowered at a speed of
2500–3000 ft/h, gamma ray and resistivity data were
recorded from the seafloor to 842 m wireline log
depth below rig floor (WRF), below which it proved
impossible to lower the tool string. After multiple
unsuccessful attempts to pass the blockage at ~842 m
WRF, we decided to log up from this depth, more
than 300 m shallower than the total hole depth. A
first repeat pass was completed at 0255 h (470 m
WRF), and the tool was run back down to 842 m
WRF for the full pass, which began at 0305 h at a
speed of 900 ft/h. At 0420 h (477 m WRF), the caliper was closed for reentry into the pipe, and the pass
was completed at 0500 h when the seafloor was
identified by a drop in the gamma ray log at 355.5 m
WRF. The tool string was back on deck at 0520 h and
rigged down at 0610 h.
Overall, the caliper of the density sonde showed a
very irregular borehole, with a hole diameter >20
inches over many intervals. We decided that, even if
the quality of FMS images was likely to be poor in
some intervals, there was no risk to the tools and the
deployment of the FMS-sonic tool string would provide worthwhile velocity and image data. The FMSsonic tool string was rigged up by 0700 h and RIH at
a speed of 2000 ft/h to record sonic velocities on the
way down. At 0800 h, a shallower obstruction was
met at 795 m WRF and could not be passed. At 0812
h, we decided to begin logging up with both the FMS
and sonic tools from the deepest depth reached (797
m WRF) at a speed of 1200 ft/h. The first pass was
completed at 0900 h at 480 m WRF, before the top of
the tool string reached the bottom of the pipe. The
FMS calipers were closed, and the tool was sent
down for a second pass. At 0920 h, the same obstruction was reached at 796 m WRF, and the second pass
started at a speed of 1500 ft/h. The FMS calipers were
closed after the second pass was completed at 1005 h
(480 m WRF), before the top of the tool string was
pulled into the pipe. Data acquisition was concluded
at 1020 h when the sonic log identified the bottom
of the drill string. The tool string was back at the surface at 1050 h and was rigged down completely at
1130 h. The initial logging plan included a VSP;
however, this was postponed until the next hole because of hole conditions, and the rig floor was
cleared to resume drilling operations at 1145 h.
Hole U1352C
The last core was recovered from Hole U1352C at
1740 h on 19 December from a total depth of 1927.5
m DSF. After sweeping the bottom of the hole with
50 bbl of high-viscosity sepiolite/attapulgite mud
and circulating two times the annular capacity of the
Proc. IODP | Volume 317
Site U1352
hole with seawater, the RCB bit was dropped at
~2300 h. In order to reserve mud supplies for the remaining sites and in consideration of the stable nature of the formation in the deepest section of the
hole, a 450 m interval was targeted between 900 and
450 m DSF to be displaced with logging mud. The
pipe was pulled to 900 m DSF, and 400 bbl of logging
mud was pumped into the borehole to stabilize this
interval, which previous logging and coring operations had shown to be unstable. Shortly thereafter,
the drill string indicated significant drag from the
formation, which required reconnection of the top
drive to restore rotation. The continued upward
progress of the pipe was slow between 880 and ~190
m DSF, suggesting a series of ledges rather than a full
hole collapse but also indicating that lowering the
logging tool could be hazardous. Considering the
potential benefits of logging this unique hole, we decided to run the triple combo tool string without radioactive sources.
At 1400 h on 20 December, the pipe was set at a logging depth of ~100 m DSF (354.6 m DRF), and rig-up
of the triple combo began at 1520 h. The modified
tool string consisted of the Hostile Environment
Natural Gamma Ray Sonde (HNGS), the Hostile Environment Litho-Density Sonde (HLDS; without
source, for the caliper only), the General Purpose Inclinometry Tool (GPIT), and the Dual Induction Tool
(DIT). The tool string was rigged up by 1600 h and
RIH at 1620 h at a speed of 4000 ft/h. A downlog was
started at 340 m WRF at a speed of 3500 ft/h, and the
seafloor and base of pipe were identified at 355 and
457 m WRF, respectively. At 1700 h, cable tension indicated signs of drag from the formation at ~550 m
WRF, similar to the shallowest trouble spot identified
by the driller as the drill string was pulled out of the
hole. After several attempts to pass the obstruction,
reaching a maximum depth of 562 m WRF, further
downhole progress was deemed hazardous. At 1720
h, the caliper was opened, recording a hole diameter
of 6 inches, significantly narrower than the 97⁄8 inch
bit size, and an uplog was started at a speed of 900 ft/
h. The caliper was closed at 490 m WRF before the
top of the tool string reached the bottom of the drill
string, and the uplog was completed at 1800 h when
the gamma ray log identified the seafloor at 355 m
WRF. The tool string was at the surface at 1825 h and
completely rigged down by 1851 h. The rig floor was
cleared to start the hole abandonment protocol at
1904 h.
Data quality
Figures F68, F69, and F70 show a summary of the
main logging data recorded in Hole U1352B. These
data were converted from original field records to
44
Expedition 317 Scientists
depth below seafloor and processed to match depths
between different logging runs. The resulting depth
scale is WMSF (see “Downhole logging” in the
“Methods” chapter).
The first indicators of the overall quality of the logs
are the size and shape of the borehole measured by
the calipers. Hole size measured by the HLDS caliper
during the triple combo run and by the FMS arms are
shown in Figures F68 and F69, respectively. Although both runs indicate an enlarged and irregular
hole, the readings of the two orthogonal FMS calipers (Fig. F69) suggest that the borehole cross section
is not circular and is probably elliptical below ~270 m
WSF. One caliper read close to 12 inches over most of
the lower half of the logged interval, whereas the
other caliper read close to ~14 inches, near the limit
of its range. The fact that the curves display variability over most of the hole suggests that both sets of
arms made some kind of contact with the formation,
possibly with only one pad in some places. The
larger HLDS caliper readings above 270 m WMSF indicate that the stronger and narrower single-arm caliper was likely pushing into the formation. Both sets
of calipers show that the hole diameter is almost uniformly larger than 16 inches (>46 cm) above 270 m
WMSF, suggesting that the tools were not able to
make good contact with the formation above this
depth. Porosity and density data, in particular,
should be used with caution, most clearly between
200 and 270 m WSF, where log data deviate significantly from the same measurements made on cores
(see below). Below this depth, the hole is apparently
less enlarged, but it remains very irregular, and many
anomalously low density readings below this depth
are likely indicative of multiple narrow washouts
that significantly affect the quality of the density
readings.
Logs recorded in Hole U1352C show similar trends
to those in Hole U1352B, reflecting the same response to lithology (Fig. F71). Lower values in
gamma ray and medium resistivity in Hole U1352C
can be explained by an extremely enlarged borehole,
which is supported by HLDS caliper readings of
>19.5 inches throughout the logged interval. Low
gamma ray and resistivity above 115 m WMSF and
between 185 and 190 m WMSF, in particular, may reflect large borehole washouts.
The quality of the logs can also be assessed by comparing log data with core measurements from the
same hole. Figure F68 shows a comparison of gamma
ray and density logs with NGR and GRA bulk density
track data and with MAD measurements made on
cores recovered from Hole U1352B (see “Physical
properties”). The gamma ray measurements generally agree well, even in the upper half of the hole,
Proc. IODP | Volume 317
Site U1352
but the density log is seriously affected by hole conditions below ~200 m WMSF, with highly variable
and anomalously low readings.
The high coherence in sonic waveforms indicated by
distinct red areas in the VP and VS tracks in Figure
F69 suggests that, despite the enlarged hole, the Dipole Sonic Imager (DSI) was able to generate strong
compressional and flexural waves and should provide reliable compressional and shear velocity values. However, all of the velocity profiles display a
high variability that is likely not representative of
formation properties and will require additional
postcruise processing to be fully characterized. In addition, VS values above ~150 m WMSF are systematically on the fastest edge of the high-coherence domain, suggesting a processing drift that should also
be corrected postcruise.
Porosity and density estimation
from the resistivity log
In order to provide a measure of porosity and density
from the logs despite the poor hole conditions, we
used Archie’s (1942) relationship to calculate porosity from the phasor deep induction log (IDPH),
which is the log least affected by borehole conditions (Schlumberger, 1989), and combined it with
MAD grain density data to derive a more reliable
density profile. Archie (1942) established an empirical relationship between porosity (φ), formation resistivity (R), and pore water resistivity (Rw) in sandy
formations:
φ = (aRw/R)1/m,
where m and a are two empirical parameters that are
often called cementation and tortuosity (or Archie)
coefficients, respectively. The resistivity of seawater
(Rw) was calculated as a function of temperature and
salinity, as described by Fofonoff (1985). Pore water
salinity was assumed to be 30 ppt (or 3%; see “Geochemistry and microbiology”), and temperature
was assumed to follow a linear gradient of 42°C/km,
as established by in situ temperature measurements
(see “Heat flow”). The most realistic value for the cementation coefficient is a = 1 because this gives a resistivity equal to formation water resistivity when
porosity is 100%. A value of m = 1.9 was chosen iteratively to provide the best baseline match with MAD
porosity data. Although Archie’s relationship was
originally defined for sand-rich formations, Jarrard et
al. (1989) showed that the effect of clay minerals is
moderate, and the relationship is commonly used to
estimate porosity in clay-rich formations with poor
borehole conditions (Collett, 1998; Jarrard et al.,
1989). The resulting porosity log is shown in Figure
45
Expedition 317 Scientists
F68, where it compares very well with MAD porosity
data. Using MAD grain density, we used this resistivity-derived porosity to calculate a new density curve,
which is in good agreement with core data and does
not display any of the anomalous variability of the
original density log.
Logging stratigraphy
The combined analysis of gamma ray, resistivity,
density, and velocity logs allows for the identification of logging units defined by characteristic trends.
Because of the uniformity of the sediments in the
logged interval (see “Lithostratigraphy”), these
units are mostly defined by subtle changes in trends
and correlations rather than by indications of significant changes in the formation. The downhole logs
were used to define two logging units.
Logging Unit 1 (82–250 m WMSF) is characterized by
relatively low amplitude variations in gamma ray, resistivity, and acoustic velocities. A distinct increasing-upward trend in gamma ray occurs at 250 m
WMSF, and a high gamma ray interval occurs between ~160 and 170 m WMSF, which corresponds to
an interval of homogeneous mud with clay beds
(Core 317-U1352B-19H). Gamma ray spectroscopy
shows a significant uranium contribution to the total radioactivity throughout this unit (Fig. F70). The
similarity between the total gamma ray and the potassium and thorium curves suggests that the increase in total gamma ray is related to variations in
mineralogy. Gamma ray decreases above 160 m
WMSF, which may reflect generally decreasing clay
content in this depth interval. The increasing-upward and then decreasing-upward pattern in gamma
ray is consistent with gamma ray logs from an equivalent depth at Site 1119 (Shipboard Scientific Party,
1999b). Resistivity decreases gradually with depth,
whereas velocity increases with depth. Caliper measurements that were consistently >19.5 inches (the
maximum reach of the HLDS caliper) indicate an enlarged borehole.
Logging Unit 2 (250–487 m WMSF) is defined by a
change to higher amplitude variations in gamma ray,
resistivity, and acoustic velocities. Gamma ray and
velocity increase with depth. Sharp peaks in uranium
(Fig. F70) are associated with high resistivity values
and can be correlated with green calcareous sandy
intervals observed in the core recovered in this unit
(see “Core-log correlation”). Another distinct feature of this unit is a 15 m thick fining-upward sequence between 435 and 450 m WMSF, characterized
by an upward increase in gamma ray that is mirrored
by an upward decrease in resistivity. The borehole diameter in this unit is smaller but highly irregular,
Proc. IODP | Volume 317
Site U1352
ranging from 6 to 19.5 inches, which may reflect the
presence of more cohesive marls in the formation.
Core-log correlation
Some of the most remarkable features in the logs recorded in Hole U1352B are the sharp peaks in uranium below ~270 m WMSF, all of which are associated with more modest peaks in resistivity and very
distinct peaks in velocity, indicating fine indurated
layers (Figs. F68, F69, F70). The higher resistivity of
these layers, although less pronounced than the
changes in velocity or uranium content, is enough to
make them appear as bright features in the FMS electrical images, making them easy to correlate with
core images. However, some of these peaks coincide
with intervals of incomplete recovery (e.g., Cores
317-U1352B-37X, 42X, and 49X), likely because of
the hardened, highly resistive layers, and the only
means of correlation are missing intervals or traces
in the core catcher (e.g., the calcareous nodule in
Section 317-U1352B-49X-CC).
Some of these layers are shown in Figure F72. The
sections of core that seem to match the high-uranium resistive layers are alternately described as dark
greenish gray calcareous sands (Core 317-U1352B32H), dark greenish gray sandy calcareous mud
(Core 317-U1352B-35H), olive calcareous sandy marl
(Core 317-U1352B-36H), olive calcareous muddy
sand and sandy mud (Core 317-U1352B-37X), or olive sandy mud with calcareous concretions (Core
317-U1352B-42X). These sections of core have in
common a calcareous sandy component and the occurrence of nodules or some level of cementation.
Log-seismic correlation
A depth–traveltime relationship can be determined
from the sonic logs and used to correlate features in the
logs, recorded in the depth domain, to features in the
seismic stratigraphy, recorded in the time domain. A
synthetic seismogram was constructed for the logged
interval in Hole U1352B (82–487 m WSF) using sonic
data and the density curve calculated from the resistivity log using Archie’s relationship. Several packages of
strong seismic reflections were reproduced in the synthetic seismogram shown in Figure F73, particularly
below ~250 m WMSF (0.77 s two-way traveltime). In
the uppermost 250 m of the borehole, sonic velocity
values measured in Hole U1352B are slightly lower
than, but in general agreement with, sonic velocities
recorded during Leg 181 at Site 1119, located in the
slope environment ~35 km from Site U1352 (Fig. F74)
(Shipboard Scientific Party, 1999b). Although the general trends in velocity between 250 and 390 m WMSF
are consistent, several discrete intervals (e.g., 250–270
46
Expedition 317 Scientists
and 300–320 m WMSF) have higher velocities at Site
1119. Below 390 m WMSF, sonic velocity at Site 1119
appears significantly higher than that in Hole U1352B.
The velocity logs recorded at these two sites are also
consistently slower than the velocity model used to determine the depth of the reflectors during the Expedition 317 transect (Fig. F74; see “Traveltime/depth
conversion” in the “Expedition 317 summary” chapter). This velocity model was derived from data recorded in Hole 1119C and in the Clipper-1 well and
predicted higher velocity values, which resulted in an
offset of ~40 m at the bottom of Hole U1352B between
the depths of the reflectors predicted by the velocity
model and those predicted by the Hole U1352B synthetic seismogram (Figs. F73, F74). Because the velocity measured on core samples recovered from Hole
U1352C suggests significantly higher velocity at depth
(see “Physical properties”), it is possible that the
time–depth relationships could have been reconciled at
depth if logging of the deep interval in Hole U1352C
had been successful.
Stratigraphic correlation
Hole U1352D, located ~20 m from Hole U1352B, was
cored as a dedicated APC hole to ~128 m and no
whole-round samples were taken. This hole was
cored specifically to provide an independent record
of Site U1352 that could be correlated with Holes
U1352A and U1352B. These three overlapping records provide an opportunity to construct a spliced
stratigraphic record against a common core composite depth below seafloor (CCSF) depth scale (see
“Stratigraphic correlation” in the “Methods” chapter). MSL and NGR data were used to facilitate correlation, and further appraisal of correlations was carried out using color and GRA bulk density records.
Because of whole-round sampling of Hole U1352A,
coupled with its relatively short length and an analysis of its correlation potential in the software Correlator, the stratigraphic record from this hole was not
required for correlation purposes. NGR and MSL data
from Holes U1352B and U1352D share a number of
key features, allowing correlative ties between the
holes to be made (Fig. F75). The depths of these correlative features (in CSF-A) often differ between the
two holes, requiring depth shifts of individual and/
or multiple cores of as much as 7.63 m. Cumulative
depth adjustments made to cores from Site U1352
are provided in Table T29. Unless they are erroneous,
these large adjustments strongly suggest that the
stratigraphic record of both holes is characterized by
localized differences in the sedimentation records of
the two localities. Depth adjustments of this size are
much larger than is typical for correlation of pelagic
Proc. IODP | Volume 317
Site U1352
sediments, reflecting the fact that the sedimentation
and/or postsedimentation history of the Site U1352
continental slope environment was extremely dynamic. Equally, however, the validity of the composite depth–adjusted records for each hole needs to be
treated with caution because of these unusually large
depth shifts. Additionally, correlation tie points are
generally ambiguous in one way or another, making
the correlation fundamentally uncertain. The lack of
unambiguous correlation ties between the two holes
may stem from a variety of sources, including genuine inconsistencies in the physical properties of contemporaneous strata, inaccuracies/imprecision in
NGR and MSL data, and drilling disturbances that affect the fidelity of the data (see “Stratigraphic correlation” in the “Methods” chapter).
The spliced record presented in Figure F75 was constructed using Hole U1352B as the basis of the record
because it was deemed to be the most complete. The
splicing in of parts of Hole U1352D was performed
only over limited intervals as close as possible to correlative ties. For the final spliced record, only four
relatively short intervals were spliced in from Hole
U1352D to cover either recovery gaps or suspected
intervals of relative incompleteness in Hole U1352B.
Correlation between Holes U1352B and U1352D in
the uppermost ~90 m was relatively straightforward
and required only minimum adjustments of core
depths in each hole. Two tie points in the uppermost
10 m suggest that ~4 m of strata may be missing in
Hole U1352D between two peaks in the NGR record;
the MSL data also support this interpretation. Because the standard shipboard protocol permits depth
adjustments only to entire cores, only one point per
core can be tied to a point in another core (see
“Stratigraphic correlation” in the “Methods” section). This means that the depth adjustment of Hole
U1352D at the ~7 m tie does not align the prominent low in NGR values at ~11 m in both holes because this feature is in the same core as the tie point
(at ~7 m). Between ~64 and ~67 m, the stepwise shift
to lower NGR values and the trough in MSL data suggest that Hole U1352B is missing ~2 m of strata relative to Hole U1352D. Edge effects and shell-hash
cave-in complicate the records farther downhole,
and care was therefore taken that these artifacts were
not used to aid in correlation.
A tie at ~102 m was made at the base of a prominent
trough in the NGR values of both records, but the
known edge effects associated with these measurements on the Whole-Round Multisensor Logger
means that this tie also takes into account the overall trend to lower values in the two data sets between
~67 and ~100 m. A smaller depth adjustment was
made in Hole U1352B at ~89 m to correlate to a
47
Expedition 317 Scientists
prominent peak in the NGR data of Hole U1352D
(Fig. F75).
Note that the depth shift in Hole U1352D that was
needed to create the tie made at ~102 m is potentially erroneous because it is particularly large (~8.3
m) and a good match between the MSL records of
the two holes could not be ascertained. In contrast,
depth shifts made to both Holes U1352B and
U1352D above 70 m resulted in a very good match
between the MSL data of both holes (particularly between 50 and 65 m). The lack of correlation between
MSL records below ~70 m potentially represents an
important caveat to the veracity of the correlations
below this depth. The use of additional data such as
color and GRA bulk density below ~102 m to aid in
correlation between the two holes did not provide
any additional support for the ties made using NGR,
and the potential correlation of these data sets is
similarly equivocal. GRA bulk density measurements
within individual APC cores are sometimes inconsistent because of the progressive compaction of sediment downcore and the sensitivity of this physical
property to coring disturbance; hence, these data
were omitted from the correlation exercise.
An additional tie was made at ~122 m between NGR
peaks present in both Holes U1352B and U1352D
(Fig. F75). Another peak in both data sets at ~118 m
also ties as a consequence of this depth adjustment.
A prominent low in the NGR record of Hole U1352D
at ~135 m (which is not considered to represent an
edge effect) can be spliced into the Hole U1352B record to cover a recovery gap in Hole U1352B based
on the correlation of a peak in NGR at ~131 m. As
with the depth shift made to create the tie at 102 m,
these correlations provide little correlation in the
MSL records, and thus these ties should be treated
with caution.
Overall, the correlation between Holes U1352B and
U1352D emphasizes the difficulty of correlating continental slope sediments within the constraints of
the correlation approach used (see “Stratigraphic
correlation” in the “Methods” chapter) and the fact
that both records are potentially stratigraphically incomplete relative to each other.
References
Archie, G.E., 1942. The electrical resistivity log as an aid in
determining some reservoir characteristics. J. Pet. Technol., 5:1–8.
Browne, G.H., 1987. In situ and intrusive sandstone in
Amuri facies limestone at Te Kaukau Point, southeast
Wairarapa, New Zealand. N. Z. J. Geol. Geophys.,
30(4):363–374.
Proc. IODP | Volume 317
Site U1352
Browne, G.H., and Naish, T.R., 2003. Facies development
and sequence architecture of a late Quaternary fluvialmarine transition, Canterbury Plains and shelf, New
Zealand: implications for forced regressive deposits. Sediment. Geol., 158(1–2):57–86. doi:10.1016/S00370738(02)00258-0
Carter, R.M., 1985. The mid-Oligocene Marshall Paraconformity, New Zealand: coincidence with global eustatic
sea-level fall or rise? J. Geol., 93(3):359–371.
doi:10.1086/628957
Carter, R.M., and Gammon, P., 2004. New Zealand maritime glaciation: millennial-scale southern climate
change since 3.9 Ma. Science, 304(5677):1659–1662.
doi:10.1126/science.1093726
Carter, R.M., McCave, I.N., and Carter, L., 2004. Leg 181
synthesis: fronts, flows, drifts, volcanoes, and the evolution of the southwestern gateway to the Pacific Ocean,
eastern New Zealand. In Richter, C. (Ed.), Proc. ODP, Sci.
Results, 181: College Station, TX (Ocean Drilling Program), 1–111. doi:10.2973/odp.proc.sr.181.210.2004
Carter, R.M., McCave, I.N., Richter, C., Carter, L., et al.,
1999. Proc. ODP, Init. Repts., 181: College Station, TX
(Ocean Drilling Program). doi:10.2973/
odp.proc.ir.181.2000
Clark, P.U., Archer, D., Pollard, D., Blum, J.D., Rial, J.A.,
Brovkin, V., Mix, A.C., Pisias, N.G., and Roy, M., 2006.
The middle Pleistocene transition: characteristics,
mechanisms, and implications for long-term changes in
atmospheric pCO2. Quat. Sci. Rev., 25(23–24):3150–
3184. doi:10.1016/j.quascirev.2006.07.008
Claypool, G.E., and Kvenvolden, K.A., 1983. Methane and
other hydrocarbon gases in marine sediment. Annu. Rev.
Earth Planet. Sci., 11(1):299–327. doi:10.1146/
annurev.ea.11.050183.001503
Collett, T.S., 1998. Well log evaluation of gas hydrate saturations. Trans. SPWLA Annu. Logging Symp., 39:MM.
Coombs, D.S., Cas, R.A., Kawachi, Y., Landis, C.A.,
McDonough, W.F., and Reay, A., 1986. Cenozoic volcanism in north, east, and central Otago. In Smith, I.E.M.
(Ed.), Late Cenozoic Volcanism in New Zealand. Bull.—
R. Soc. N. Z., 23:278–312.
Cooper, R.A., 2004. The New Zealand Geological Timescale.
Inst. Geol. Nucl. Sci. Monogr., 22.
Dehairs, F., Chesselet, R., and Jedwab, J., 1980. Discrete
suspended particles of barite and the barium cycle in
the open ocean. Earth Planet. Sci. Lett., 49(2):528–550.
doi:10.1016/0012-821X(80)90094-1
Eade, J.V., 1967. New Zealand recent foraminifera of the
families Islandiellidae and Cassidulinidae. N. Z. J. Mar.
Freshwater Res., 1(4):421–454. doi:10.1080/
00288330.1967.9515217
Edwards, A.R., Hornibrook, N. de B., Lewis, D.W., Smale,
D., and van der Lingen, G.J., 1979. The depositional
depth of the Oxford Chalk—Comment. Newsl. Geol. Soc.
N. Z., 47:40–43.
Field, B.D., and Browne, G.H., 1989. New Zealand Geological Survey Basin Studies (Vol. 2): Cretaceous and Cenozoic
Sedimentary Basins and Geological Evolution of Canterbury
48
Expedition 317 Scientists
Region, South Island, New Zealand: Christchurch, New
Zealand (DSIR).
Fofonoff, N.P., 1985. Physical properties of seawater: a new
salinity scale and equation of state for seawater. J. Geophys. Res., [Oceans], 90(C2):3332–3342. doi:10.1029/
JC090iC02p03332
Fulthorpe, C.S., and Carter, R.M., 1991. Continental-shelf
progradation by sediment-drift accretion. Geol. Soc. Am.
Bull., 103(2):300–309. doi:10.1130/
0016-7606(1991)103<0300:CSPBSD>2.3.CO;2
Fulthorpe, C.S., Carter, R.M., Miller, K.G., and Wilson, J.,
1996. Marshall paraconformity: a mid-Oligocene record
of inception of the Antarctic Circumpolar Current and
coeval glacio-eustatic lowstand? Mar. Pet. Geol.,
13(1):61–77. doi:10.1016/0264-8172(95)00033-X
Gartner, S., 1977. Calcareous nannofossil biostratigraphy
and revised zonation of the Pleistocene. Mar. Micropaleontol., 2:1–25. doi:10.1016/0377-8398(77)90002-0
Heiden, K.A., and Holmes, M.A., 1998. Grain-size distribution and significance of clay and clay-sized minerals in
Eocene to Holocene sediments from Sites 918 and 919
in the Irminger Basin. In Saunders, A.D., Larsen, H.C.,
and Wise, S.W., Jr. (Eds.), Proc. ODP, Sci. Results, 152:
College Station, TX (Ocean Drilling Program), 39–49.
doi:10.2973/odp.proc.sr.152.248.1998
Jarrard, R.D., Dadey, K.A., and Busch, W.H., 1989. Velocity
and density of sediments of Eirik Ridge, Labrador Sea:
control by porosity and mineralogy. In Srivastava, S.P.,
Arthur, M.A., Clement, B., et al., Proc. ODP, Sci. Results,
105: College Station, TX (Ocean Drilling Program), 811–
835. doi:10.2973/odp.proc.sr.105.146.1989
Jolly, R.H., and Lonergan, L., 2002. Mechanisms and controls on the formation of sand intrusions. J. Geol. Soc.
(London, U. K.), 159(5):605–617. doi:10.1144/0016764902-025
Kameo, K., and Bralower, T.J., 2000. Neogene calcareous
nannofossil biostratigraphy of Sites 998, 999, and 1000,
Caribbean Sea. In Leckie, R.M., Sigurdsson, H., Acton,
G.D., and Draper, G. (Eds.), Proc. ODP, Sci. Results, 165:
College Station, TX (Ocean Drilling Program), 3–17.
doi:10.2973/odp.proc.sr.165.012.2000
Lewis, D.W., 1973. Polyphase limestone dikes in the
Oamaru region, New Zealand. J. Sediment. Petrol.,
43(4):1031–1045. doi:10.1306/74D728E2-2B21-11D78648000102C1865D
Lind, I.L., Janecek, T.R., Krissek, L.A., Prentice, M.L., and
Stax, R., 1993. Color bands in Ontong Java Plateau carbonate oozes and chalks. In Berger, W.H., Kroenke, L.W.,
Mayer, L.A., et al., Proc. ODP, Sci. Results, 130: College
Station, TX (Ocean Drilling Program), 453–470.
doi:10.2973/odp.proc.sr.130.007.1993
Lu, H., and Fulthorpe, C.S., 2004. Controls on sequence
stratigraphy of a middle Miocene–Holocene, currentswept, passive margin: offshore Canterbury Basin, New
Zealand. Geol. Soc. Am. Bull., 116(11):1345–1366.
doi:10.1130/B2525401.1
Lu, H., Fulthorpe, C.S., and Mann, P., 2003. Three-dimensional architecture of shelf-building sediment drifts in
the offshore Canterbury Basin, New Zealand. Mar. Geol.,
193(1–2):19–47. doi:10.1016/S0025-3227(02)00612-6
Proc. IODP | Volume 317
Site U1352
McDougall, I., and Coombs, D.S., 1973. Potassium-argon
ages for the Dunedin volcano and outlying volcanics.
N. Z. J. Geol. Geophys., 16(2):179–188.
McDuff, R.E., and Gieskes, J.M., 1976. Calcium and magnesium profiles in DSDP interstitial waters: diffusion or
reaction? Earth Planet. Sci. Lett., 33(1):1–10.
doi:10.1016/0012-821X(76)90151-5
McGonigal, K., and Di Stefano, A., 2002. Calcareous nannofossil biostratigraphy of the Eocene–Oligocene Transition, ODP Sites 1123 and 1124. In Richter, C. (Ed.),
Proc. ODP, Sci. Results, 181: College Station, TX (Ocean
Drilling Program), 1–22. doi:10.2973/
odp.proc.sr.181.207.2002
Newman, J., Eckersley, K.M., Francis, D.A., and Moore,
N.A., 2000. Application of vitrinite-inertinite reflectance and fluorescence (VIRF) to maturity assessment in
the East Coast and Canterbury Basins of New Zealand.
N. Z. Pet. Conf. Proc., 314–333. http://www.crownminerals.govt.nz/cms/pdf-library/petroleum-conferences-1/2000-conference-proceedings/newman-375kb-pdf
Paytan, A., Kastner, M., and Chavez, F.P., 1996. Glacial to
interglacial fluctuations in productivity in the equatorial Pacific as indicated by marine barite. Science,
274(5291):1355–1357. doi:10.1126/science.274.5291.1355
Peuraniemi, V., Aario, R., and Pulkkinen, P., 1997. Mineralogy and geochemistry of the clay fraction of till in
northern Finland. Sediment. Geol., 111(1–4)313–327.
doi:10.1016/S0037-0738(97)00023-7
Reyes, A.G., 2007. Abandoned oil and gas wells: a reconnaissance study of an unconventional geothermal
resource. GNS Sci. Rep., 2007/23.
Richter, C., Acton, G., Endris, C., and Radsted, M., 2007.
Handbook for shipboard paleomagnetists. ODP Tech.
Note, 34. doi:10.2973/odp.tn.34.2007
Roussel, E.G., Cambon Bonavita, M.-A., Querellou, J.,
Cragg, B.A., Webster, G., Prieur, D., and Parkes, J.R.,
2008. Extending the sub-sea-floor biosphere. Science,
320(5879):1046. doi:10.1126/science.1154545
Schlumberger, 1989. Log Interpretation Principles/Applications: Houston (Schlumberger Educ. Serv.), SMP–7017.
Sewell, R.J., 1988. Late Miocene volcanic stratigraphy of
central Banks Peninsula, Canterbury, New Zealand. N.
Z. J. Geol. Geophys., 31(1):41–64.
Sewell, R.J., Weaver, S.D., and Reay, M.B., 1992. Geology of
Banks Peninsula, 1:100,000. Inst. Geol. Nucl. Sci. Map, 3.
Shipboard Scientific Party, 1988. Site 682. In Suess, E., von
Huene, R., et al., Proc. ODP, Init. Repts., 112: College Station, TX (Ocean Drilling Program), 363–435.
doi:10.2973/odp.proc.ir.112.113.1988
Shipboard Scientific Party, 1999a. Leg 181 summary:
southwest Pacific paleoceanography. In Carter, R.M.,
McCave, I.N., Richter, C., Carter, L., et al., Proc. ODP,
Init. Repts., 181: College Station, TX (Ocean Drilling Program), 1–80. doi:10.2973/odp.proc.ir.181.101.2000
Shipboard Scientific Party, 1999b. Site 1119: drift accretion
on Canterbury Slope. In Carter, R.M., McCave, I.N.,
Richter, C., Carter, L., et al., Proc. ODP, Init. Repts., 181:
49
Expedition 317 Scientists
College Station, TX (Ocean Drilling Program), 1–112.
doi:10.2973/odp.proc.ir.181.103.2000
Smith, D.C., Spivack, A.J., Fisk, M.R., Haveman, S.A., Staudigel, H., and the Leg 185 Shipboard Scientific Party,
2000. Methods for quantifying potential microbial contamination during deep ocean coring. ODP Tech. Note,
28. doi:10.2973/odp.tn.28.2000
Suggate, R.P., 1990. Late Pliocene and Quaternary glaciations of New Zealand. Quat. Sci. Rev., 9(2–3):175–197.
doi:10.1016/0277-3791(90)90017-5
Sykes, R., 2004. Peat biomass and early diagenetic controls
on the paraffinic oil potential of humic coals, Canterbury Basin, New Zealand. Pet. Geosci., 10(4):283–303.
doi:10.1144/1354-079302-568
Sykes, R., and Funnel, R.H., 2002. Petroleum source rock
potential and generation history in the offshore Canterbury Basin. Min. Econ. Dev. N. Z. Pet. Rep., PR2707.
Sykes, R., and Johansen, P.E., 2009. Maturation characteristics of the New Zealand coal band, Part 1. Evolution of
Proc. IODP | Volume 317
Site U1352
oil and gas products. Abstr. Rep. Int. Congr. Org. Geochem., 23:571–572.
van der Lingen, G.J., Smale, D., and Lewis, D.W., 1978.
Alteration of a pelagic chalk below a paleokarst surface,
Oxford, South Island, New Zealand. Sediment. Geol.,
21(1):45–66. doi:10.1016/0037-0738(78)90033-7
van Morkhoven, F.P.C.M., Berggren, W.A., and Edwards, A.S.,
1986. Cenozoic cosmopolitan deep-water benthic foraminifera. Bull. Cent. Rech. Explor.—Prod. Elf-Aquitaine, 11.
Ward, D.M., and Lewis, D.W., 1975. Paleoenvironmental
implications of storm-scoured, ichnofossiliferous midTertiary limestones, Waihao District, South Canterbury,
New Zealand. N. Z. J. Geol. Geophys., 18(6):881–908.
Publication: 4 January 2011
MS 317-104
50
Expedition 317 Scientists
Site U1352
Figure F1. Dip Profile EW00-01-60 showing Site U1352 and proposed alternate Site CB-04A. There is no
crossing strike profile at this site. Yellow = proposed penetration, red = actual penetration (1927.5 m CSF-A).
Authorization to exceed the proposed maximum penetration of 1913 m CSF-A was granted because the Marshall Paraconformity (MP) was deeper than expected. CDP = common depth point.
CDP
2400
0
NW
2200
2000
Profile 60
SPHDIV display
(gain 180)
1800
1600
SE
0.25
CB-04A
Site U1352
0.50 18
17
19
0.75
12
16
Two-way traveltime (s)
1.00
15
14
1.25
13
1.50
11
9
1.75
8
7
6
2.00
MP (“GREEN”)
2.25
2.50
1 km
Proc. IODP | Volume 317
51
Expedition 317 Scientists
Site U1352
Figure F2. Map of drilled and proposed Expedition 317 sites, together with EW00-01 high-resolution (frequencies up to 300 Hz) MCS grid (thick straight lines), low-resolution CB-82 commercial MCS grid (thin
straight lines), exploration wells Clipper and Resolution, and Ocean Drilling Program (ODP) Site 1119. The
EW00-01 survey was designed to provide improved vertical resolution (~5 m in the upper 1 s) to enhance our
ability to define high-frequency sedimentary sequences. Also shown is the distribution of seismically resolvable
sediment drifts D1–D11, along with D8 and D9 subdrifts. Blue curved lines = crests of drift mounds, dashed
blue lines = drifts identified on CB-82 profiles. Dip Profiles EW00-01-66, EW00-01-60, EW00-01-01, EW00-0107a, and CB-82-25 are also labeled.
44°
S
Ra
ng
ita
N
ta
D5
ra
-1
eo
D9
Pa
r
D
D8 7
-1
D8-2
D83
D8
So
ut
h
Is
l
an
d
Resolution
0
20
D9
-2
0
40
0
D1
3
D9-
CB-05B/D/E
D1
D6
D3
D4
D11
CB-05C
600
CB-05F
Site U1353 (CB-01A)
CB-01B
Site U1354 (CB-02A)
CB-01C
Site U1351 (CB-03B)
1
CB-02B
0
80
0
100
ODP Site 1119
0
120
0
82
New Zealand
CB-06B
ut
h
B-
nd
5
-2
CB-04C
Clipper
60 66
45°
Site U1352 (CB-04B)
a
07
173°E
CB-04A
la
D2
CB-03A
Is
Waita
ki
So
C
0
20
45°
S
Canterbury Basin
km
171°E
172°
Drilled sites
Proc. IODP | Volume 317
173°
Proposed sites
52
Expedition 317 Scientists
Site U1352
Figure F3. Summary of core recovery, lithology, lithologic units, unit descriptions, physical property data (colorimetry, magnetic susceptibility, and NGR), and gamma ray data from downhole logging, Hole U1352B.
Downhole logging data are plotted on the WMSF depth scale. NGR = natural gamma radiation, cps = counts
per second.
Hole U1352B
50
100
150
200
250
Depth CSF-A (m)
300
350
400
450
500
550
600
650
700
750
800
1H
2H
3H
4H
5H
6H
7H
8H
9H
10H
11H
12H
13H
14H
15H
16H
17H
18H
19H
20H
21H
22H
23H
24H
25H
26H
27H
28H
29H
30H
31H
32H
33H
34H
35H
36H
37X
38X
39X
40X
41X
42X
43X
44X
45X
46X
47X
48X
49X
50X
51X
52X
53X
54X
55X
56X
57X
58X
59X
60X
61X
62X
63X
64X
65X
66X
67X
68X
69X
70X
71X
72X
73X
74X
75X
76X
77X
78X
79X
80X
81X
82X
83X
84X
85X
86X
87X
88X
89X
90X
91X
92X
93X
94X
850
Proc. IODP | Volume 317
Lith.
unit
Age
IC
IIA
Unit description
-18
-8
2
12
0
20
40
NGR
(cps)
30
80
Gamma
ray
(gAPI)
10 35 60 85
Subunit IA contains
predominantly mud-rich
sediment consisting of
mud, interbedded
mud-clay-sand, calcareous
sandy mud, shelly muds,
and muddy (sometimes
shelly) sand. Bioturbation
is common and ranges
between an ichnofabric
index of 1 and 5.
IA
IB
Magnetic
susceptibility
(instrument units)
b*
Pleistocene-Holocene
0
Lithology
Pliocene
Core
recovery
Clay
Silt
Very fine sand
Fine sand
Medium sand
Coarse sand
Very coarse sand
Granules
Pebbles
Cobbles
Grain size
Subunit IB has thicker
more homogeneous mud
intervals, as well as
greenish gray marl,
interbedded sand and
mud, mottled sandy mud,
and muddy sand.
Carbonate cementations
become common in the
greenish gray sandy muds
and muddy sands.
Bioturbation is common
and ranges between an
ichnofabric index of 1 and 4.
Subunit IC represents a
transition from muddominated sedimentation
to carbonate-rich deposits,
with an increasing
dominance of marls. Mud
is still the dominant
sediment over much of this
interval, as well as marl,
marlstone, sandy mud and
muddy sand, and (rarely)
muddy sandstone.
Cemented intervals
increase in frequency and
thickness. Bioturbation is
common and ranges
between an ichnofabric
index of 1 and 4.
Subunit IIA contains more
homogeneous calcareous
deposits, although still
containing a significant
amount of detrital sand,
silt, and clay, and with
more cemented layers
and concretions.
Bioturbated marl
dominates in Hole
U1352B. The ichnofabric
index is 1.
53
Expedition 317 Scientists
Site U1352
Figure F4. Summary of core recovery, lithology, lithologic units, unit descriptions, physical property data (colorimetry, magnetic susceptibility, and NGR), and gamma ray data from downhole logging, Hole U1352C.
Downhole logging data are plotted on the WMSF depth scale. NGR = natural gamma radiation. A. 0–1200 m
CSF-A. (Continued on next page.)
Hole U1352C
A
Core
recovery Lithology
Clay
Silt
Very fine sand
Fine sand
Medium sand
Coarse sand
Very coarse sand
Granules
Pebbles
Cobbles
Grain size
Magnetic
susceptibility
(instrument units)
b*
Lith.
unit Age
Unit description
-14
-4
6
-4 16 36 56 0
NGR
(cps)
Gamma
ray
(gAPI)
20 40 60 8 28 48 68
0
50
100
150
200
250
1D
300
350
400
Subunit IC represents a
transition from muddominated sedimentation
to carbonate-rich deposits,
with an increasing
dominance of marls. Mud
is still the dominant
sediment over much of
this interval, as well as
marl, marlstone, sandy
mud and muddy sand,
and (rarely) muddy
sandstone. Cemented
intervals increase in
frequency and thickness
downhole. Bioturbation is
common and ranges
between an ichnofabric
index of 1 and 4.
450
550
600
2R
3R
4R
5D
IC
650
700
750
800
850
900
950
1000
1050
1100
1150
1200
6R
7R
8R
9R
10R
11R
12R
13R
14R
15R
16R
17R
18R
19R
20R
21R
22R
23R
24R
25R
26R
27R
28R
29R
30R
31R
32R
33R
34R
35R
36R
37R
38R
39R
40R
41R
42R
43R
44R
45R
46R
47R
48R
49R
50R
51R
52R
53R
54R
55R
56R
57R
58R
59R
60R
61R
Proc. IODP | Volume 317
Pliocene
Depth CSF-A (m)
500
IIA
Subunit IIA contains
more homogeneous
calcareous deposits,
although still containing a
significant amount of
detrital sand, silt, and
clay, and with more
cemented deposits.
Bioturbated marlstone
dominates, with sandy
mud, mud, muddy sand,
marl, muddy sandstone,
and chalk as subordinate
lithologies. Bioturbation
is common and ranges
between an ichnofabric
index of 1 and 5.
54
Expedition 317 Scientists
Site U1352
Figure F4 (continued). B. 1200–1930 m CSF-A.
Hole U1352C
B
1300
1350
1400
Depth CSF-A (m)
1450
1500
1550
1600
1650
1700
1750
1800
1850
1900
Proc. IODP | Volume 317
Unit description
-14
-4
6
-4 16 36 56 0
NGR
(cps)
Gamma
ray
(gAPI)
20 40 60 8 28 48 68
Pliocene
1250
61R
62R
63R
64R
65R
66R
67R
68R
69R
70R
71R
72R
73R
74R
75R
76R
77R
78R
79R
80R
81R
82R
83R
84R
85R
86R
87R
88R
89R
90R
91R
92R
93R
94R
95R
96R
97R
98R
99R
100R
101R
102R
103R
104R
105R
106R
107R
108R
109R
110R
111R
112R
113R
114R
115R
116R
117R
118R
119R
120R
121R
122R
123R
124R
125R
126R
127R
128R
129R
130R
131R
132R
133R
134R
135R
136R
137R
138R
139R
140R
141R
142R
143R
144R
145R
146R
147R
148R
IIB
Miocene
1200
Magnetic
susceptibility
(instrument units)
b*
Lith.
unit Age
Subunit IIB is distinguished by
the common occurrence of thin
layers of dark mudstone between
1170 and 1392 m, although
marlstone still dominates. An
increase in current bedding
features within the marlstone,
especially below the interval of
common mudstone occurrences,
means that other lithologies also
appear, including very fine to fine
sandstone and (rarely) medium
to coarse sandstone. Slumps
occur at 1376, 1440, and 1541
m. An unconformity detected by
biostratigraphy around 1400 m is
associated with a change in
mineralogy and an increase in
organic matter (plant derived) in
the cores. A banded appearance
possibly caused by incipient
pressure solution seams begins
to develop from 1600 m.
Bioturbation is common and
ranges between an ichnofabric
index of 1 and 5.
IIC
Subunit IIC contains a transition
from sandy marlstones to sandy
limestones, with minor mudstone
and very fine sandstone beds. This
unit is distinguished from all
overlying units by a high
percentage of glauconite, and by
glauconitic sandstone or marlstone
beds that occur with increasing
frequency toward the base of the
unit and sometimes crosscut the
limestone texture. Bioturbation is
common and ranges between an
ichnofabric index of 1 and 5.
III
Unit III comprises foraminiferbearing nannofossil limestone with
only minor quantities of silt-sized
quartz and glauconite or pyrite
infilling foraminifers. This unit is
correlative to the onshore Amuri
Limestone. Toward the base of the
unit, mudstone or marlstone beds
appear, becoming more common
toward the base of the hole.
Bioturbation is common and ranges
between an ichnofabric index of
1 and 5.
Eocene-Oligocene
Core
recovery Lithology
Clay
Silt
Very fine sand
Fine sand
Medium sand
Coarse sand
Very coarse sand
Granules
Pebbles
Cobbles
Grain size
55
Expedition 317 Scientists
Site U1352
Figure F5. Core recovery and lithology shown as a proportion of the recovered interval (lithology [%] × recovery [%]/100) in Hole U1352B. Depth scale in CSF-A (m) is variable, but for graphical purposes each core is
depicted by an equally thick horizontal bar.
Hole U1352B
Core
recovery (%) Core
0
50
100
0
Weighted lithology (%)
10
20
30
40
50
60
70
80
90
100
Lith.
unit
1H
5H
IA
10H
100
15H
20H
200
25H
IB
30H
Depth CSF-A (m)
35H
300
40X
45X
400
50X
55X
500
60X
65X
600
IC
70X
75X
700
80X
85X
IIA
90X
800
Interbedded clay and mud
Interbedded silt and mud
Interbedded sand and mud
Very fine sand
Proc. IODP | Volume 317
Sandy mud
Marl
Muddy sand
Mud
56
Expedition 317 Scientists
Site U1352
Figure F6. Core photographs of various mud lithofacies from Subunit IA. All pictures have the same vertical
scale. A. Mud-sand-clay interbedding, dragged down at the edges of the core liner by drilling (interval 317U1352B-2H-3, 114–134 cm). B. Interbedded mud and clay (interval 317-U1352A-1H-2, 91–111 cm). C. Mottled
massive mud (interval 317-U1352B-7H-1, 1–21 cm). D. Homogeneous mud (interval 317-U1352B-9H-6, 5–25
cm).
B
C
D
5 cm
A
Proc. IODP | Volume 317
57
Expedition 317 Scientists
Site U1352
Figure F7. Core photographs of green calcareous sandy beds with sharp to bioturbated bases and concretions
in the lower part of Unit I. A. Green sandy material in burrows below the contact between gray mud (below)
and green muddy sand (above) (interval 317-U1352B-7H-6, 43–125 cm). B. Bioturbation of green sandy material into underlying gray mud (interval 317-U1352B-28H-1, 25–65 cm). C. Concretions in a green sandy marlstone (interval 317-U1352B-52X-7, 2–17 cm). D. Sharp contact between green muddy sandstone and gray mud
(interval 317-U1352B-17H-4, 1–21 cm). B, C, and D have the same vertical scale.
A
B
C
5 cm
5 cm
D
Proc. IODP | Volume 317
58
Expedition 317 Scientists
Site U1352
Figure F8. Core photographs of deformation in Subunit IA. All photographs are at the same vertical scale.
A. Normal fault offsetting layers of mud, clay, and sand (interval 317-U1352A-2H-2, 15–50 cm). B. Distortion
of clay and mud layers from drilling disturbance (interval 317-U1352A-2H-2, 49–61 cm). C. Folding below a
thick sand layer (interval 317-U1352A-2H-4, 90–127 cm). D. Slump folds in interbedded sand and mud (interval 317-U1352B-2H-3, 20–40 cm).
A
B
C
5 cm
D
Proc. IODP | Volume 317
59
Expedition 317 Scientists
Site U1352
Figure F9. Core photographs of various interbedded lithofacies in Subunit IA. All photographs are at the same
vertical scale. A. Mud-sand-clay interbedding (interval 317-U1352B-2H-5, 0–25 cm). B. Interbedded mud and
clay (interval 317-U1352A-2H-5, 20–45 cm). C. Thinly interbedded clay-mud and sand-mud (interval 317U1352B-8H-3, 5–30 cm).
B
C
5 cm
A
Proc. IODP | Volume 317
60
Glauconite
(%)
CaCO3
(%)
0
0
40
80
0
20
Mica
(%)
0
40
Ferromagnesian
minerals
(%)
0
8
0
Dense
minerals
(%)
10
Quartz and
feldspar
(%)
0
40
80 0
Clay
(%)
40
Siliceous
bioclasts
(%)
Sand
(%)
80 0
40
80 0
10
20
Lith.
unit
30
Expedition 317 Scientists
Proc. IODP | Volume 317
Figure F10. Mineral and textural percentage estimates based on smear slide observations, Site U1352. CaCO3 estimates are plotted against data
from coulometry analyses for comparison (see “Geochemistry and microbiology”). Clay and sand fractions tend to vary opposite each other,
siliceous bioclasts and ferromagnesian minerals both diminish in Unit II, and glauconite increases abruptly in Subunit IIC. Mica refers to undifferentiated micas, biotite, muscovite, and chlorite.
IA
IB
400
Depth CSF-A (m)
IC
800
IIA
1200
IIB
1600
IIC
III
2000
Coulometry, Hole U1352B
Coulometry, Hole U1352C
Thin section
Carbonate zones
Subunit boundary
Unit boundary
61
Site U1352
Hole U1352A
Hole U1352B
Hole U1352C
Maximum peak intensity (counts)
Quartz
800
Total clay
1600
80
160
Total mica
3500
Chlorite
7000
2000
Plagioclase
4000
2000
Hornblende
4000
120
240
Calcite
4500
Lith.
unit
9000
0
IA
200
Expedition 317 Scientists
Proc. IODP | Volume 317
Figure F11. Relative maximum peak intensity in XRD analyses for common minerals, Holes U1352B and U1352C.
IB
400
IC
Depth CSF-A (m)
600
800
IIA
1000
1200
1400
IIB
1600
IIC
1800
III
Hole U1352B
Hole U1352C
Subunit boundary
Unit boundary
Site U1352
62
Expedition 317 Scientists
Site U1352
Figure F12. Thin section photomicrographs illustrating changes in lithology and diagenesis in the succession.
A. Partly dissolved aragonitic gastropod set in a muddy micritic matrix with sparse feldspar and quartz silt
(marlstone; Subunit IB); moldic pores occur elsewhere in the slide. Note the authigenic opaque minerals
(pyrite?) partly lining shell interior (Sample 317-U1352B-17H-4, 69 cm). B. Bioturbated marlstone in Subunit
IIA showing the variability of texture/lithology present on a microscopic scale in these mixed (zoned) carbonate/siliciclastic rocks: sandy marlstone (left), marlstone (center), carbonate-cemented sandstone (right)
(Sample 317-U1352C-26R-1, 23 cm). C. Sandy marlstone in Subunit IB consisting of sand-sized foraminifers
and feldspar/quartz grains set in a fine micritic matrix. Reddish brown milliolid foraminifers are less competent
than other foraminifer types and show evidence of fracturing, interpenetration, and possibly pressure solution.
The more competent foraminifers are locally recrystallized and exhibit fine carbonate overgrowths (far right)
and local glauconitic infill (upper left). High residual porosity in this rock is mostly intraparticle within foraminifer tests (Sample 317-U1352B-42X-5, 5 cm). D. Section of high-amplitude stylolite passing through pelagic
limestone with sparse foraminifers in Unit III. The stylolite is marked by brownish clay, along which the foraminifers are truncated, presumably by pressure solution (Sample 317-U1352B-141R-2, 15 cm). E, F. Abrupt
high-angle contact between muddy limestone to limestone (transitional) below (within host rock) and glauconitic sandy marlstone above (potential injection material) in Subunit IIC. Limestone color is a function of the
foraminifer/nannofossil ratio (higher in light limestone; lower in dark limestone) (Sample 317-U1352C-137R2, 58 cm).
A
B
2 mm
C
5 mm
D
2 mm
F
E
5 mm
Proc. IODP | Volume 317
2 mm
5 mm
63
Expedition 317 Scientists
Site U1352
Figure F13. Total clay minerals normalized by calcite content from XRD analyses, Holes U1352B and U1352C.
Lith.
unit
Total clays/calcite
0
0
1
2
3
4
5
6
IA
200
IB
400
IC
600
Depth CSF-A (m)
800
IIA
1000
1200
1400
IIB
1600
IIC
1800
III
2000
Hole U1352B
Proc. IODP | Volume 317
Hole U1352C
64
Expedition 317 Scientists
Site U1352
Figure F14. Core recovery and lithology shown as a proportion of the recovered interval (lithology [%] × recovery [%]/100) in Hole U1352C. Depth scale in CSF-A (m) is variable, but for graphical purposes each core is
depicted by an equally thick horizontal bar.
Hole U1352C
Core
recovery (%)
0
50
Core
100
Weighted lithology (%)
0
10
20
30
40
50
60
70
80
90
100
Lith.
unit
2R
600
700
IC
10R
15R
800
20R
25R
900
30R
IIA
35R
1000
40R
45R
50R
1100
Depth CSF-A (m)
55R
60R
1200
65R
70R
1300
80R
85R
1400
IIB
90R
95R
1500
105R
110R
1600
115R
120R
1700
125R
130R
1800
IIC
135R
140R
1900
Proc. IODP | Volume 317
III
145R
Marl/Marlstone
Very fine sand
Medium sand
Chalk/Limestone
Sandy mud
Mudstone
Muddy sand
65
Expedition 317 Scientists
Site U1352
Figure F15. Representative core photographs showing features of sandy marlstones encountered in Unit II. All
photographs are at the same vertical scale. A. Dark mudstone beds with sharp bases and burrows infilled with
dark mud extending into underlying sandy marlstone (interval 317-U1352C-81R-1, 28–66 cm). B. Bioturbated
sandy marlstone (interval 317-U1352C-42R-2, 1–16 cm). C. Large, pale burrows overprinting a small, darker
burrow presumably belonging to an earlier episode of burrowing (interval 317-U1352C-59R-3, 31–46 cm).
D. Lighter colored layer with higher carbonate content (interval 317-U1352C-44R-3, 10–25 cm). E. Patches of
sand and wavy, discontinuous laminations (interval 317-U1352C-95R-6, 2–17 cm). F. Laminated sandy marlstone to calcareous sandstone (interval 317-U1352C-114R-6, 84–98 cm). G. Wavy laminations and sand lenses
in a sandy marlstone (interval 317-U1352C-114R-7, 15–32 cm).
A
B
F
D
G
5 cm
E
C
Proc. IODP | Volume 317
66
Expedition 317 Scientists
Site U1352
Figure F16. Core photographs, interpretation, and sketches of soft-sediment deformation intervals in Subunit
IIB. The base of Section 317-U1352C-94R-7 (1438.2 m) was described as surface U1352C-S12. A. Fluidized structures associated with soft-sediment deformation (Core 317-U1352C-94R). Shades of gray indicate color changes
in lithology. B. Very fine sandy marlstone showing soft-sediment deformation structures: recumbent and isoclinal folding, fluidized structures, and coarse to medium sand (Section 317-U1352C-106R-4). C. Poorly sorted
sandy marlstone with fine to coarse, angular, lithic clasts near the base of an interval of soft-sediment deformation extending from Sections 317-U1352C-105R-1 to 106R-5 (interval 317-U1352C-106R-4, 49–59 cm).
A
C
cm
50
cm 25
30
40
35
45
55
50
55
B
cm 20
25
30
Proc. IODP | Volume 317
35
40
45
50
55
60
65
70
75
80
85
90
95
100
67
Expedition 317 Scientists
Site U1352
Figure F17. Core photographs showing characteristic features of Subunit IIC. A. Characteristic striped pattern
of marlstones or limestones and glauconitic sandstone or marlstone layers (interval 317-U1352C-135R-3, 67–
128 cm). B. Laminated layers of glauconitic sandstone (interval 317-U1352C-135R-3, 66–76 cm). C. Intersecting and high-angle layers of glauconitic sandstone (interval 317-U1352C-139R-5, 19–29 cm). D. Converging layers of glauconitic sandstone (interval 317-U1352C-135R-2, 62–72 cm). E. High-angle layers between
blocks of glauconitic limestone (interval 317-U1352C-136R-3, 6–16 cm). F. Example of truncation of limestone
fabric (burrow) by glauconitic sandstone layer (interval 317-U1352C-134R-1, 71–81 cm). B–F are at the same
vertical scale.
A
B
C
D
E
5 cm
5 cm
F
Proc. IODP | Volume 317
68
Expedition 317 Scientists
Site U1352
Figure F18. Core photographs of contact between Units II and III. Although only two limestone pieces and
some rubble were recovered, it is possible the drilled succession included relatively unconsolidated glauconitic
sandstone between the two limestones. A. Unconformity (arrow) between glauconitic sandy limestone
(Miocene) and white limestone (Oligocene) (interval 317-U1352C-140R-2, 0–55 cm). B. Glauconitic limestone
above the unconformity, Subunit IIC (interval 317-U1352C-140R-1, 6–26 cm). C. Stylolitic white limestone
below the unconformity, Unit III (interval 317-U1352C-141R-2, 14–34 cm). D. Stylolitic white limestone with
purple banding and pyrite concentrated in a stylolite, Unit III (interval 317-U1352C-144R-2, 41–61 cm).
E. Muddy limestone with layers of marlstone and purple banding, Unit III (interval 317-U1352C-148R-1, 62–
68 cm). B–E are at the same vertical scale.
B
C
D
E
5 cm
5 cm
A
Proc. IODP | Volume 317
69
Expedition 317 Scientists
Site U1352
A
0
Core
Clay
Silt
Very fine sand
Fine sand
Medium sand
Coarse sand
Very coarse sand
Figure F19. Comparison of downhole gamma ray logs, magnetic susceptibility, lithology, lithologic surfaces
identified in core description, and predicted depths of seismic sequence boundaries, Hole U1352B. Downhole
logging data are plotted on the WMSF depth scale. A. 0–150 m CSF-A. (Continued on next two pages.)
Magnetic
susceptibility
Color of mud
(instrument units)
Blue Green
0
10
20
Gamma ray
(gAPI)
30 40 20 40
60
80 100
Lith.
unit
Seismic
Lithologic sequence
boundary
surface
1H
10
2H
20
3H
30
Depth CSF-A (m)
40
4H
5H
50
6H
60
7H
IA
S1
E1
70
8H
80
9H
90
100
U19
U19
10H
11H
12H
110
13H
120
IB
14H
130
15H
140
U18
U18
16H
E2
S2
150
Proc. IODP | Volume 317
Slump
Shell fragment
Shell layer
Burrow
70
Expedition 317 Scientists
Site U1352
B
150
Core
Clay
Silt
Very fine sand
Fine sand
Medium sand
Coarse sand
Very coarse sand
Figure F19 (continued). B. 150–300 m CSF-A. (Continued on next page.)
Magnetic
susceptibility
Color of mud (instrument units)
Blue Green
0
10 20
Gamma ray
(gAPI)
30 40 20 40
60
80 100
Lith.
unit
Seismic
Lithologic sequence
surface boundary
17H
160
18H
170
19H
20H
180
21H
190
22H
U17
U17
E3
S3
200
23H
Depth CSF-A (m)
210
24H
220
25H
IB
230
240
250
26H
27H
28H
E4
S4
U16
U16
29H
260
30H
270
31H
32H
280
33H
290
300
34H
35H
36H
37X
Proc. IODP | Volume 317
Slump
Shell fragment
Shell layer
Burrow
71
Expedition 317 Scientists
Site U1352
C
300
Core
Clay
Silt
Very fine sand
Fine sand
Medium sand
Coarse sand
Very coarse sand
Figure F19 (continued). C. 300–450 m CSF-A.
Magnetic
susceptibility
Color of mud (instrument units)
Blue Green
37X
0
10
20
Gamma ray
(gAPI)
30 40 20 40
60
80 100
Lith.
unit
Seismic
Lithologic sequence
surface boundary
38X
310
39X
320
40X
330
41X
340
42X
350
43X
Depth CSF-A (m)
360
44X
370
IB
45X
380
46X
390
47X
400
48X
410
49X
420
50X
E5
S5
430
U15
51X
440
450
52X
IC
53X
Proc. IODP | Volume 317
Slump
Shell fragment
Shell layer
Burrow
U14
72
Expedition 317 Scientists
Site U1352
Figure F20. Core photographs and drawings of Type A, B, and C contacts and facies associations. Type A surfaces are characterized by sharp, basal contacts commonly separating two distinct lithologies below >1 m thick
deposits. Type B surfaces represent amalgamation packages of >1 m thick thinly bedded (decimeter scale), distinctly contrasting lithologies (e.g., turbidites, interbedded mud and clay, and well-cemented marlstone/chalk)
with sharp basal contacts. Type C surfaces are always sharp and separate contrasting lithologies such as sand
and mud or marlstone and chalk. However, the total thickness of beds associated with these contacts could not
be clearly delineated on board ship. A. Surface U1352B-S1 (interval 317-U1352B-7H-6, 76–118 cm). Shells are
present, and extensive bioturbation occurs beneath the contact. B. Surface U1352B-S3 (interval 317-U1352B23H-1, 116–148 cm). Heavy bioturbation occurs beneath the contact. C. Thinly bedded (1–2 cm thick) sand
beds and mud interval (interval 317-U1352B-3H-4, 113–140 cm). D. Very fine grained, well-sorted, quartz-rich
sand beds (2–5 cm thick) (interval 317-U1352B-2H-6, 17–48 cm). (Figure shown on next page.)
Proc. IODP | Volume 317
73
Expedition 317 Scientists
Site U1352
Figure F20 (continued). (Caption shown on previous page.)
Type A
A
cm
80
Type A surface
}
>1 m thick
B
cm
120
85
125
90
S3
Contact
95
130
S1
135
105
140
Type B surface
}
>1 m thick
110
145
115
Type B
C
Contact
Type C surface
}
Total
thickness
undetermined
D
cm
cm
115
20
120
25
125
30
Contact
35
130
40
Contact
135
45
140
Proc. IODP | Volume 317
74
Expedition 317 Scientists
Site U1352
Figure F21. Thicknesses of muddy sand, sandy mud, and clay beds in each core of Hole U1352B. A. Sand-mudclay beds illustrating lithologies associated with Type A. B. Interbedded intervals (sand/mud and mud/clay) illustrating lithologies associated with Type B. Red = sand or muddy sand, blue = sandy mud or clay.
Bed thickness
(cm)
A
0
200
400
B
Interbed thickness
(cm)
600
0
0
Core 317-U1352B-7H
0
400
800
Sand/Mud
Type B
10
100
200
Mud/Clay
Depth (m)
Mud-sand, Type A
20
Type B
30
Clay
40
300
50
Depth (m)
Mud , Type A
400
Sand,
Type A
500
600
700
Sand,
Type A
800
Proc. IODP | Volume 317
75
Expedition 317 Scientists
Site U1352
Figure F22. Core photographs of an unusual very calcareous interval likely corresponding to the Pliocene/
Miocene boundary. All photographs are at the same vertical scale. A. Interval 317-U1352C-73R-2, 18–48 cm
(top of calcareous section). B. Interval 317-U1352C-73R-3, 53–83 cm. C. Interval 317-U1352C-73R-4, 5–35 cm
(base of calcareous section, designated as surface U1352C-S11).
A
B
C
5 cm
S11
Proc. IODP | Volume 317
76
Expedition 317 Scientists
Site U1352
Figure F23. Hole U1352B interpretation of correlation between sand bed thickness, surfaces U1352B-S1 to
U1352B-S6, and seismic sequence boundaries U19–U13.
Hole U1352B
Sand bed thickness (cm)
0
0
200
Core 7H
400
600
S1 (64 m)
U19 (68 m)
100
U18 (142 m)
S2 (147 m)
200
U17 (195 m)
S3 (200 m)
U16 (249 m)
S4 (250 m)
Depth (m)
300
400
U15 (428 m)
S5 (428.5 m)
U14 (448 m)
S5.1 (453 m)
S6 (482 m)
500
U13 (500 m)
600
700
800
Proc. IODP | Volume 317
77
Expedition 317 Scientists
Site U1352
Figure F24. Core recovery, epochs, calcareous nannofossil (NN) zones, and New Zealand (NZ) stage correlation for planktonic foraminifers and bolboformids (PF) and benthic foraminifers (BF), Site U1352. Solid wavy lines = hiatuses between
zonal boundaries, dashed wavy lines = hiatuses within biozones, dashed straight line = poorly constrained zonal boundary. See Figure F6 in the “Methods” chapter for NZ stage abbreviations. A. 0–1050 m CSF-A. (Continued on next page.)
2H
48X
24H
49X
25H
50X
26H
51X
27H
52X
BF
600
850
73X
33H
58X
60X
40X
350
150
17H
550
43X
63X
64X
18H
44X
65X
19H
22H
66X
67X
46X
Wn-Wc
21H
45X
NN19
20H
84X
750
69X
600
86X
87X
88X
2R
89X
3R
90X
68X
47X
400
83X
85X
middle
42X
16H
82X
62X
Pliocene
NN20
15H
61X
41X
80X
32R
8R
33R
9R
34R
10R
35R
81X
14H
14H
700
59X
39X
13H
Wp-Wc
Wp-Wc
500
79X
upper Wo-Wp
38X
57X
Wn-Wc
34H
35H
36H
300 37X
7R
4R
800
11R
12R
950
36R
37R
13R
38R
14R
39R
15R
40R
16R
1000
Wp-Wc
78X
31R
Pliocene
Wq
56X
NN16-NN19
13H
77X
32H
NN16
12H
Pleistocene-Holocene
12H
Wq
9H
55X
6R
Wp-Wc
9H
76X
upper Wo-Wp
31H
8H
30R
900
NN16
30H
75X
early
54X
650
Pliocene
middle
NN21
8H
29R
53X
Wq
450
29H
7H
11H
5D
74X
NN19
250 28H
7H
11H
middle
72X
Pleistocene
6H
6H
10H
25R
28R
5H
10H
BF
71X
5H
5H
PF
27R
4H
200
PF
4H
4H
Depth CSF-A (m)
400
NZ stage
NN correlation
Hole
BF
U1352C Epoch zone PF
24R
26R
3H
100
BF
3H
3H
50
PF
23H
Hole
Hole
NN
U1352B U1352C Epoch zone
4R
70X
Core
recovery
NZ stage
correlation
upper Wo-Wp
200
NN
Hole
Hole
U1352B U1352C Epoch zone
Core
recovery
NZ stage
correlation
upper Wo
BF
Core
recovery
NZ stage
correlation
NN16
PF
NN
Hole
U1352B Epoch zone
NN19
0
NN
Hole
Hole
Hole
U1352A U1352B U1352D Epoch zone
1H
1H
1H
2H
2H
Core
recovery
NZ stage
correlation
NN12-NN15
Core
recovery
Pleistocene
A
41R
17R
42R
18R
43R
19R
44R
20R
45R
91X
92X
21R
1050
46R
93X
22R
94X
23R
24R
840
Proc. IODP | Volume 317
78
Expedition 317 Scientists
Site U1352
Figure F24 (continued). B. 1050–1930 m CSF-A. The Marshall Paraconformity (at least 12 m.y. missing) is between Samples 317-U1352C-139R-CC and 140R-CC (1848.49–1852.71 m).
104R
126R
148R
105R
127R
86R
58R
131R
132R
NN5
Tk-Tt
?NN8-NN12
89R
111R
133R
112R
1800 134R
1600
Sc?-lower Sl
91R
113R
92R
114R
lower Tt
64R
93R
65R
66R
Proc. IODP | Volume 317
136R
115R
137R
116R
138R
96R
NN7
1450
117R
1650
NN4NN5
Wo
95R
NN6-
69R
70R
135R
94R
68R
1250
Ar-Wo
90R
1400
63R
67R
130R
middle
88R
110R
62R
129R
87R
59R
1200
Sl
108R
109R
61R
1750
Ar-Wo
1550 107R
Miocene
85R
128R
1930
early
1350
106R
lower Lwh
147R
Wo-Ar
84R
?NN8-NN12
83R
upper Tk
82R
Miocene
late
57R
60R
NN4-NN5
125R
Miocene
Wp-Wc
upper Wo
Pliocene
early
Depth CSF-A (m)
55R
81R
Ar
103R
146R
80R
54R
1150
1900
124R
78R
79R
56R
145R
123R
1700
1500 102R
77R
53R
144R
NP23
76R
101R
122R
142R
143R
NP19
75R
121R
141R
Po-lower Pl
52R
1300
100R
140R
Oligocene
early
99R
120R
1850
Sc-Iower Sl
98R
119R
NN2?-NN3
51R
BF
NZ stage
NN correlation
Hole
U1352C Epoch zone PF
BF
Eocene
late
NN12-NN15
1100
97R
74R
PF
1650
118R
73R
50R
BF
Pl
72R
49R
PF
96R
NN4
48R
1450
Core
recovery
NZ stage
correlation
NN3
71R
BF
middle
47R
PF
NN
Hole
U1352C Epoch zone
Sw
70R
NN6-NN7
1250
Core
recovery
NZ stage
correlation
NN6
BF
NN
Hole
U1352C Epoch zone
late
PF
46R
NN
Hole
U1352C Epoch zone
Core
recovery
NZ stage
correlation
lower Wo
1050
NN
Hole
U1352C Epoch zone
Core
recovery
NZ stage
correlation
lower Wo
Core
recovery
Pliocene
early
B
139R
1850
79
Expedition 317 Scientists
Site U1352
Figure F25. Planktonic foraminiferal abundance relative to total foraminifers and oceanicity, Site U1352.
Pliocene
Holocene-Pleistocene
Miocene
Ol. Eoc.
100
Oceanic
Planktonic foraminiferal abundance (%)
90
80
Suboceanic
70
60
50
Extraneritic
40
30
Outer neritic
20
10
Inner neritic
Hole U1352B
Hole U1352C
0
0
100
200
300
400
500
600
700
800
900 1000 1100 1200 1300 1400 1500 1600 1700 1800 1900 2000
Depth CSF-A (m)
Proc. IODP | Volume 317
80
Expedition 317 Scientists
Site U1352
Figure F26. Thin section photomicrograph of sparry calcite and glauconite-infilled planktonic foraminifer in
a li