ArticlePDF Available
Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean
Ajay K. Singh
a
, Franco Marcantonio
a,
, Mitchell Lyle
b
a
Department of Geology and Geophysics, Texas A&M University, College Station, TX 77843, USA
b
Department of Oceanography, Texas A&M University, College Station, TX 77843, USA
abstractarticle info
Article history:
Received 31 January 2011
Received in revised form 8 June 2011
Accepted 20 June 2011
Available online xxxx
Keywords:
thorium-230
sediment focusing
boundary scavenging
mass accumulation rate
Panama basin
eastern equatorial Pacic Ocean
Age-model derived sediment mass accumulation rates (MARs) are consistently higher than
230
Th-normalized
MARs in the Equatorial Pacic Ocean during the past 25 ka. The offset, being highest in the Panama Basin,
suggests a signicant role for deep-sea sediment redistribution (i.e., sediment focusing) in this region. Here,
we test the hypothesis that downslope transport of sediments from topographically high regions that
surround the Panama Basin is the cause of higher-than-expected xs
230
Th inventories over the past 25 ka in
the deeper parts of the basin. We nd little difference in xs
230
Th inventories between the highest and lowest
reaches of the basin. Furthermore, there is no correlation between xs
230
Th-derived sediment focusing factors
and water depth which suggests that the topographic highs do not serve as a source of xs
230
Th. A spatial
analysis suggests that there may be an enhanced scavenging effect on xs
230
Th concentrations in sediment
closest to the equator where productivity is the highest, although further data is necessary to corroborate this.
At the equator xs
230
Th-derived focusing factors are high and range from about 1 to 5 during the Holocene and
about 1 to 11 during the last glacial. In contrast, non-equatorial cores show a smaller range in variability from
about 0.7 to 2.8 during the Holocene and from 0.7 to 3.6 during the last glacial. Based on
232
Th ux
measurements, we hypothesize that the location at which eolian detrital uxes surpass the riverine detrital
uxes is approximately 300 km from the margin. While riverine uxes from coastal margins were higher
during the Holocene, eolian uxes were higher during the last glacial.
© 2011 Elsevier B.V. All rights reserved.
1. Introduction
Essentially all sediments that reach the pelagic ocean oor are
either derived from the continents through weathering processes or
formed as a result of biological productivity in surface water.
Contemporaneous climatic conditions largely affect the processes
and mechanisms that bring these sediments to the ocean oor. For
example, sediment intervals that record higher biogenic uxes are
often interpreted as being deposited at a time during which export
production to the seaoor was increased. Similarly, intervals during
which lithogenic particle uxes are high can represent an intensied
transport of continental material via rivers (if close to a continental
margin) and/or wind. Thus, reconstruction of particle uxes from
oceanic sedimentary archives can broaden our understanding of past
climate conditions.
Historically and to the present, sedimentary mass accumulation
rates (MARs) have been estimated by multiplying the linear
sedimentation rate (LSR), estimated using dated horizons (oxygen-
isotope- or radiocarbon-derived), with sediment dry bulk density.
This method measures the amount of sediment preserved at the sea
oor but does not discriminate between vertically falling particles and
those redistributed by a variety of horizontal advection processes. The
xs
230
Th constant-ux proxy (CFP) method of determining mass
accumulation rates is thought to see throughsuch sediment
redistribution processes and is purported to measure the true vertical
ux (Bacon, 1984; Francois et al., 2004). The idea behind this
constant-ux proxy lies in the different geochemical behavior of
thorium and uranium in the oceanic water column. In seawater,
230
Th
is produced by the α-decay of
234
U. Unlike uranium, which has a
constant seawater concentration, thorium is extremely particle
reactive, and the decay product,
230
Th, is rapidly scavenged onto
sinking particles so that the ux of
230
Th to the ocean oor is identical
to its rate of production in the water column (Bacon, 1984). Hence,
MARs within an interval of sediment can be calculated by dividing the
known production rate of
230
Th by the concentration of
230
Th within
thesameinterval.Hence,theverticalux of any component
preserved in the sediment, F
i
,can be calculated theoretically using
the following equation:
F1=conc
ðÞ
iβZ
xs230Th0
 ð1Þ
in which (conc)
i
is the concentration of component i; βis the production
rate of
230
Th in the water column (0.0267 dpm m
3
yr
1
); Z is the
Earth and Planetary Science Letters xxx (2011) xxxxxx
Corresponding author at: Department of Geology and Geophysics, 3115 MS, TAMU,
Texas A&M University, College Station, TX 77843, USA. Tel.: +1 979 845 9240; fax: +1
979 845 6162.
E-mail addresses: asingh1@neo.tamu.edu (A.K. Singh),
marcantonio@geos.tamu.edu (F. Marcantonio), mlyle@ocean.tamu.edu (M. Lyle).
EPSL-11000; No of Pages 12
0012-821X/$ see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.epsl.2011.06.020
Contents lists available at ScienceDirect
Earth and Planetary Science Letters
journal homepage: www.elsevier.com/locate/epsl
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
water depth in m; and [xs
230
Th
o
] is the measured sedimentary
230
Th
activity corrected for decay, in situ production of
230
Th from authigenic
234
U, and detrital
230
Th. The elegance of Eq. 1is that the derived
sedimentary ux is, by denition, solely the vertical component of the
preserved sedimentary ux. Furthermore, one can solve for the
normalizedsedimentary ux by measuring the concentration of
xs
230
Th alone. The extent that such normalization works depends on
how realistic the assumption is that the production of
230
Th in the water
column is equal to the ux of the scavenged
230
Th to the underlying
sediments (Bacon, 1984;Francois et al., 2004).If the postulated behavior
of oceanic
230
Th is correct, an added benet to the constant-ux proxy
methodology is that syndepositional sediment redistribution can be
quantied by integrating the xs
230
Th inventory within an interval of
sediment and comparing it to the integrated production of
230
Th in the
overlying water column over the time of accumulation (Suman and
Bacon, 1989). Indeed, the ratio of these two parameters has been
dened by a physical parameter called the focusing factor, (Ψ)
(Suman and Bacon, 1989). A Ψvalue of one implies that sediment has
not been redistributed at the studied site. A Ψvalue greater than one
implies sediment in excess of what has been delivered vertically has
been advected by deep-sea horizontal advection (i.e., focusing) to the
studied site, while a Ψvalue less than one implies winnowing or
removal of sediment from the studied site at the time of sediment
deposition (Francois et al., 2004). Model studies have shownthat 70% of
the ocean oor receives a
230
Th ux within 30% of its production in the
water column (Henderson and Anderson, 2003; Siddall et al., 2008),
implying a sensitivity of the xs
230
Th proling technique that is typically
within +/30%. This estimate must be considered somewhat tentative
given the reliance of this result on the assumption that isopycnal and
vertical diffusion is a reasonable approximation of ocean mixing
processes inherent in many ocean models (Siddall et al., 2008).
Although age-model-derived and xs
230
Th-normalized MARs have
been widely used in paleoceanographic research, in some cases the
differently calculated MARs are signicantly different, and, therefore,
yield competing interpretations. Perhaps the best known of these
discrepancies exists in the equatorial (west, central and east) Pacic
Ocean (Broecker, 2008; Francois et al., 2007; Higgins et al., 1999;
Kienast et al., 2007; Koutavas et al., 2002; Koutavas and Sachs, 2008;
Kowsmann, 1973; Loubere et al., 2004; Lyle et al., 2005, 2007;
Marcantonio et al., 1996; Marcantonio et al., 2001a; Paytan et al.,
1996; Thomas et al., 2000). Here, xs
230
Th-derived focusing factors,
suggest that horizontal sediment transport almost always is higher
(sometimes several times higher) than the vertical ux. The highest
focusing factors (as high as 5.5; Kienast et al., 2007) are observed
during the last glacial in the eastern equatorial Pacic (EEP) Ocean in
the Panama Basin.
In the Panama Basin, using age-model-derived MARs, many
investigators have concluded that particle uxes during the last
glacial were as much as 100% higher than those during the Holocene,
and are caused by enhanced primary productivity (Lyle, 1988; Lyle et
al., 2002; Paytan et al., 1996; Pedersen, 1983). However, xs
230
Th
normalized MARs for sediments deposited during the last glacial
suggest calcite uxes that are 3050% lower than those during the
Holocene (Loubere et al., 2004). These authors contend that the
higher glacial age-model-derived uxes are due to sediment focusing
processes in the Panama Basin. In addition, Kienast et al. (2007)
reexamined several sites that were studied by others (Loubere et al.,
2004; Lyle et al., 2005) in the Panama Basin and came to a similar
conclusion; namely, that xs
230
Th-normalized MARs are lower and less
variable than age-model-derived MARs, indicating varying degrees of
sediment focusing.
Lyle et al. (2005) disagree with the interpretation that sediment
focusing is widespread in the Panama Basin, and argue that xs
230
Th
normalization overestimates the degree to which sediment redistri-
bution processes are occurring in the EEP. They reason that the
observed larger-than-expected inventories of sedimentary xs
230
Th in
the EEP, that are in excess of those expected from a constant water
column production rate of
230
Th, can be attributed to increased
boundary scavenging at the surface due to increased productivity
close to the equator, in agreement with an analysis by Broecker
(2008). However, within an efcient (low resolution) ocean circula-
tion model, Siddall et al. (2008) found that particle scavenging effects
are not sufcient to explain the additional xs
230
Th inventories
measured in the Panama Basin. Using the Bern3D ocean model, they
considered particle scavenging over a broad range of particle uxes
reaching up to 10 times higher than actual measurements in the
equatorial Pacic region. Even at the highest end of this range of
particle uxes, the model by Siddall et al. (2008) suggests only a two-
fold increase in the ux of
230
Th over the production of
230
Th in water
column due to particle scavenging effects.
Kienast et al. (2007) propose that downslope transport of
sediment from the east-west trending Carnegie Ridge, which forms
the southern boundary of the Panama Basin, might explain the
additional xs
230
Th in sediments of the Panama Basin. In this study, we
test this downslope transport hypothesis by measuring xs
230
Th
inventories of sediments deposited on the Cocos and Carnegie
Ridgesregional topographic highs that surround the Panama Basin.
In general, xs
230
Th inventories in sediment from the tops of ridges
suggest sediment focusing factors that are greater than 1 for both the
Holocene and glacial sediments. More importantly, sediment xs
230
Th
inventories on the ridge tops are similar to those in the previously
studied deeper cores (Kienast et al., 2007). If ridge tops were the
source of extra xs
230
Th inventory in the basin, one would expect their
focusing factors to be less than one and/or lower than those measured
in the basin. We explore the potential causes for the larger-than-
expected xs
230
Th inventories throughout the Panama Basin, including
the effects of particle scavenging on
230
Th uxes to the seaoor.
2. Methodology
2.1. Site selection and sampling strategies
We chose sites to test whether downslope transport from
surrounding-ridge and within-basin topographic highs can explain
the higher inventories of sedimentary xs
230
Th in the Panama Basin as
suggested by Kienast et al. (2007). Cores were retrieved from the
Carnegie and Cocos Ridges that ranged in depth from 712 m to
2230 m (Fig. 1;Table 1). We also selected two deeper cores (TR 163
22, just west of the Galapagos platform, and Y69-106P, just south of
the Cocos Ridge) to add to the literature data collected from sites in
the central basin or from the foot of the Carnegie Ridge. Our
philosophy in approaching the sampling of cores here differs from
that of previous studies in that we have sampled intervals at a lower
resolution in order to obtain a broader spatial sampling of the
sedimentary inventory throughout the Panama Basin (a total of 9
cores have been sampled). The consistency between average
sediment focusing factors calculated here and those calculated by
Kienast et al. (2007) at nearby sites (see Fig. 1) shows that we are
justied in our sampling methodology.
Selected cores, in addition to those studied previously (Kienast et
al., 2007; Loubere et al., 2004) at depths greater than 2300 m, provide
for a more complete assessment of the xs
230
Th inventory in the
Panama Basin. Cores in this study were obtained from the core
repositories of Lamont-Doherty Earth Observatory, Oregon State
University and University of Rhode Island. Six or seven sediment
intervals spanning the past 25 ka were sampled from each core.
Sample selection for the Holocene and the last glacial was based on
published age models (Benway et al., 2006; Kienast et al., 2007;
Koutavas and Lynch-Stieglitz, 2003; Lea et al., 2006; Martinez et al.,
2003; Pisias and Mix, 1997)(Table 1). In addition to age models for
each core, we also need information on the dry bulk density (DBD) in
order to calculate average MARs and focusing factors. DBDs for cores
2A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
V19-27 and Y69-106P were estimated based on the CaCO
3
content in
Lyle et al. (2002). For cores RC8-102 and MV0005A-27JC, DBDs were
estimated from CaCO
3
concentrations in Ruddiman (1992) and
Kienast et al. (2007) using the equation in Snoeckx and Rea (1994)
(DBD (g cm
3
)=1/(3.60.0279×% CaCO3)). DBD for core TR163-22
was estimated in Lea et al. (2006). For cores TR163-11, TR163-33 and
ME0005A-43JC, average DBDs are estimated using carbonate content
of nearby cores ODP 1241, ME0005A-27JC and ODP 1242, respectively
(Mix et al., 2003 [ODP Leg 202 Scientic Results]). For core TR163-38
we averaged the DBD of three nearby cores (TR163-33, V19-27 and
ME0005A-27JC) with similar sedimentary lithologies and histories.
The average DBD is the least constrained for core V21-29, for which
we assumed an average value of 0.6 g cm
3
. In order to compare our
dataset with that of Kienast et al. (2007) and Loubere et al. (2004),we
averaged our xs
230
Th-derived MARs and age-model-derived MARs
similarly for two time slices, one for a specied interval covering the
entire Holocene (013 ka), and one for a specied interval in the last
glacial (1325 ka). We used broad time intervals in order to reduce
errors in sedimentation rates caused by errors from assigned ages.
2.2. Radionuclide isotope measurement
Radionuclide measurements followed the procedures described
in Pourmand et al. (2004). Approximately 0.30.5 g of dried and
homogenized sediment was spiked with known amounts of
236
U
and
229
Th (spikes used for isotope dilution analysis of uranium and
thorium isotopes). Sample and spike mixtures were digested using
HCl, HNO
3
,HClO
4
and HF acids. After complete sediment digestion
Fig. 1. Map showing location of studied cores in the Panama Basin. Red circles represent cores analyzed here and yellow circles represent cores studied previously (Kienast et al.,
2007; Loubere et al., 2004). Focusing factors are bracketed next to each core identication (rst number in bracket represents Holocene (013 ka) focusing factor and second number
represents glacial (1325 ka) focusing factor. This map has been generated using GeoMapApp software available at http://www.geomapapp.org.
Table 1
230
Th-derived Mass Accumulation Rates (MARs), age-model-derived MARs, and focusing factors in Panama Basin sediments during the Holocene and glacial. Equatorial cores
(within ±2°N and S) are shown in bold letters (other cores are non-equatorial). Margin cores are shown in light shade of gray (all other cores are non-margin cores), and deep
cores are located at water depths greater than 2300 m (see text for discussion). Subdivision of these cores into equatorial and non-equatorial, margin and non-margin, and, shallow and
deep cores are mutually inclusive. All MAR values have a common unit (g cm
2
ka
1
). Data from Kienast et al. (2007) are italicized.
3A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
Th and U were separated and puried from the digested solution
using ion exchange chromatography. Uranium and thorium isotope
ratios were measured on an Element XR magnetic sector ICP-MS at
Texas A&M University (Table 2). Th and U were analyzed separately
to avoid isotopic interferences during mass spectrometer measure-
ments.
238
U abundance sensitivity at lower masses (
234
U,
235
Uand
236
U) was corrected assuming an exponential decrease in
238
U
counts toward lower masses. In most cases the
238
Uabundance
sensitivity was about 1.5 ppm at 3 amu and 0.6 ppm at 4 amu.
Similarly, the
232
Th abundance was approximately 3.2 ppm at 2 amu
and 1.3 ppm at 3 amu. Mass bias was determined by measuring
the
238
U/
235
U in each sample in the case of uranium analyses, and by
bracketing thorium analyses with measurements of the
238
U/
235
U
in the U500 standard and assuming similar fractionation between
uranium and thorium. Mass bias corrections ranged from about 0.1
to 0.3/amu.
Table 2
Uranium, thorium results for Panama Basin cores (see Fig. 1 for location).
Cores Depth
(cm)
Years
(a)
238
U
(dpm/g)
xs
230
Th
o
(dpm/g)
232
Th
(dpm/g)
V19-27 Water depth =1373 m Latitude=0.46 Longitude = 82
35.5 4500 3.66 1.65 1.08
45.5 5500 1.58 1.17 1.73
64.5 10091 1.48 1.98 0.59
84.5 15218 2.23 1.94 0.23
94.5 17389 2.65 1.94 0.20
114.5 22000 3.82 2.19 0.22
RC8-102 Water depth =2180 m Latitude=1.4 Longitude = 86.8
2 5600 0.73 4.22 0.10
15 9050 5.14 3.85 0.10
45 15500 4.65 2.04 0.05
60 17850 2.53 3.19 0.09
75 19800 2.73 3.30 0.09
105 23800 5.93 2.43 0.09
TR163-11 Water depth= 1950 m Latitude= 6.4 Longitude = 85.8
11 4900 1.83 5.11 0.28
16 9000 2.59 5.21 0.31
22 13200 3.25 4.67 0.31
28 16200 4.92 5.25 0.28
33 18830 5.00 4.62 0.27
39 25340 3.93 4.85 0.26
V21-29 Water depth =720 m Latitude=1.0 Longitude = 89.3
19 3933 2.17 1.77 0.04
57 9394 3.20 1.80 0.05
76 12269 3.44 1.71 0.05
95 14856 7.06 1.51 0.05
114 16146 4.96 1.52 0.04
133 16808 4.98 2.02 0.05
ME0005A-43JC Water depth= 1368 m Latitude= 7.8 Longitude = 83.6
3 857 4.52 1.92 0.40
43 6808 5.01 1.82 0.38
83 10710 4.53 1.48 0.32
123 14726 5.57 1.62 0.37
163 18600 4.35 1.80 0.37
206 21920 3.91 1.37 0.32
TR163-22 Water depth= 2830 m Latitude= 0.5 Longitude = 92.4
37 4000 2.23 6.65 0.09
75 10600 3.69 5.29 0.09
112 14500 3.72 5.10 0.11
150 18200 5.01 5.55 0.15
187 21800 4.58 5.29 0.20
225 25300 3.77 5.27 0.17
Y69-106P Water depth= 2870 m Latitude= 2.9 Longitude =86.5
2 4456 1.01 6.52 0.17
11 7861 2.89 5.73 0.19
20 11876 3.62 6.24 0.26
35 20783 2.51 4.57 0.20
44 26127.2 3.15 6.14 0.23
TR163-38 Water depth =2200 m Latitude=1.3 Longitude = 81.5
19 4560 2.76 2.07 0.68
38 8510 3.39 1.82 0.90
57 10910 6.43 2.48 0.50
76 13730 6.60 2.11 0.48
95 16500 7.32 1.91 0.59
114 19000 7.56 1.85 0.47
133 22750 8.74 2.25 0.56
TR163-33 Water depth =2230 m Latitude=1.9 Longitude = 82.5
30 5580 3.78 3.48 0.38
45 8410 3.47 3.13 0.29
60 10900 3.22 2.69 0.25
75 19700 3.87 2.72 0.30
90 21200 4.13 2.67 0.33
4A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
In order to determine the true unsupported sedimentary
230
Th
xs
,
measured concentrations of
230
Th (Table 2) were corrected for
detrital
230
Th and in situ growth of
230
Th from authigenic
234
U
using Eq. 2.
230Thxs
= 230Thmeas
233U
232Th
!
det
×232Thmeas
"#
"238Umeas
233U
232Th
!
det
×232Thmeas
()
×(1eλ230t

+λ230
λ230λ234
×eλ234t
eλ230t

×
234U
233U
!
sw
1
!)#
ð2Þ
The subscripts xs, meas, det and sw refer to excess (unsupported),
measured, detrital and seawater, respectively. λ
230
and λ
234
are the
decay constants for
230
Th and
234
U, respectively. For the detrital
238
U/
232
Th activity ratio, we assumed a ratio of 0.7, which is the best
estimate for the U/Th activity ratio of detrital material delivered to the
Pacic Ocean (Henderson and Anderson, 2003). All calculations used
a seawater
234
U/
238
U activity of 1.146 (Robinson et al., 2004). In all
cases, except for Holocene intervals in cores closest to the margin (see
our denition of margin coresbelow), the detrital
230
Th amounted
to less than 5% of the total measured
230
Th. Only for one of the
Holocene samples of one margin core was the correction signicant
(~50% in sample from 45 to 46 cm in core V19-27). Authigenically-
produced
230
Th made up between 20 and 30% of the measured
230
Th
in all samples, similar to the amount found in sediments studied by
Kienast et al. (2007). Finally, xs
230
Th activities were corrected for
decay since time of sedimentary deposition.
2.3. Replicates and blanks
One quadruplicate and three replicates were run (Table 3). The
average external reproducibility was 12.4% for total uranium
concentration, 2.6% for
230
Th concentration and 2.1% for
232
Th
concentration. The poor average reproducibility for uranium was
due to the uranium reproducibility of one sediment sample from core
V21-29. This one sample had an average external reproducibility of
42.8%. We are uncertain as to why the uranium reproducibility of this
one sample is so poor. Not including the uranium reproducibility
results for core V21-29 leads to an average uranium reproducibility of
about 2.2%, similar to the thorium reproducibility. For samples that
have been replicated, the replicate averages are displayed in Table 3.
Our thorium and uranium blanks consistently make up less than 1% of
analyte, and therefore no blank corrections were necessary.
2.4. Core chronologies
Focusing factors are the ratio of the inventory of sedimentary
230
Th
xs
averaged over some depth interval to its production in the
overlying water column. The greatest uncertainty in the focusing
factor is caused by inaccuracies in the age model (Francois et al., 2004;
Kienast et al., 2007). We tried to avoid age model problems in our
calculations by choosing well-constrained, best available age models
for our cores, and to average the data over longer time spans.
Age models for cores V19-27, RC8-120 and V21-29 are based on
oxygen isotope records of planktonic foraminfera (G. sacculifer and G.
ruber) and 10 planktonic radiocarbon dates (N. dutertrei:Koutavas
and Lynch-Stieglitz, 2003). Age models for cores TR163-38 and
TR163-33 are based on high-resolution (~0.5 to 1 ka) planktonic δ
18
O
records and 3 radiocarbon dates on N. dutertrei, while the age model
for TR163-11 is based on high-resolution (~0.5 to 1 ka) planktonic
δ
18
O(Martinez et al., 2003). The age model for core ME0005A-43JC is
based on combination of planktonic δ
18
O stratigraphy as well as six
radiocarbon dates on N. dutertrei (Benway et al., 2006). Core
chronology of TR163-22 is based on δ
18
OinG. ruber and nine
radiocarbon dates (Lea et al., 2006)onN. dutertrei. The chronology of
Y69-106 is based on planktonic δ
18
O stratigraphy (Pisias and Mix,
1997). Hence, of the nine cores analyzed here, all have oxygen isotope
stratigraphy, and seven have ages calibrated by radiocarbon dating.
Errors introduced into our focusing factor calculations because of
chronological uncertainty are less than 30%. A 30% misestimate of the
Holocene MAR, for example, would require that we misplace the MIS
2/1 boundary by ~ 4 kyr.
Table 3
Uranium, thorium isotopes reproducibility results (STDev represents one standard
deviation from the mean, %rsd represents the relative standard deviation in %).
Core ID Depth (cm) [
235
U]
(dpm/g)
[
238
U]
(dpm/g)
[
230
Th]
(dpm/g)
[
232
Th]
(dpm/g)
V21-29 76 0.21 4.52 2.16 0.04
76 0.29 6.28 2.22 0.05
76 0.43 9.40 2.27 0.05
76 0.17 3.62 2.00 0.05
Average= 0.27 5.95 2.16 0.05
STDev= 0.12 2.55 0.12 0.001
%rsd= 42.84 42.84 5.51 2.80
TR163-33 30 0.17 3.78 3.77 0.38
30 0.19 4.04 3.93 0.39
30 0.18 3.86 3.83 0.40
Average= 0.18 3.89 3.84 0.39
STDev= 0.01 0.13 0.08 0.01
%rsd= 3.41 3.43 2.20 2.63
ME0005A-43JC 3 0.19 4.16 No data No data
3 0.18 4.00 2.17 0.39
3 0.19 4.09 2.12 0.38
Average= 0.19 4.08 2.14 0.38
STDev= 0.004 0.08 0.03 0.01
%rsd= 1.91 1.91 1.41 1.72
ME0005A-43JC 123 0.25 5.53 2.37 0.37
123 0.25 5.41 2.37 0.38
123 0.26 5.55 2.42 0.37
Average= 0.25 5.50 2.38 0.38
STDev= 0.004 0.08 0.03 0.005
%rsd= 1.42 1.42 1.31 1.27
Table 4
Spatio-temporal variability of
232
Th ux in the Panama Basin.
232
Th ux data for the rst
ve cores are from a previous study (Kienast et al., 2007), while the remaining data are
from this study. Bold letters represent cores that are close to continent (see Results and
discussion).
Core ID Approximate distance
from continent
(km)
232
Th ux
(dpm/m
2
/a)
Holocene Glacial Glacial/holocene
ME0005-24JC 600 1.9 2.6 1.3
Y69-71 600 1.8 2.5 1.4
ME0005-27JC 155 6.3 5.4 0.9
TR163-19 1250 0.8 1.2 1.4
TR163-31 250 7.2 8.9 1.2
V19-27 100 29.2 3.9 0.1
RC8-102 650 1.4 1.6 1.2
TR163-11 400 2.9 3.0 1.0
V21-29 900 0.5 0.5 1.1
ME0005A-43JC 100 7.7 8.0 1.0
TR163-22 1300 1.1 2.2 1.9
Y69-106P 850 2.6 3.0 1.2
TR163-38 40 19.8 14.9 0.8
TR163-33 140 5.8 6.8 1.2
5A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
3. Results
Our MAR and xs
230
Th results in addition to those studied
previously (Kienast et al., 2007) are presented (Table 1) with respect
to their temporal and spatial (latitudinal, bathymetric and distance
from continental margin) variability. In order to investigate temporal
variability of MARs and sediment focusing factors, and to be
consistent with the study of Kienast et al. (2007), we average such
parameters for intervals of sediment deposited during the Holocene
(013 ka; the Holocene) and the last glacial (1325 ka; glacial). To
describe the latitudinal spatial variability of xs
230
Th-derived MARs
and focusing factors, we have divided all cores (Table 1)into
equatorial cores (nine out of fteen cores that are located within
±2° of equator) and non-equatorial cores (six out of fteen cores that
are located outside ±2° of equator). Our subdivision of latitudinal
spatial variability into equatorial and non-equatorial cores is based on
the fact that most of the upwelling driven productivity, which may
have an effect on the scavenging efciency of
230
Th, is taking place
close to the equator (Broecker, 2008; Thomas et al., 2000). To describe
the bathymetric spatial variability, which in effect is a way to test the
downslope transport mechanism to explain higher inventory of
xs
230
Th in the deeper parts of the basin, we have chosen a bathymetric
division of 2300 m. We use the terms shallowfor cores collected at
depths b2300 m (Table 1) and deepfor those collected at depths
N2300 m (Table 1). The shallowest sill depth in the Panama Basin is
close to 2300 m (Lonsdale, 1977; Lonsdale and Malfait, 1974), and is
located at a saddle in the Carnegie Ridge. This location may serve as
one entryway through which deep water enters the Panama Basin
from the Peru Basin (Tsuchiya and Talley, 1998)the other being the
Peru trench, at a similar depth, in the eastern part of the basin. An
additional spatial variability constraint is distance from the continen-
tal margin. We have used our
232
Th ux data (discussed in Section 4.3)
to divide the studied cores into margin(within 300 km of con-
tinental margin; Table 4) and non-margincores (more than 300 km
away from continental margin; Table 4).
232
Th ux is a proxy for
continentally-derived detrital material (Anderson et al., 2006;
Marcantonio et al., 2001b; McGee et al., 2007; Pourmand et al.,
2004; Winckler et al., 2008). The division of cores into margin and
non-margin cores is based on a dramatic decrease in the detrital ux
with respect to distance from the continental margin (occurs at
approximately 300 km; see Section 4.3).
3.1. Spatio-temporal variability of mass accumulation rates (MARs)
For all nine cores analyzed here and the six cores from Kienast et al.
(2007),xs
230
Th-normalized and oxygen isotope age-model-derived
MARs were calculated and averaged for sediments deposited during the
Holocene and glacial (Table 1). For the equatorial cores, average
xs
230
Th-normalized MARs for sediment deposited during the Holocene
and glacial are 1.7 and 1.8 g cm
2
ka
1
, respectively, suggesting no
signicant temporal change in xs
230
Th-normalized MARs. In contrast,
the average oxygen isotope age-model-derived MAR for the same cores
was higher by ~50% during glacial compared to that measured during
Holocene (5.4 g cm
2
ka
1
versus 3.6 g cm
2
ka
1
). Similarly,
xs
230
Th-normalized average MARs for sediment in the non-equatorial
cores are 1.5 and 1.6 (g cm
2
ka
1
) during the Holocene and glacial,
respectively. Again,there is an insignicanttemporal change in xs
230
Th-
normalized MARs measured in non-equatorial cores for sediment
deposited during the Holocene and glacial. Average oxygen isotope age-
model-derived MARs for sediment deposited in the non-equatorial
cores are 2.4 and 3.5 (g cm
2
ka
1
) during the Holocene and glacial,
respectively. There is no difference between the Holocene and glacial
average xs
230
Th-normalized MARs for sediments analyzed in the
margin cores (each 2.1 g cm
2
ka
1
). In contrast, in the same
margin cores, the average age-model-derived MARs are higher (35%)
during the last glacial (4.2 g cm
2
ka
1
) than during the Holocene
(3.1 g cm
2
ka
1
)(Table 1). Similarly, the average xs
230
Th-normalized
MAR of the non-margin cores is 1.3 g cm
2
ka
1
during the Holocene
compared to 1.4 g cm
2
ka
1
during glacial. For the non margin
cores, average age-model-derived MARs are always higher (3.1 and
4.9 g cm
2
ka
1
) for the Holocene and glacial, respectively; (Table 1)
than MARs determined using xs
230
Th normalization.
The average xs
230
Th-normalized MARs for shallow cores are 1.8 and
1.9 (g cm
2
ka
1
) for sediments deposited during the Holocene and
glacial,respectively. Forthe same shallow cores, the averageage-model-
derived MARs are 3.0 and 4.5 g cm
2
ka
1
for the Holocene and glacial,
respectively (Table 1). For the deep cores, the average xs
230
Th-derived
MARs are 1.3 and 1.5 (g cm
2
ka
1
) during the Holocene and glacial,
respectively, versusage-model-derived MARs of 3.3 and4.7 (g/cm
2
/ka)
during the same time periods. The xs
230
Th-normalized method of
calculating MARs reveals higher MARs for the shallower cores in
comparison to those for the deeper cores, in contrast to the age-model
method, which suggests similar MARs regardless of depth.
3.2. Spatio-temporal variability of xs
230
Th-derived sediment focusing
factors
Sediment focusing factors were calculated and averaged for
sediments deposited during the Holocene and glacial. During the
last glacial, average focusing factors of all cores are 50% more than
those during the Holocene (3 versus 2; our data and data from Kienast
et al., 2007;Table 1). Focusing factors in equatorial cores display
greater variability than those in non-equatorial cores, and range from
~1 to 5 in the Holocene, and from ~1 to 11 in glacial. The percentage
increase in Holocene to glacial change of focusing factors in the
equatorial cores ranges from no change (TR163-38) to about 170%
(V21-29). Average focusing factors for equatorial cores are 2.3 and 3.6
during the Holocene and last glacial, respectively. In the non-equator
cores, the percentage increase in focusing factor from the Holocene
to glacial ranges from 12% (TR163-11) to 118% (TR163-31). Only one
non-equatorial core, Y69-106P, has a lower focusing factor (~36%)
during the last glacial (Table 1). This core resides within the Panama
Basin at the base of the Cocos Ridge at a water depth of 2870 m. Also,
average focusing factors for non-equatorial cores are 1.6 and 2.1 for
the Holocene and glacial, respectively (Table 1). Cores closest to the
equator not only have the highest focusing factors in the Panama
Basin over the past 25 ka, but also the greatest relative temporal
change in focusing factors, i.e., higher focusing factors during glacial
than in the Holocene.
For margin cores, the average focusing factors during the Holocene
and glacial are 1.6 and 2.1, respectively, so glacial focusing factors are
on average 31% greater than the Holocene. For non-margin cores
these factors are 2.3 for the Holocene and 3.5 for the glacial (Table 1).
For non-margin cores, the increase in average focusing factor during
glacial compared to the Holocene is greater (52%).
Cores studied here and by Kienast et al. (2007) have water depths
that range from 712 to 3209 m (Table 1). Nine of these fteen cores
were at or close to the tops of ridges that bound the Panama Basin
(Carnegie and Cocos Ridges), and have depths of less than 2300 m.
These shallower cores have average focusing factors during the
Holocene and glacial of 1.8 and 2.9, respectively. Focusing factors for
these shallower cores vary from 0.7 (TR163-11, depth 1950 m) to 3.9
(V21-29, depth 712 m) in the Holocene (Fig. 4), and from 0.8 (TR163-
11, depth 1950 m) to 10.5 (V21-29, depth 712 m) in the glacial. The
shallowest core, V21-29 has the highest focusing factors during both
the Holocene (4.5) and glacial (12). Similarly, the average focusing
factor of our deeper cores during the Holocene and glacial is 2.3
and 3.0, respectively. For cores Y69-106P and TR163-22, the deepest
cores studied by us here within the basin proper, focusing factors
were 1.1 and 2.3 during the Holocene, and 0.7 and 3.7 during the
glacial, respectively 2. No signicant correlation exists between
focusing factors and depth of cores (Fig. 2).
6A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
3.3. Spatio-temporal variability of
232
Th uxes
Average
232
Th uxes, estimated using the xs
230
Th-normalized
MARs are calculated for the Holocene and last glacial (Table 4). During
the Holocene and last glacial an apparently exponential decay of
detrital uxes away from the continent is observed.
232
Th uxes were
higher during glacial than those during the Holocene in all cores with
the exceptions of Carnegie Ridge cores V19-27, ME0005-27JC and
TR163-38, which are closest to the South American margin (Fig. 3).
232
Th uxes derived using oxygen isotope age models show similar
glacial-interglacial trends in margin and non-margin cores. This is
because changes in
232
Th uxes are controlled almost entirely by
changes in
232
Th concentration. In general, sites that are within about
300 km of a continental margin (see Section 4.3) have
232
Th uxes
that are up to an order of magnitude higher throughout the past 25 ka
compared to the same uxes at sites that are further than 300 km
from a continental margin.
4. Discussion
4.1. xs
230
Th from ridge tops surrounding Panama Basin
Previous
230
Th studies suggest signicant amounts of lateral
redistribution of sediments (i.e., focusing factor values N1) in the
deeper sections (27003200 m) of the Panama Basin (Kienast et al.,
2007; Kusch et al., 2010; Loubere et al., 2004). Similar xs
230
Th
inventories with focusing factor values N1 have been found
throughout the equatorial Pacic Ocean suggesting that sediment
focusing is a widespread phenomenon throughout the equatorial
sector of the western (Higgins et al., 2002), central (Marcantonio et
al., 1996, 2001a), and eastern (Broecker, 2008; Kienast et al., 2007;
Loubere et al., 2004; McGee et al., 2007) Pacic Ocean. Focusing factor
values are highest for the equatorial Pacic Ocean in the Panama Basin
where much contention over their meaning has arisen (Broecker,
2008; Francois et al., 2007; Kienast et al., 2007; Loubere et al., 2004;
Lyle et al., 2005, 2007).
Kienast et al. (2007) hypothesized that this extra sedimentary
xs
230
Th in the Panama Basin may have been derived from the top
(shallower regions) of the Carnegie Ridge through downslope
transport of sediments. Here, we further test this hypothesis by
measuring xs
230
Th in seven cores located in the topographically
highest regions (i.e., ridge tops; 7122230 m) of the Panama Basin to
test downslope transport. While there clearly is erosion over parts of
the Carnegie Ridge (Malfait and van Andel, 1980), other parts of the
Carnegie Ridge, e.g., around V19-27 are clearly depositional, as we will
discuss later.
For all but one of seven shallow cores (TR163-11), focusing factors
are greater than 1 for both the Holocene and glacial (Fig. 2). If the
underlying assumption of the xs
230
Th normalization technique is
correct (i.e., that the ux of
230
Th to the ocean oor is constant), the
ubiquitous presence (hills and basins) of inventories of xs
230
Th that
are greater than what is expected from water column production
alone means that sediment focusing is taking place everywhere in the
Panama Basin, even near the tops of ridges (Fig. 2). Moreover, the
average focusing factors in the topographically highest regions of the
basin (1.8 and 2.9 for Holocene and last glacial, respectively) are
similar to those recorded in the deepest parts of the basin (2.3 and 3.0
for Holocene and last glacial, respectively).
Our nding that focusing factors are greater than 1 on or near the
tops of ridges does not agree with the typical observation from high
resolution seismic reection studies that show basins catch more
sediment than hills, with occasional erosion from highs (Mollenhauer
et al., 2002; Tominaga et al., 2011). Two hypotheses could explain
the observation, either that (1) the sediments sampled on top of
the ridges exhibit ponding (local focusing from surrounding ridge
Fig. 3. Variation of
232
Th ux as measured by distance between core location and
continental margin for sediment deposited during A) the Holocene (013 ka) and
B) the last glacial (1325 ka). Panel C) displays the ratio of the glacial
232
Th ux to the
Holocene
232
Th ux.
232
Th ux has been plotted at the same vertical scale for the
Holocene and glacial for comparison. Long-dashed line in panel C separates cores which
are proximal to the continent from cores which are more distant.
Fig. 2. Bathymetric distribution of average focusing factors for Panama Basin sediment
deposited during the Holoc ene (013 ka) and glacial (1325 ka). Gray squares
represent data from Kienast et al. (2007) and black circles represent data obtained in
this study.
7A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
terrain), or (2) the cores receive their extra inventories of
230
Th in
another way that is not reective of horizontal sediment movement.
Several seismic reection proles of the seaoor in the Panama
Basin (available at www.geomapapp.org) clearly show some erosion-
al and nondepositional surfaces in the topographically highest regions
of the basin. For example, the seismic prole from the Vema cruise
V2104 that passed across the V21-29 site (core with the highest
focusing factor) shows that about 2/3 of the seaoor in the vicinity of
the core is bare, and ts with the idea that some horizontal advection
has taken place at the site of V21-29. Areas around the Carnegie Ridge
gap (~8586°W) are clearly erosional (Lonsdale and Malfait, 1974;
Malfait and van Andel, 1980). These regions have not yet been studied
for their
230
Th systematics but parts have been surveyed in 2010 and
will be studied in the future.
It is clear that in deeper parts of the Panama basin that the xs
230
Th
technique suggests extensive sediment focusing, and that the ux of
laterally advected sediments are 24 times greater than the ux of
that which is rained vertically through the ocean (Kienast et al., 2007).
Sedimentological evidence for signicant horizontal transport around
the Panama basin is also extensive (Lonsdale and Malfait, 1974;
Malfait and Van Andel, 1980). The earliest sedimentological mapping
of the Panama Basin is based on grain size distribution (Dowding,
1977), visual observations of erosional surfaces (Heezen and Rawson,
1977), mineralogy (Heath et al., 1974), coarse component of surface
sedimentary cover (Kowsmann, 1973), distribution of suspended
particles (Plank et al., 1973), and textural and dispersal patterns (Van
Andel, 1973). Much of this and seismic work on the ridges (Lonsdale
and Malfait, 1974; Malfait and van Andel, 1980; Van Andel, 1973)
suggest that the sedimentary cover in signicant parts of the basin is
heavily reworked by deep water currents, most notably at saddles in
ridges and to the north of saddle. The studies by Lonsdale and Malfait
(1974) and by Malfait and Van Andel (1980), in the Carnegie Gap area
of Carnegie Ridge, found areas of erosion that have a complete
absence of sediment cover, and where downslope transport probably
occurred. One can surmise that such regions of erosion or non-
deposition should contain decits of
230
Th or, at the very least, show
much smaller inventories of
230
Th than those in the deeper parts of
the basin. With more detailed studies (in progress) near where the
sediments should have been transported, these movements can be
better quantied.
Other parts of the Carnegie Ridge, around V19-27 for example, are
clearly depositional and probably have not supplied much additional
sediment to the Panama Basin (Fig. 4). On ridges especially, one must
worry about sampling bias, i.e., that piston cores tend to be taken from
basins where the sediments are thickest (Francois et al., 2007). The
area around V19-27 was surveyed for drilling on ODP Leg 202 (Site
1239; Mix et al., 2003), so good information exists about the sediment
cover (Fig. 4). About 20% of the ridge top was surveyed while trying to
locate Site 1239. Of the area surveyed about 75% was thickly covered
with sediment. All the smooth topography in Fig. 4 represents
sediment-covered terrain, for example. Even the highest portion of
the ridge just north of V19-27 has discontinuous sediment cover.
Along line 6 it is possible to estimate how variable the average
sedimentation rates have been based on the depth to the rst major
seismic horizon compared to its depth at Site 1239. Site 1239,
incidentally, has a sedimentation rate of 4.8 cm/kyr in the upper 50 m,
only slightly lower than that of V19-27 (5.2 cm/kyr). Along line 6, 60%
of the prole has a sedimentation rate between 0.5 and 1.5× that of
Site 1239, 15% of the prole has rates N1.5× that of Site 1239, and a
little less than 20% of the prole has sedimentation rates ~0.5× that of
Site 1239. Based on this and inspection of the other seismic lines, the
site at V19-27 actually has a sedimentation rate only slightly higher
than average, and is, therefore not an anomalously pondedsite in
comparison to the rest of the ridge where sedimentation appears
uniform. Lastly, the observation that focusing factors are not less than
one at V19-27 suggests that this site and the entire ridge area
surrounding it likely did not serve as a source of focused sediment in
the deeper parts of the Panama Basin.
Other site surveys were made along Cocos and Carnegie Ridge in
order to locate other shallow drillsites for ODP Leg 202 (b2300 m)
sites 1238, 1239, 1241 and 1242 (Mix et al., 2003; see UTIG marine
seismic data portal, http://www.ig.utexas.edu/sdc/cruise.php?
cruiseIn=nemo03mv). The seismic reection proles show erosional
areas but also many areas of sediment accumulation. The anks of the
ridges, e.g. around Site 1238 on the S ank of Carnegie Ridge, are
mostly depositional regions. Areas of nondeposition on the ridges are
signicantly smaller than the areas estimated in the simple box model
used to articulate the problem of ridge sources of
230
Th and sediment
(Lyle et al., 2007).
If sampling bias is not an issue, and the source of additional
sedimentary xs
230
Th inventory is not derived within the basin from
the ridge tops, then it is possible that there is an extra-basinal source
of xs
230
Th which derives from the Peru Basin located south of the
Panama Basin. Sedimentary xs
230
Th from the Peru Basin might be
laterally advected into the Panama Basin along with bottom water
through the Ecuador Trench and/or across the central saddle of the
Carnegie Ridge (Lonsdale, 1977). However, we do not have any
evidence that Peru Basin sediment is a source of sedimentary xs
230
Th.
Based on the strong temporal correspondence of alkenone, total
organic carbon and foraminifera fractions (from ne-grained to
coarse-grained) in late-glacial to Holocene sediments of the Panama
Basin, Kusch et al. (2010) argue that the source of any additional
xs
230
Th has to be transported in a syn-depositional fashion from a
local source of xs
230
Th. If the source is extra-basinal, one would expect
to see a temporal decoupling among sediment fractions with different
grain sizes since the nest grain particles that might be transported
long distances (Kusch et al., 2010) would be, presumably, older. There
is the possibility that upwelled intermediate and surface currents
might transport very ne particles from surrounding, less distal, shelf
regions toward the Panama Basin bringing additional xs
230
Th which
could explain higher focusing factors found throughout the region.
Indeed, research has shown that ner particles have higher
inventories of xs
230
Th (Kretschmer et al., 2010; McGee et al., 2010)
and, in some cases, cause an overestimation of focusing factors
Although ne-grained sediment transport from-mid-depth waters is a
possibility, there is no data to suggest that this occurs. We therefore
consider alternative explanations.
4.2. Potential for boundary scavenging effects in the Panama Basin
The main assumption of the
230
Th CFP technique is that the ux of
230
Th to the ocean oor is equal to the production of
230
Th by the decay
of
234
U in the water column. For
230
Th to be a perfect CFP, the residence
time of
230
Th should be zero, and it should be scavenged instanta-
neously as modeled in Eq. (1). Most researchers agree that there is a
limit to this assumption. The extent of lateral movement of
230
Th in the
water column is dened by its residence time in the ocean (τ
230Th
). In
turn, its residence time is dened by the extent to which thorium is
particle reactive (i.e., the degree to which Th is insoluble in seawater).
It is obvious from measurements of dissolved
230
Th in the water
column that the τ
230Th
in the ocean is not zero. General circulation
models which impart oceanic particle-ux elds (e.g., Henderson
et al., 1999; Marchal et al., 2000; Siddall et al., 2005, 2008) estimate a
τ
230Th
of about 20 yrs, suggesting that within 70% of the world's oceans
the ux of
230
Th is within about 30% of its production rate.
The margins of the Pacic Ocean, for example (Anderson et al.,
1983; 1990; Lao et al., 1992, 1993), are regions with increased particle
uxes where the ux of
230
Th is higher than its known production
rate. The increased scavenging efciency of
230
Th in such regions leads
to what is known as the boundary scavenging effect. In such cases, the
230
Th normalization technique would underestimate MARs, and
overestimate the degree to which sediment focusing takes places.
8A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
Broecker (2008) suggests that enhanced particle ux due to
enhanced upwelling and associated higher productivity along the
equator delivers larger-than-expected inventories of xs
230
Th to the
underlying sediments. This extra xs
230
Th is speculated to be supplied
by lateral mixing of equatorial waters with water adjacent to the
equator(Broecker, 2008), in essence, a boundary scavenging effect
along the equator. In contrast, a recent model paper detailing the
behavior of
230
Th in the open-ocean equatorial Pacic suggests that a
particle ux effect cannot explain the higher than expected inventories
of
230
Th in equatorial Pacic sediments (Siddall et al., 2008). However,
this result relies on the approximation of ocean mixing processes at the
equator by isopycnal and vertical diffusion terms (Siddall et al., 2008).
The exchange of deep waters through the Panama Basin is
relatively rapid, and potentially the large particulate uxes and
rapid exchange of water could interact to be an effective stripping
mechanism of
230
Th. The residence time of water within the basin is
short because of high geothermal heating (5080 yrs; Detrick et al.,
1974). Other estimates of the Panama Basin water residence time
range from 42 yrs (Mix et al., 1995 [ODP138 Scientic Results]) to less
than 50 yrs (Lonsdale, 1977). These estimates are on the order of the
residence time of thorium in sea water, suggesting the potential for a
greater ux of
230
Th to underlying sediments.
To investigate the possibility that there are enhanced removal
rates of
230
Th in regions of high particle ux within the Panama Basin,
Fig. 4. Bathymetric and seismic proles from ODP Leg 202 Site survey (ODP Leg 202 Initial Reports volume, seismic data from NEMO-3, available at the University of Texas Marine
Seismic Data Portal; http://www.ig.utexas.edu) from the top of Carnegie Ridge. It shows the position of core V19-27 studied here relative to the location of ODP Site 1239.
Bathymetric map shows locations of bare zone near rugged topography towards the north as well as sediment accumulation zones just south of ridge tops.
9A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
and given the potential for overestimating the degree of sediment
redistribution, we compared
230
Th-derived focusing factors (Fig. 1;
Table 1) between equatorial and non-equatorial regions (see Section 3
for denition of regions) using our data and data from Kienast et al.
(2007). We excluded continental margin cores that are inuenced by
high riverine inputs (Section 4.3) because we wanted to restrict our
analysis to the open ocean. There does seem to be a general
relationship between proximity to the equator (a region of high
particle ux) and higher apparent focusing factors (Fig. 1). This holds
true during both the Holocene and the last glacial, with the latter
period having proportionally higher apparent focusing factors.
Indeed, cores closest to the equator where primary production is
the greatest today have focusing factors that are about twice as high
during the last glacial than during the Holocene (average focusing
factor of 3.6 versus 2.3, respectively; Table 1;Fig. 1). This nding is
corroborated by a recent latitudinal transect study in the eastern
equatorial Pacic, west of Panama Basin at 110°W, that show
latitudinal distribution in focusing factor where the highest focusing
factor was recorded near the equator (McGee et al., 2007). For the
non-equatorial cores, the average xs
230
Th-derived focusing factors are
smaller and more similar (~1.6 versus 2.1) during both the Holocene
and last glacial, respectively (Table 1;Fig. 1). If we are to take the
model of Siddall et al. (2008) at face value, then
230
Th-derived uxes
can, at most, only overestimate focusing (or underestimate MARs) by
a factor of 1.3. For the equatorial cores, therefore, average focusing
factors can, at most be reduced to about 2.8 and 1.8 for the last glacial
and Holocene, respectively. Although our data point toward the
possibility that enhanced focusing factors nearest the equator may be
a consequence of increased productivity and enhanced scavenging
efciency of
230
Th (greater than that predicted by the model of Siddall
et al., 2008), additional sediment data is required to further evaluate
this possibility. There is the additional caveat that surface reactions
with different particle types, such as carbonates (Chase et al., 2002)
and hydrothermal manganese (Frank et al., 1994), may also have an
effect on the scavenging efciency of xs
230
Th. Hence, it may be
possible that current models underestimate the
230
Th-particle ux (or
scavenging efciency) effect, which may be greater than 30%. The
pertinent question is: is it possible to obtain better estimates of the
scavenging effect so that we can quantitatively unravel it from
calculated focusing factors based on
230
Th systematics? Additional
water column
230
Th data is required to rigorously measure the
particle-ux effect on
230
Th. Specically the extent to which a
latitudinal diffusion gradient in dissolved
230
Th concentration exists,
with lowest dissolved concentrations at the equator where produc-
tivity is the greatest, needs to be investigated.
4.3.
232
Th uxes in the Panama Basin
232
Th uxes are used as a proxy for the ux of detrital continental
material, and have been useful in deciphering past dust dynamics and
changes in wind patterns/strength associated with climate change
(Anderson et al., 2006; Marcantonio et al., 2001b; McGee et al., 2007;
Pourmand et al., 2004; Winckler et al., 2008). Here, we use
230
Th-
derived
232
Th uxes in order to easily compare our detrital uxes with
other studies. More importantly (as we note in Section 3.3), although
absolute detrital uxes calculated using age-model-derived MARs are
not identical with those calculated using
230
Th-derived MARs, relative
differences from core to core remain the same. Dust ux analyses
based on
232
Th data suggest that there were two-fold increases in
eolian uxes during glacial in the central equatorial (Anderson et al.,
2006) and eastern equatorial Pacic Ocean (McGee et al., 2007). In our
study, the core furthest to the west at 92.4°W (core TR 16322), the
glacial/the Holocene
232
Th ux ratio is about 2, identical to the
Glacial/Holocene ratios at both 110°W (McGee et al., 2007) and
140°W (Anderson et al., 2006). It is clear that the major part of the
detrital fractions at these last two locations is eolian. Hence, based on
the similar Glacial/Holocene
232
Th ux ratios, we interpret our detrital
signal at 92.4°W to be predominantly eolian derived. Comparing
average
232
Th ux over the past 25 ka, we nd similar values at 140°W
and 110°W (~ 0.6 dpm m
2
a
1
). At 92°W, however, the average
232
Th ux is 1.8 dpm m
2
a
1
(Table 3), or 3 times higher than that at
140°W and 110°W. East of 92°W, great increases in the detrital ux
(average ux over 25 ka as high as 17 dpm m
2
a
1
) are due to the
non-eolian detrital component, which is most likely made up of clays
transported to the margin by riverine runoff and diffused out to sea.
Although most of the measured
232
Th uxes in our non-margin
cores are higher in glacial than in the Holocene (Fig. 3;Table 3), three
out of the six margin cores closest to the margin (Table 3), including
those studied by Kienast et al. (2007), have higher
232
Th uxes during
the Holocene (Fig. 3). As one moves away from the continent the
glacial/the Holocene detrital ux ratio (Fig. 3C) increases. Margin
cores which record higher terrigenous runoff during the Holocene
have glacial/Holocene
232
Th ux ratios less than 1 (Fig. 3C), while
non-margin cores have glacial/the Holocene
232
Th ux ratios greater
than 1. We suggest that the increase in this ratio, as distance to the
margin increases, represents an increase in the detrital component
being more inuenced by eolian-relative to riverine-borne material.
Furthermore, we believe the higher detrital uxes for the cores closest
to the margin are due to increased continental runoff during the
Holocene in Central America and northern South America in
agreement with a recent study by Rincón-Martínez et al. (2010).
The signicant spatial distribution of our cores (Fig. 1), in addition to
those studied previously (Kienast et al., 2007),enablesusto
approximate the threshold distance where the detrital component
transitions from being mainly composed of river-borne material
versus being mainly composed of wind-blown material (Fig. 3).
Indeed, this threshold distance may be coincident with the location at
which the glacial/Holocene detrital ux ratio is approximately equal
to 1, i.e., at about 300 km (Fig. 3).
5. Summary and conclusions
In the Panama Basin, xs
230
Th-derived MARs are lower than age-
model derived MARs, and lead to the prediction that signicant
sediment focusing (i.e., lateral redistribution of sediments by deep-
sea currents) occurs. Downslope transport from surrounding ridge
tops has been proposed as a source for excess inventory of xs
230
Th
found in the deepest parts of the basin. We have tested this hypothesis
and nd a ubiquitous presence of larger-than-expected inventories of
xs
230
Th on the tops and anks of ridges that surround the Panama
Basin. Focusing factors in these regions are as high as those in the
deeper parts of the basins suggesting the ridges and anks are not
supplying the high inventories of xs
230
Th to the deep basin.
The spatio-temporal distribution of focusing factors and MARs is
such that the highest average values are those determined for
sediment deposited during the last glacial in the equatorial cores.
Lowest sediment focusing factors (still greater than 1, for the most
part) are determined for the non-equatorial cores during the
Holocene. Higher equatorial focusing factors during the glacial could
be related to scavenging effects on
230
Th driven by higher productivity
in the Panama Basin. To determine whether this is the case, more data
is needed: specically, a complementary latitudinal transect study of
water column
230
Th between high- and low-particle ux regions, and
better control on the erosional areas. Based on
232
Th ux measure-
ments, we hypothesize that the location at which eolian (as opposed
to riverine) uxes dominate the detrital ux occurs at approximately
300 km from the margin.
Acknowledgements
Sediment samples were provided by the core repositories at
LamontDoherty Earth Observatory (supported by NSF grant OCE-07-
10 A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
51761), Oregon State University (supported by NSF grant OCE-
0648164), and University of Rhode Island (supported by NSF grant
OCE-06-44625). This research is funded by NSF grant OCE-0851056 to
FM and ML. We thank Ken and Jane Williams for their generous
support of the radiogenic geochemistry isotope facility at Texas A&M
University. We thank Mark Siddall and two anonymous reviewers for
constructive comments.
References
Anderson, R.F., Bacon, M.P., Brewer, P.G., 1983. Removal of Th-230 and Pa-231 from the
Open Ocean. Earth and Planetary Science Letters 62 (1), 723.
Anderson, R.F., Lao, Y., Broecker, W.S., Trumbore, S.E., Hofmann, H.J., Wol, W., 1990.
Boundary scavenging in the Pacic-Ocean a comparison of Be-10 and Pa-231.
Earth and Planetary Science Letters 96 (34), 287304.
Anderson, R.F., Fleisher, M.Q., Lao, Y., 2006. Glacial-interglacial variability in the
delivery of dust to the central equatorial Pacic Ocean. Earth and Planetary Science
Letters 242 (34), 406414.
Bacon, M.P., 1984. Glacial to interglacial changes in carbonate and clay sedimentation in
the Atlantic-Ocean estimated from Th-230 measurements. Isotope Geoscience 2
(2), 97111.
Benway, H.M., Mix, A.C., Haley, B.A., Klinkhammer, G.P., 2006. Eastern Pacic warm pool
paleosalinity and climate variability: 030 kyr. Paleoceanography 21, PA3008.
doi:10.1029/2005PA001208.
Broecker, W., 2008. Excess sediment
230
Th: transport along the sea oor or enhanced
water column scavenging? Global Biogeochem. Cycles 22, GB1006. doi:10.1029/
2007 GB003057.
Chase, Z., Anderson, R.F., Fleisher, M.Q., Kubik, P.W., 2002. The inuence of particle
composition and particle ux on scavenging of Th, Pa and Be in the ocean. Earth and
Planetary Science Letters 204, 215229.
Detrick, R.S., Williams, D.L., Mudie, J.D., Sclater, J.G., 1974. Galapagos spreading center
bottom-water temperatures and signicance of geothermal heating. Geophysical
Journal of the Royal Astronomical Society 38 (3), 627637.
Dowding, L.G., 1977. Sediment dispersal within Cocos Gap, Panama Basin. Journal of
Sedimentary Petrology 47 (3), 11321156.
Francois, R., Frank, M., Rutgers van der Loeff, M.M., Bacon, M.P., 2004. 230Th
normalization: an essential tool for interpreting sedimentary uxes during the
late Quaternary. Paleoceanography 19, PA1018. doi:10.1029/2003PA000939.
Francois, R., et al., 2007. Comment on Do geochemical estimates of sediment focusing
pass the sediment test in the equatorial Pacic?by M. Lyle et al. (2005).
Paleoceanography 22, PA1216. doi:10.1029/2005PA001235.
Frank, M., Eckhardt, J.-D., Eisenhauer, A., Kubik, P.W., Dittrich-Hannen, B., Segl, M.,
Mangini, A., 1994. Beryllium 10, thorium 230, and protactinium 231 in Galapagos
Microplate sediments: implications of hydrothermal activity and paleoproductivity
changes during the last 100,000 years. Paleoceanography 9, 559578.
Heath, G.R., Moore, T.C., Roberts, G.L., 1974. Mineralogy of surface sediments from
Panama-Basin, Eastern-Equatorial Pacic. Journal of Geology 82 (2), 145160.
Heezen, B.C., Rawson, M., 1977. Visual observations of contemporary current erosion
and tectonic deformation on Cocos Ridge Crest. Marine Geology 23 (12), 173196.
Henderson, G.M., Anderson, R.F., 2003. The U-series toolbox for paleoceanography.
Reviews of Mineralology and Geochemistry 52, 493531.
Henderson, G.M., Heinze, C., Anderson, R.F., Winguth, A.M.E., 1999. Global distribution
of the Th-230 ux to ocean sediments constrained by GCM modelling. Deep-Sea
Research Part I-Oceanographic Research Papers 46 (11), 18611893.
Higgins, S.M., Broecker, W., Anderson, R., McCorkle, D.C., Timothy, D., 1999. Enhanced
sedimentation along the equator in the western Pacic. Geophysical Research
Letters 26 (23), 34893492.
Higgins, S.M., Anderson, R.F., Marcantonio, F., Schlosser, P., Stute, M., 2002. Sediment
focusing creates 100-ka cycles in interplanetary dust accumulation on the Ontong
Java Plateau. Earth and Planetary Science Letters 203 (1), 383397.
Kienast, S.S., Kienast, M., Mix, A.C., Calvert, S.E., Francois, R., 2007. Thorium-230
normalized particle ux and sediment focusing in the Panama Basin region during
last 30,000. Paleoceanography. doi:10.1029/2006PA001357.
Koutavas, A., Lynch-Stieglitz, J., 2003. Glacial-interglacial dynamics of the eastern
equatorial PaciccoldtongueIntertropical Convergence Zone system reconstructed
from oxygen isotope records. Paleoceanography 18. doi:10.1029/2003PA000894.
Koutavas, A., Sachs, J.P., 2008. Northern timing of deglaciation in the eastern equatorial
Pacic from alkenone paleothermometry. Paleoceanography 23, PA4205.
doi:10.1029/2008PA001593.
Koutavas, A., Lynch-Stieglitz, J., Marchitto, T.M., Sachs, J.P., 2002. El Nino-like pattern in
ice age tropical Pacic sea surface temperature. Science 297 (5579), 226230.
Kowsmann, R.O., 1973. Coarse components in surface sediments of Panama-Basin,
Eastern Equatorial Pacic. Journal of Geology 81 (4), 473494.
Kretschmer, S., Geibert, W., Rutgers van der Loeff, M.M., Mollenhauer, G., 2010. Grain
size effects on
230
Thxs inventories in opal-rich and carbonate-rich marine
sediments. Earth and Planetary Science Letters 294, 131142.
Kusch, S., Eglinton, T.I., Mix, A.C., Mollenhauer, G., 2010. Timescales of lateral sediment
transport in the Panama Basin as revealed by radiocarbon ages of alkenones, total
organic carbon and foraminifera. Earth and Planetary Science Letters 290 (34),
340350. doi:10.1016/j.epsl.2009.12.030.
Lao, Y., Anderson, Broecker, W.S., Trumbore, S.E., Hofmann, H.J., Woli, W., 1992.
Transport and burial rates of Be-10 and Pa-231 in the Pacic-Ocean during the
Holocene Period. Earth and Planetary Science Letters 113 (12), 173189.
Lao, Y., Anderson, R.F., Broecker, W.S., Hofmann, H.J., Woli, W., 1993. Particulate uxes
of Th-230, Pa-231, and Be-10 in the northeastern Pacic-Ocean. Geochimica Et
Cosmochimica Acta 57 (1), 205217.
Lea, D.W., Pak, D.K., Belanger, C.L., Spero, H.J., Hall, M.A., Shackleton, N.J., 2006.
Paleoclimate history of Galapagos surf ace waters over the last 135,000 yr.
Quaternary Science Reviews 25 (1112), 11521167.
Lonsdale, P., 1977. Inow of bottom water to Panama Basin. Deep-Sea Research 24 (12),
10651101.
Lonsdale, P., Malfait, B., 1974. Abyssal dunes of foraminiferal sand on Carnegie Ridge.
Geological Society of America Bulletin 85, 16971712.
Loubere, P., Mekik, F., Francois, R., Pichat, S., 2004. Export uxes of calcite in the eastern
equatorial Pacic from the Last Glacial Maximum to the present. Paleoceanography
19, PA2018. doi:10.1029/2003PA000986.
Lyle, M., 1988. Climatically forced organic-carbon burial in equatorial Atlantic and
Pacic Oceans. Nature 335 (6190), 529532.
Lyle, M., Mix, A., Pisias, N., 2002. Patterns of CaCO3 deposition in the eastern tropical
Pacic Ocean for the last 150 kyr: evidence for a southeast Pacic depositional spike
during marine isotope stage (MIS) 2. Paleoceanography 17 (2), 1013. doi:10.1029/
2000PA000538.
Lyle, M., Mitchell, N., Pisias, N., Mix, A., Martinez, J.I., Paytan, A., 2005. Do geochemical
estimates of sediment focusing pass the sediment test in the equatorial Pacic?
Paleoceanography 20, PA1005. doi:10.1029/2004PA001019.
Lyle, M., Pisias, N., Paytan, A., Martinez, J.I., Mix, A., 2007. Reply to comment by R.
Francois et al. on Do geochemical estimates of sediment focusing pass the
sediment test in the equatorial Pacic?: further explorations of 230Th normal-
ization. Paleoceanography 22, PA1217. doi:10.1029/2006PA001373.
Malfait, B.T., Van Andel, T.H., 1980. A modern oceanic hardground on the Carnegie
Ridge in the eastern Equatorial Pacic. Sedimentology 27, 467496.
Marcantonio, F., Anderson, R.R., Stute, M., Kumar, N., Schlosser, P., Mix, A.C., 1996.
Extraterrestrial 3He as a tracer of marine sediment transport and accumulation.
Nature 383, 705707.
Marcantonio, F., Anderson, R.F., Higgins, S., Stute, M., Schlosser, P., Kubik, P.W., 2001a.
Sediment focusing in the central equatorial Pacic ocean. Paleoceanography 16, 260267.
Marcantonio, F., Anderson, R.F., Higgins, S., Fleisher, M.Q., Stute, M., Schlosser, P., 2001b.
Abrupt intensication of the SW Indian Ocean monsoon during the last
deglaciation: constraints from Th, Pa, and He isotopes. Earth and Planetary Science
Letters 184, 505514.
Marchal, O., Francois, R., Stocker, T.F., Joos, F., 2000. Ocean thermohaline circulation and
sedimentary Pa-231/Th-230 ratio. Paleoceanography 15 (6), 625641.
Martinez, I., Keigwin, L., Barrows, T.T., Yokoyama, Y., Southon, J., 2003. La Nina-like
conditions in the eastern equatorial Pacic and a stronger Choco jet in the northern
Andes during the last glaciation. Paleoceanography 18 (2), 1033.
McGee, D., Marcantonio, F., Lynch-Stieglitz, J., 2007. Deglacial changes in dust ux in the
eastern equatorial Pacic. Earth and Planetary Science Letters 257 (12), 215230.
McGee, D., Marcantonio, F., McManus, J.F., Winckler, G., 2010. The response of excess
230
Th and extraterrestrial
3
He to sediment redistribution at the Blake Ridge,
western North Atlantic (2010). Earth and Planetary Science Letters 299 (12),
138149. doi:10.1016/j.epsl.2010.08.029.
Mix, A.C., Pisias, N.G., Rugh, W., Wilson, J., Morey, A., Hagelberg, T., 1995. Benthic
foraminiferal stable isotope record from Site 849, 05 Ma: Local and global climate
changes. In: Pisias, N.G., Mayer, L., Janecek, T., Palmer-Julson, A., van Andel, T.H.
(Eds.), Proc. ODP,/Scientic Res ults/138, College Station, TX (Ocean Drilling
Program), pp. 371412.
Mix, A.C., Tiedemann, R., Blum, P., et al., 2003. Proc. ODP, Init. Repts., 202: College Station,
TX (Ocean Drilling Program). doi:10.2973/odp.proc.ir.202.2003.
Mollenhauer, G., Schneider, R.R., Muller, P.J., Spiess, V., Wefer, G., 2002. Glacial/
interglacial variability in the Benguela upwelling system: spatial distribution and
budgets of organic carbon accumulation, Global Biogeochem. Cycles 16 (4), 1134.
doi:10.1029/2001GB001488.
Paytan, A., Kastner, M., Chavez, F.P., 1996. Glacial to interglacial uctuations in
productivity in the equatorial Pacic as indicated by marine barite. Science 274
(5291), 13551357.
Pedersen, T.F., 1983. Increased productivity in the eastern Equatorial Pacic during the
last glacial maximum (19,000 to 14,000 Yr Bp). Geology 11 (1), 1619.
Pisias, N.G., Mix, A.C., 1997. Spatial and temporal oceanographic variability of the
eastern equatorial Pacic during the late Pleistocene: evidence from Radiolaria
microfossils. Paleoceanography 12 (3), 381393.
Plank, W.S., Ronald, J., Pak, H., Zaneveld, V., 1973. Distribution of suspended matter in
Panama Basin. Journal of Geophysical Research 78 (30), 71137121.
Pourmand, A., Marcantonio, F., Schulz, H., 2004. Variation in productivity and eolian
uxes in the northeastern Arabian Sea during the past 110 ka. Earth and Planetary
Science Letters 221, 3954.
Rincón-Martínez, D., Lamy, F., Contreras, S., Leduc, G., Bard, E., Saukel, C., Blanz, T.,
Mackensen, A., Tiedemann, R., 2010. More humid interglacials in Ecuador during
the past 500 kyr linked to latitudinal shifts of the equatorial front and the
Intertropical Convergence Zone in the eastern tropical Pacic. Paleoceanography
25, PA2210. doi:10.1029/2009PA001868.
Robinson, L.F., Belshaw, N.S., Henderson, G.M., 2004. U and Th concentrations and
isotope ratios in modern carbonates and waters from the Bahamas. Geochimica et
Cosmochimica Acta 98, 17771789.
Ruddiman, W., 1992. Calcium carbonate database. IGBP PAGES/World Data Center for
Paleoclimatology Data Contribution Series #92-001. NOAA/NGDC Paleoclimatology
Program, Boulder, Colorado, USA.
Siddall, M., Henderson, G.M., Edwards, N.R., Muller, S.A., Stocker, T.F., Joos, F., Frank, M., 2005.
231 Pa/
230
Th fractionation by ocean transport, biogenic particle ux and particle type.
Earth and Planetary Science Letters 237, 135155. doi:10.1016/j.epsl.2005.05.031.
11A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
Siddall, M., Anderson, R.F., Winckler, G., Henderson, G.M., Bradtmiller, L.I., McGee, D.,
Franzese, A., Stocker, T.F., Muller, S.A., 2008. Modeling the particle ux effect on
distribution of 230Th in the equatorial Pacic. Paleoceanography 23, PA2208.
doi:10.1029/2007PA001556.
Snoeckx,H., Rea, D.K., 1994.Dry bulk densityand CaCO3 relationships in upperQuaternary
sediments of the eastern equatorial Pacic. Marine Geology 120, 327333.
Suman, D.O., Bacon, M.P., 1989. Variations in Holocene sedimentation in the North-
American basin determined from Th-230 measurements. Deep-Sea Research Part
a-Oceanographic Research Papers 36 (6), 869878.
Thomas, E., Turekian, K.K., Wei, K.Y., 2000. Productivity control of ne particle transport
to equatorial Pacic sediment. Global Biogeochemical Cycles 14 (3), 945955.
Tominaga, M. , Lyle, M., Mitchell, N.C., 2011. Se ismic interpretation of pelagic
sedimentation regimes in the 1853 Ma eastern equatorial Pacic: basin-scale
sedimentation and inlling of abyssal valleys. Geochemistry, Geophysics, Geosys-
tems 12 (3). doi:10.1029/2010GC003347 22 pp.
Tsuchiya, M., Talley, L.D., 1998. A Pacic hydrographic section at 88 degrees W: Water-
property distribution. Journal of Geophysical Research-Oceans 103 (C6), 1289912918.
Van Andel, T.H., 1973. Texture and dispersal of sediments in Panama-Basin. Journal of
Geology 81 (4), 434457.
Winckler, G., Anderson, R.F., Fleisher, M.Q., Mcgee, D., Mahowald, N., 2008. Covariant
glacialinterglacial dust uxes in the equatorial Pacic and Antarctica. Science 320
(5872), 9396.
12 A.K. Singh et al. / Earth and Planetary Science Letters xxx (2011) xxxxxx
Please cite this article as: Singh, A.K., et al., Sediment focusing in the Panama Basin, Eastern Equatorial Pacic Ocean, Earth Planet. Sci. Lett.
(2011), doi:10.1016/j.epsl.2011.06.020
... However, there are reports of higher velocities (20-30 cm/s) and also dune fields in some parts of the Carnegie and Malpelo Ridges. Those fluctuations cause the winnowing and focusing processes, both related to deep sea currents, which affect sedimentation rates in the basin and result in moderate sorting of the biogenic oozes (Lonsdale and Malfait, 1974;Heezen and Rawson, 1977;Lonsdale and Fornari, 1980;Singh et al., 2011;Dubois and Mitchel, 2012;Bista et al., 2016). ...
... For instance, the biogenic component is higher around the Galapagos Archipelago and the oceanic ridges, decreasing near the continental slopes. Furthermore, the non-biogenic component of the deep sea sediments comprises terrigenous material and volcanic material from the oceanic ridges (Moore et al., 1973;Lyle, 1992;Singh et al., 2011). The terrestrial input of organic matter is particularly high along the continental margin of southern Colombia and the Panama Bight. ...
... Restrepo and Kjerfve, 2000). Therefore, higher sedimentation rates (>4 cm/ky) are present in the eastern part of the basin, close to the continental slope (Kienast et al., 2007;Singh et al., 2011). ...
Article
Environmental controls affecting the composition and distribution of deep sea benthic foraminifera were analyzed in the Panama Basin (Eastern Equatorial Pacific). For this study, results of the analysis of a new set of 33 core-top samples from the upper-lower slope (714–3819 mbsf) were integrated with previous data from this area. The composition of the foraminiferal assemblages resembles a typical deep sea association, illustrating a close relationship with dissolved oxygen at the seafloor and indirectly with surface productivity. Low oxygen content and organic matter influx, as inferred in the eastern Panama Basin, are connected to terrigenous supply and seasonal upwelling on the Panama Bight and southern Costa Rica. Buliminids, bolivinids, and Epistominella spp. are common elements of the continental slope. In addition, physiographic aspects such as the type of substrate, turbidity fluxes, and bottom currents also exert control over some of the described taxa and defined morphogroups. A broad exploration of the infaunal/epifaunal relationships in the basin does not suggest a clear relationship with dissolved oxygen content. However, there are fluctuations in the proportion of some of the defined epifaunal morphogroups and selected taxa (e.g. Cibicidoides wuellerstorfi, Chilostomella oolina, and Hoeglundina elegans, among others) along the Carnegie Ridge. These changes are explained by the variable physiographic features (surface upwelling, strong bottom currents) that coalesce in this part of the Panama Basin. Furthermore, re-evaluation of the foraminiferal morphogroups in two deep sea cores that comprised the last 25 ky, indicate that some taxa of the recent samples (Cassidulina spp. and Siphouvigerina peregrina, among others) reacted to paleoproductivity and bottom current fluctuations at the end of the Last Glacial Maximum and the Younger Dryas. Despite the lack of ecological data in the foraminiferal assemblages for a large proportion of the Panama Basin, the application of selected taxa or morphogroups as proxies for paleoceanographic conditions is promising.
... Therefore, further work is required to discern the variability of past and present-day surface productivity in this equatorial region and how it affects the global carbon cycle and climate change (Pennington et al., 2006;Sigman & Boyle, 2000;Takahashi et al., 2002). In the EEP, paleoproductivity estimates are primarily based on evaluating fluxes of organic carbon, calcite, and opal in pelagic sediments using age-model-based (Lyle, 1988) or 230 Th-based accumulation rates (Bradtmiller et al., 2010;Kienast et al., 2007;Loveley et al., 2017;Marcantonio et al., 2014;Singh et al., 2011). Variable dissolution of carbonate and very low preservation efficiency of opal and particulate organic carbon may obscure the true paleoproductivity signals in a pelagic sediment column (Jaccard et al., 2009;Lyle et al., 2014;Treguer et al., 1995). ...
... We collected sediment samples deposited during marine oxygen isotope stages 1 and 2 (MIS 1 = 0-12 kyr and MIS 2 = 12-25 kyr; see Table 1 for information on age models) in upwelling regions south of equatorial front (i.e., sites TR163-31, TR163-33, RC8-102, and V21-29), within the equatorial divergence and cold tongue region (TR163-22, MV1014-2-9MC, and MV1014-2-16MC), continental marginal regions (within 300 km, Singh et al., 2011) within which detrital sediments are abundant (TR163-33, 163-38, V19-27, and ME0005-43JC), and less-productive regions located north of the equatorial front in warmer waters (Y69-106, TR163-11, ME0005-43JC, MV1014-1-1MC, and MV1014-1-7MC). In addition, our sample collection spans water depths which range from 712 to 2,870 m. ...
... Age models for sediment cores used in this study are well-constrained with radiocarbon ages in conjunction with oxygen isotope records of planktonic foraminifera from previous studies ( Table 1). Description of the age models and their accuracy is discussed in detail in Singh et al. (2011) and Marcantonio et al. (2014). ...
Article
Full-text available
We present the first regional‐scale records of biogenic Barium (xsBa) fluxes in the Panama basin of the eastern‐equatorial margin of the Pacific Ocean in order to assess xsBa as a paleoproductivity proxy. Measurements of xsBa from 13 cores that range in water depths from about 700 to 3,000 m show an increase in ²³⁰Th‐normalized xsBa mass accumulation rates (MARs) with increasing water depth during both marine oxygen isotope stages (MIS) 1 and 2. The correlation of xsBa MARs with depth are strong despite differences in bulk sediment MARs and differing degrees of sediment redistribution. We interpret the increasing xsBa with water depth as likely due to the continued decomposition and remineralization of falling and/or resuspended biogenic particles. xsBa does not seem to be affected by diagenetic sulfate reduction in most of the cores. Calculated estimates of xsBa preservation in the sediment pile are high and fluctuate between 45% and 52% throughout the last 25 kyr. Although xsBa fluxes can be a robust indicator of paleoproductivity, caution is needed if (a) there is evidence of sulfate reduction in sediments being analyzed, and (b) one is trying to quantify differences in paleoproductivity among sites that are located at different depths in the water column.
... Microbial respiration of organic matter lowers porewater oxygen concentrations relative to BWO concentrations, such that uranium may precipitate even though the overlying bottom waters are oxic. For example, uranium enrichment has been observed in the Eastern Equatorial Pacific (Singh et al., 2011), the California Margin (McManus et al., 1998), and the Bering Sea (Serno et al., 2014) at sites that all have BWO concentrations >90 lmol/kg. In some settings, particularly on continental margins, high organic carbon content (>2 wt% organic carbon, often >5 wt%) is the dominant control on porewater oxygen, and thus uranium precipitation is primarily interpreted as a productivity rather than oxygen signal (e.g., McManus et al., 2006). ...
... This study leverages several previously published databases. Core tops, analyzed by Wang et al. (2022), include the full spectrum of major and trace elements including U, Ba, and P. The core top Ba database of Hayes et al. (2021) was cross-referenced with the U-Th database of Costa et al. (2020) as well as the original publications in order to compile U, Ba, and P where available (Dorta and Rona, 1971;Sarin et al., 1979;Calvert and Price, 1983;Frank et al., 1995;McManus et al., 1998;Bonn et al., 1998;Ceccaroni et al., 1998;Weber, 1998;Veeh et al., 1999;Murray et al., 2000;Sirocko et al., 2000;Veeh et al., 2000;Pfeifer et al., 2001;Nameroff et al., 2002;Prakash Babu et al., 2002;Chase et al., 2003;Eagle et al., 2003;Neimann and Geibert, 2003;Anderson and Winckler, 2005;McManus et al., 2005;Gonneea and Paytan, 2006;Shigemitsu et al., 2007;Bradtmiller et al., 2007;Anderson et al., 2009;Mills et al., 2010;Singh et al., 2011;Anderson et al., 2014;Serno et al., 2014;Calvert et al., 2015;Costa et al., 2016;Hickey, 2016;Lippold et al., 2016;Loveley et al., 2017;Wang et al., 2022). Several additional sites were included for which both U and Ba were available. ...
Article
Oxygen is essential for marine ecosystems, and it is linked by respiration to carbon storage in the deep ocean. Reconstructing oxygen concentrations in the past has been limited by the absence of quantitative, rather than qualitative, proxies, but several new (semi-) quantitative oxygen proxies have recently been developed. In this study we explore the possibility of adding bulk sedimentary uranium (U) to this list by normalizing it to barium (Ba). First, U/Ba and bottom water oxygen concentrations are compared on a global scale, using a core top database, in pelagic environments greater than 200 m water depth. Then, the relationships between U/Ba and bottom water oxygen are examined on smaller spatial scales: within each ocean basin and regionally within the Eastern Equatorial Pacific, the Arabian Sea, and Western Equatorial Atlantic. At this regional scale, where secondary influences on the behavior of both U and Ba may be more spatially uniform, empirical piecewise linear calibrations are developed and subsequently tested on downcore records. U/Ba-based oxygen reconstructions generally agree with those derived from previously published alkenone preservation and benthic foraminiferal surface porosity records. Several limitations to the utility of U/Ba as a proxy for oxygen have also been identified. The proxy should only be applied in the uppermost sedimentary intervals that contain porewater sulfate to minimize barite diagenesis, and phosphorus contents should be monitored for the potential influence of apatite on uranium content. U/Ba is more successful at recording oxygen concentrations during mean glacial and interglacial periods than during climate transitions, when the timing and amplitude may be more sensitive to burndown and smoothing. Conservative errors on the calibrations result in the greatest utility of U/Ba in regions with relatively high oxygen concentrations (e.g., >50 μmol/kg) and large oxygen variability (±10s of μmol/kg). Even with these caveats, U/Ba is only one of two quantitative oxygen proxies potentially capable of recording variability above 50 μmol/kg, and further investigation into its functionality in different environmental settings is worthwhile in the endeavor to reconstruct the full marine range of oxygen concentrations in the past.
... Following Singh et al. (2011) and Kienast et al. (2016), we excluded sites located less than 300 km from the coast, except for the equatorial Atlantic Ocean off the coast of Brazil, for which this distance is 600 km (Holocene) and 800 km (LGM). This rule does not hold for studies that isolate dust from riverine contributions to terrigenous sediment (e.g., McGee et al., 2013). ...
Article
Full-text available
Mineral dust aerosol concentrations in the atmosphere varied greatly on glacial–interglacial timescales. The greatest changes in global dust activity occurred in response to changes in orbital parameters (which affect dust emission intensity through glacial activity) and the lifetime of dust in the atmosphere (caused by changes in the global hydrological cycle). Long-term changes in the surface dust deposition rate are registered in geological archives such as loess, peats, lakes, marine sediments, and ice. Data provided by these archives are crucial for guiding simulations of dust and for better understanding the natural global dust cycle. However, the methods employed to derive paleo-dust deposition rates differ markedly between archives and are subject to different sources of uncertainty. Here, we present Paleo±Dust, an updated compilation of bulk and <10 µm paleo-dust deposition rates with quantitative 1σ uncertainties that are inter-comparable among archive types. Paleo±Dust incorporates a total of 285 pre-industrial Holocene (pi-HOL) and 209 Last Glacial Maximum (LGM) dust flux constraints from studies published until December 2022, including, for the first time, peat records. We also recalculate previously published dust fluxes to exclude data from the last deglaciation and thus obtain more representative constraints for the last pre-industrial interglacial and glacial end-member climate states. Based on Paleo±Dust, the global LGM:pi-HOL ratio of <10 µm dust deposition rates is 3.1 ± 0.7 (1σ). We expect Paleo±Dust to be of use for future paleoclimate dust studies and simulations using Earth system models of high to intermediate complexity. Paleo±Dust is publicly accessible at 10.1594/PANGAEA.962969 (Cosentino et al., 2024).
... Following Singh et al. (2011) and Kienast et al. (2016), we excluded sites located less than 300 km from the coast, except for the equatorial Atlantic Ocean off the coast of Brazil, for which this distance is 600 km (Holocene) and 800 km (LGM). This 290 rule does not hold for studies that isolate dust from riverine contributions to terrigenous sediment (e.g., McGee et al., 2013). ...
Preprint
Full-text available
Mineral dust aerosol concentrations in the atmosphere varied greatly on glacial-interglacial timescales. The greatest changes in global dust activity occurred in response to changes in orbital parameters that affect dust emission intensity through glacial activity, and dust lifetime in the atmosphere through changes in the global hydrological cycle. Long-term changes in surface dust deposition rate are registered in geological archives such as loess, peats, lakes, marine sediments, and ice. Data provided by these archives is crucial for guiding simulations of dust, and for better understanding the natural global dust cycle. However, the methods employed to derive paleo-dust deposition rates differ markedly between archives and are subject to different sources of uncertainty. Here, we present Paleo±Dust, an updated compilation of bulk and <10-µm paleo-dust deposition rate with quantitative 1-σ uncertainties that are inter-comparable among archive types. Paleo±Dust incorporates a total of 284 pre-industrial Holocene (pi-HOL) and 208 Last Glacial Maximum (LGM) dust flux constraints from studies published until December 2022, including for the first time peat records. We also recalculate previously published dust fluxes to exclude data from the last deglaciation and thus obtain more representative constraints for the last pre-industrial interglacial and glacial end-member climate states. Based on Paleo±Dust, the global LGM:pi-HOL ratio of <10-µm dust deposition rate is 3.1 ± 0.8 (1σ). We expect Paleo±Dust to be of use for future paleoclimate dust studies and simulations using Earth system models of high to intermediate complexity.
... Long chain n-alkanes with an odd number of carbon atoms are found in the waxes of higher plant tissues (Eglinton & Hamilton, 1967;Eglinton & Eglinton, 2008) and are transported to the ocean by wind or rivers (Gagosian et al., 1987), so these compounds have commonly been used as a continental material and dust input proxy (Calvo et al., 2004(Calvo et al., , 2011Lamy et al., 2014;Martínez-Garcia et al., 2009. In the ODP Site 1240 location, 600 km away from the continent, eolian dust input represents the main source of land-originated particles (Singh et al., 2011). Here, we only present data from the C 29 n-alkane homolog, as it was the most abundant n-alkane (the distribution of these compounds had a strong odd to even preference, and the C 29 n-alkane showed the same trend as the total n-alkane abundance). ...
Article
Full-text available
Modern biogeochemical conditions of the Eastern Equatorial Pacific (EEP) region are characterized by high macronutrient concentrations but low phytoplankton abundance due to both iron and silicic acid limitation. Since primary producers significantly impact the global carbon cycle, paleoproductivity in relation to climate change and nutrient availability in this region has been a topic of a number of studies. However, the complex dynamics of this region, especially east of the Galapagos Islands, has led to some discrepancies when linking reconstructed paleoproductivity with potential mechanisms for higher primary productivity. Here we focus on reconstructing primary productivity of haptophyte algae and diatoms, as well as continental material input, sea surface salinity, and sea surface temperature, and compare these reconstructions with existing records for the period comprised between 150 and 110 ka (the penultimate deglaciation period) with the aim to understand the mechanisms that most significantly influence phytoplankton growth over the EEP region east of the Galapagos Islands. Our results suggest enhanced upwelling in the EEP system during the penultimate deglaciation and increased phytoplankton abundance mainly as the result of both the increasing influence of nutrient‐rich Southern Ocean sourced waters through the Equatorial Undercurrent and a higher input of iron through atmospheric deposition. The highest phytoplankton abundances recorded at the study site during the penultimate deglaciation also suggest that maximum input of nutrients might have occurred during the millennial‐scale event Heinrich Event 11 in the North Atlantic as a result of global atmospheric and oceanic reorganizations.
... Over 50 years ' (1966-2019) worth of data have been compiled to create the global thorium database (n = 1,167) presented here (Adkins et al., 2006;Anderson et al., 2006Bausch, 2018;Böhm et al., 2015;Bohrmann, 2013;Borole, 1993;Bradtmiller et al., 2006Bradtmiller et al., , 2007Bradtmiller et al., , 2009Broecker, 2008;Broecker et al., 1993;Brunelle et al., 2007Brunelle et al., , 2010Causse & Hillaire-Marcel, 1989;Chase et al., 2003Chase et al., , 2014Chong et al., 2016;Costa, McManus, & Anderson, 2017;Crusius et al., 2004;Dekov, 1994;Denis et al., 2009;Dezileau et al., 2000Dezileau et al., , 2004Durand et al., 2017;Fagel et al., 2002;Francois et al., 1990, Francois et al., 1993Frank, Eisenhauer, Bonn, et al., 1995Fukuda et al., 2013;Galbraith et al., 2007;Geibert et al., 2005;Gherardi et al., 2005Gherardi et al., , 2009Gottschalk et al., 2016;Hickey, 2010;Hillaire-Marcel et al., 2017;Hoffmann et al., 2013Hoffmann et al., , 2018Jaccard et al., 2009Jaccard et al., , 2013Jacobel et al., 2017a;Jonkers et al., 2015;Kienast et al., 2007;Ku & Broecker, 1966;Kumar et al., 1995;Lam et al., 2013;Lamy et al., 2014;Lippold et al., 2009, Lippold et al., 2011, Lippold et al., 2012Loubere et al., 2004;Loveley et al., 2017;Lund et al., 2019;Mangini & Dominik, 1978;Marcantonio et al., 1996, Marcantonio et al., 2001Martínez-Garcia et al., 2009;McGee et al., 2007, 2010, McGee & Mukhopadhyay, 2013McManus et al., 1998McManus et al., , 2004Meier, 2015;Middleton et al., 2020;Missiaen et al., 2018;Mohamed et al., 1996;Mollenhauer et al., 2011;Moran et al., 2005;Mulitza et al., 2008Muller et al., 2012;Nave et al., 2007;Negre et al., 2010;Neimann & Geibert, 2003;Ng et al., 2018;Not & Hillaire-Marcel Claude, 2010;Nuttin, 2014 Paleoceanography and Paleoclimatology Plain, 2004;Pourmand et al., 2004, Pourmand et al., 2007Purcell, 2019;Roberts et al., 2014;Robinson et al., 2008;Rowland et al., 2017;Ruhlemann et al., 1996;Sarin et al., 1979;Saukel, 2011;Scholten et al., 1990, Scholten et al., 2008Serno et al., 2014Serno et al., , 2015Shiau et al., 2012;Shimmield et al., 1986;Shimmield & Mowbray, 1991;Shimmield & Price, 1988;Singh et al., 2011;Skonieczny et al., 2019;Studer et al., 2015;Sukumaran, 1994;Thiagarajan & McManus, 2019;Thöle et al., 2019;Thomas et al., 2007;Thomson et al., 1993, Thomson et al., 1995, Thomson et al., 1999Vallieres, 1997;Veeh et al., 1999Veeh et al., , 2000Veiga-Pires & Hillaire-Marcel, 1999;Voigt et al., 2017;Waelbroeck et al., 2018;Walter et al., 1997;Wengler et al., 2019;Williams et al., 2016;Winckler et al., 2008;Yang & Elderfield, 1990;Yang et al., 1995;Yu, 1994;Zhou & McManus, 2020). ...
Article
Full-text available
Th normalization is a valuable paleoceanographic tool for reconstructing high‐resolution sediment fluxes during the late Pleistocene (last ~500,000 years). As its application has expanded to ever more diverse marine environments, the nuances of ²³⁰Th systematics, with regard to particle type, particle size, lateral advective/diffusive redistribution, and other processes, have emerged. We synthesized over 1000 sedimentary records of ²³⁰Th from across the global ocean at two time slices, the late Holocene (0–5,000 years ago, or 0–5 ka) and the Last Glacial Maximum (18.5–23.5 ka), and investigated the spatial structure of ²³⁰Th‐normalized mass fluxes. On a global scale, sedimentary mass fluxes were significantly higher during the Last Glacial Maximum (1.79–2.17 g/cm²kyr, 95% confidence) relative to the Holocene (1.48–1.68 g/cm²kyr, 95% confidence). We then examined the potential confounding influences of boundary scavenging, nepheloid layers, hydrothermal scavenging, size‐dependent sediment fractionation, and carbonate dissolution on the efficacy of ²³⁰Th as a constant flux proxy. Anomalous ²³⁰Th behavior is sometimes observed proximal to hydrothermal ridges and in continental margins where high particle fluxes and steep continental slopes can lead to the combined effects of boundary scavenging and nepheloid interference. Notwithstanding these limitations, we found that ²³⁰Th normalization is a robust tool for determining sediment mass accumulation rates in the majority of pelagic marine settings (>1,000 m water depth).
Article
Aeolian dust has a crucial impact on the marine carbon cycle, through its leverage on marine export production. To reconstruct dust depositions, 232Th in marine sediment is widely used as a proxy, with different 232Th concentrations of upper continental crust (UCC). Here, by comparing a novel compilation of 230Th-normalized 232Th in globally distributed marine sediment and the directly measured dust flux, we show that 14 ppm of 232Th concentration in UCC is the optimized parameter for deriving dust flux. This concentration was subsequently applied to recalculate the thorium-based dust flux during both the Holocene and the last glacial maximum (LGM), which yielded elevated LGM dust deposition over Holocene in most cores. A closer look at the dust comparison reveals a marked increase in the LGM dust flux at regions between 45°S to 55°S of the Southern Ocean owing to the variation in the dynamics of the Antarctic Circumpolar Current. Furthermore, with a compilation of the 230Th-normalized excess barium flux, these thorium-derived dust depositions suggest that no dust-induced iron fertilization occurred in the equatorial Pacific and the Antarctic zone of the Southern Ocean, as upwelling-derived iron was much greater than dust-derived iron, while the same case in the north Pacific Ocean was caused by the reduced vertical supply of nutrients. In contrast, dust-born iron boosted export productivity in most regions of the Subantarctic zone of the Southern Ocean.
Article
The Eastern Equatorial Pacific (EEP) affects the ocean-atmosphere exchange of CO_2 on seasonal and interannual time scales through a balance of upwelling of CO_2-rich waters and the drawdown of CO_2 by biological productivity in the surface waters. The EEP accounts for almost 3/4ths of the global oceanic outgassing of CO_2 to the atmosphere, and it is known that the size of this EEP source of CO_2 varies significantly during El Nino events (Feely et al., 1999). There has been much effort to determine the El Nino Southern Oscillation (ENSO) state of the Equatorial Pacific during the past, particularly at the Last Glacial Maximum (LGM) when the global atmospheric [CO_2] was low, yet the glacial ENSO state remains a source of considerable controversy (Ford et al., 2015; Herguera, 2000; Koutavas et al., 2002; Loubere, 2001; Loubere et al., 2004; Lyle, 1988; Paytan et al., 1996; Pedersen, 1983; Sarnthein et al., 1988). Reconstructing past changes in equatorial productivity could help establish the prevailing ENSO state of the Pacific during the LGM, as the El Nino-related deepening of the thermocline in the East Pacific reduces productivity in the EEP and increases it in the Western Equatorial Pacific. Here we investigate changes in productivity in four cores from the equatorial Pacific, in the heart of the modern equatorial cold tongue. We determine changes in productivity using measurements of ^(231)Pa, ^(230)Th, ^(232)Th, ^(235)U and ^(238)U along with sedimentary fluxes. We also compare our findings to other sediment cores in the Pacific. We find elevated (^(231)Pa/^(230)Th)_(xs_ values (higher than production values) in general across the cores, indicating a net sink for oceanic ^(231)Pa in the EEP. We also find evidence for low levels of lateral sediment focusing, as well as lower productivity during the glacial in reduced ^(230)Th-normalized opal fluxes and decreased (^(231)Pa/^(230)Th)_(xs) at multiple sites. Examination of authigenic uranium at our sites in conjunction with previous work (Jacobel et al., 2017) shows that between 2 and 3.5 km depth in the Equatorial Pacific, there was a floating pool of respired carbon associated with the southward return flow of North Pacific Deep Water, sequestering CO_2 from the atmosphere during the LGM. We also compile Pacific basin wide records of productivity and Pa/Th during the Holocene (0-11kya) and LGM (18-22kya) and find evidence consistent with a more frequent or persistent glacial El Niño state throughout much of the Pacific (North Pacific, Western Equatorial Pacific and EEP).
Article
A compilation of ages from 67 core tops in the eastern equatorial Pacific (EEP) does not display an easily discernible regional pattern. The ages range from 790 to over 15,000 years. The youngest core tops with the highest sediment focusing factors are located in the Panama Basin. There are weak but statistically significant inverse relationships between core top age and age-model based mass accumulation rates, bioturbation depth, linear sedimentation rate and sediment focusing factors. However, we found no statistically significant relationship between core top age and calcite dissolution in sediments or ²³⁰Th-normalized mass accumulation rates. We found evidence suggesting that greater amount of sediment focusing helps to preserve the carbonate fraction of the sediment where focusing is taking place. When focusing factors are plotted against percent calcite dissolved, we observe a strong inverse relationship, and core tops younger than 4500 years tend to occur where focusing factors are high and percent calcite dissolved values are low. Using labile organic carbon fluxes to estimate bioturbation depth in the sediments results in the observation that where bioturbation depth is shallow (<4 cm), the core top age has a strong, inverse relationship with sediment accumulation rate. We used the Globorotalia menardii Fragmentation Index (MFI) as an indicator of percent calcite dissolved in deep sea sediments. There is a distinct pattern to core top calcite dissolution in the EEP which delineates bands of high surface ocean productivity as well as the clear increase in dissolution downward on the flanks of the East Pacific Rise.
Article
Full-text available
Multiproxy geologic records of delta18O and Mg/Ca in fossil foraminifera from sediments under the Eastern Pacific Warm Pool (EPWP) region west of Central America document variations in upper ocean temperature, pycnocline strength, and salinity (i.e., net precipitation) over the past 30 kyr. Although evident in the paleotemperature record, there is no glacial-interglacial difference in paleosalinity, suggesting that tropical hydrologic changes do not respond passively to high-latitude ice sheets and oceans. Millennial variations in paleosalinity with amplitudes as high as ~4 practical salinity units occur with a dominant period of ~3-5 ky during the glacial/deglacial interval and ~1.0-1.5 ky during the Holocene. The amplitude of the EPWP paleosalinity changes greatly exceeds that of published Caribbean and western tropical Pacific paleosalinity records. EPWP paleosalinity changes correspond to millennial-scale climate changes in the surface and deep Atlantic and the high northern latitudes, with generally higher (lower) paleosalinity during cold (warm) events. In addition to Intertropical Convergence Zone (ITCZ) dynamics, which play an important role in tropical hydrologic variability, changes in Atlantic-Pacific moisture transport, which is closely linked to ITCZ dynamics, may also contribute to hydrologic variations in the EPWP. Calculations of interbasin salinity average and interbasin salinity contrast between the EPWP and the Caribbean help differentiate long-term changes in mean ITCZ position and Atlantic-Pacific moisture transport, respectively.
Article
Full-text available
Studying past changes in the eastern equatorial Pacific Ocean dynamics and their impact on precipitation on land gives us insight into how the Intertropical Convergence Zone (ITCZ) movements and the El Niño-Southern Oscillation modulate regional and global climate. In this study we present a multiproxy record of terrigenous input from marine sediments collected off the Ecuadorian coast spanning the last 500 kyr. In parallel we estimate sea surface temperatures (SST) derived from alkenone paleothermometry for the sediments off the Ecuadorian coast and complement them with alkenone-based SST data from the Panama Basin to the north in order to investigate SST gradients across the equatorial front. Near the equator, today's river runoff is tightly linked to SST, reaching its maximum either during the austral summer when the ITCZ migrates southward or during El Niño events. Our multiproxy reconstruction of riverine runoff indicates that interglacial periods experienced more humid conditions than the glacial periods. The north-south SST gradient is systematically steeper during glacial times, suggesting a mean background climatic state with a vigorous oceanic cold tongue, resembling modern La Niña conditions. This enhanced north-south SST gradient would also imply a glacial northward shift of the Intertropical Convergence Zone at least in vicinity of the cold tongue: a pattern that has not yet been reproduced in climate models.
Article
Holocene sediments from the Atlantic are characterized by 231Pa/230Th ratios below the production ratio of the two radionuclides in the water column (0.093), whereas Holocene sediments from the Southern Ocean have 231Pa/230Th > 0.093. This pattern of 231Pa deficit and excess was ascribed to southward 231Pa export from the Atlantic by the Atlantic thermohaline circulation (THC) as Pa is scavenged less efficiently by marine particles and more effectively transported by the THC than Th. The same pattern is observed in deposits of the Last Glacial Maximum (LGM), which led to a previous contention that the THC strength did not vary markedly through the last glacial termination. Here we embed a description of trace metal scavenging into a zonally averaged, circulation-biogeochemistry ocean model to explore the sensitivity of 231Pa/230Th in Atlantic and Southern Ocean sediments to THC changes. Our results show that the production of biogenic opal (which, unlike other marine particles, poorly fractionates Th and Pa) in the Southern Ocean water column determines the spatial pattern of the sensitivity. Also, 231Pa/230Th increases in the North Atlantic but changes little in the South Atlantic and decreases in the Southern Ocean as THC is reduced. The mean 231Pa/230Th of the whole Atlantic is therefore less sensitive to THC changes than the mean 231Pa/230Th of the North Atlantic. The current uncertainties in Atlantic mean 231Pa/230Th are too large to rule out a twofold reduction of the THC at the LGM. However, the increase in North Atlantic mean 231Pa/230Th simulated in response to a twofold THC reduction is larger than the observed change in the North Atlantic mean 231Pa/230Th from the LGM to Holocene. Comparing this change with the modeled sensitivity of North Atlantic 231Pa/230Th to THC variations indicates that the THC at the LGM could not have been reduced by >30% of its present strength. Experiments of transient THC changes indicate that high-resolution 231Pa/230Th records from North Atlantic sediments could also document thermohaline oscillations on century-to-millennial timescales.
Article
The constant-flux proxies excess 230Th (230Thxs) and extraterrestrial 3He (3HeET) are commonly used to calculate sedimentary mass accumulation rates and to quantify lateral advection of sediment at core sites. In settings with significant lateral input or removal of sediment, these calculations depend on the assumption that concentrations of 230Thxs and 3HeET are the same in both advected sediment and sediment falling through the water column above the core site. Sediment redistribution is known to fractionate grain sizes, preferentially transporting fine grains; though relatively few studies have examined the grain size distribution of 230Thxs and 3HeET, presently available data indicate that both are concentrated in fine grains, suggesting that fractionation during advection may bias accumulation rate and lateral advection estimates based on these proxies. In this study, we evaluate the behavior of 230Thxs and 3HeET in Holocene and last glacial samples from two cores from the Blake Ridge, a drift deposit in the western North Atlantic. At the end of the last glacial period, both cores received large amounts of laterally transported sediment enriched in fine-grained material. We find that accumulation rates calculated by normalization to 230Th and 3He are internally consistent despite large spatial and temporal differences in sediment advection. Our analyses of grain size fractions indicate that ~ 70% of 3HeET-bearing grains are in the < 20 μm fraction, with roughly equal amounts in the < 4 and 4–20 μm fractions. 230Thxs is concentrated in <4-μm grains relative to 4- to 20-μm grains by approximately a factor of 2 in Holocene samples and by a much larger factor (averaging a factor of 10) in glacial samples. Despite these enrichments of both constant-flux proxies in fine particles, the fidelity of 230Th- and 3He-based accumulation rate estimates appears to be preserved even in settings with extreme sediment redistribution, perhaps due to the cohesive behavior of fine particles in marine settings.
Article
Sediments of the Panama Basin contain >90% nonbiogenous components adjacent to Central and South America, but <5% in the area west of the Galapagos Islands. The nearshore deposits are quartz rich, with clay fractions consisting of roughly equal amounts of chlorite, illite, and kaolinite. These sediments are dispersed over distances of several hundred kilometers by currents in waters of intermediate depth (below the photic zone, but not deep enough to be affected by bottom topography), and to a lesser extent by bottom currents that flow through the deepest parts of the Basin. In the outer parts of the basin, the nonbiogenous deposits are quartz poor, with clay fractions dominated by authigenic montmorillonite. Winnowing and lateral reworking at the sea floor, which strongly affect the distribution of biogenous components, have little effect on the patterns of occurrence of the clay minerals.
Article
The short residence times of Th and Pa in seawater make them very responsive to changes in the ocean environment. We use a new multi-ion-counting technique to make Th and Pa isotope measurements in seawaters from a near-shore environment in which oceanic chemical tracers are not overwhelmed by terrestrial inputs (the Bahamas). An unusual feature of the Bahamas setting is the shallow depth of water residing on the bank tops. These waters have significantly lower ²³²Th/²³⁰Th (∼10,000) than those immediately adjacent to the banks (24,000–31,000) and a (²³¹Pa/²³⁰Th) near the production ratio (∼0.1). The change in ²³²Th/²³⁰Th and (²³¹Pa/²³⁰Th) on the bank tops is explained by almost quantitative removal of Th and Pa by scavenging, and their replacement with a mixture of ²³⁰Th and ²³¹Pa alpha-recoiled from the underlying carbonates, together with Th from dust dissolution. Analysis of a water profile in the Tongue of the Ocean, which separates the Great and Little Bahama Banks, allows us to trace the movement of bank-top water to depth. A distinct minimum in both ²³²Th/²³⁰Th (∼13,000) and (²³¹Pa/²³⁰Th) (∼0.5) is observed at ∼430 m and is interpreted to reflect density cascading of bank-top water with entrained carbonate sediment. These results suggest that Th and Pa can be used as water-mass tracers in near-shore environments. Uranium concentration measurements on the same waters demonstrate that U is conservative across a range in salinity of 2 psu, with a concentration of 3.33 ppb (at a salinity of 35).
Article
Biogenic particle fluxes from highly productive surface waters, boundary scavenging, and hydrothermal activity are the main factors influencing the deposition of radionuclides in the area of the Galapagos microplate, eastern Equatorial Pacific. In order to evaluate the importance of these three processes throughout the last 100 kyr, concentrations of the radionuclides 10Be, 230Th, and 231Pa, and of Mn and Fe were measured at high resolution in sediment samples from two gravity cores KLH 068 and KLH 093. High biological productivity in the surface waters overlying the investigated area has led to 10Be and 231Pa fluxes exceeding production during at least the last 30 kyr and probably the last 100 kyr. However, during periods of high productivity at the up welling centers off Peru and extension of the equatorial high-productivity zone, a relative loss of 10Be and 231Pa may have occurred in these sediment cores because of boundary scavenging. The effects of hydrothermal activity were investigated by comparing the 230Thex concentrations to the Mn/Fe ratios and by comparing the fluxes of 230Th and 10Be which exceed production. The results suggest an enhanced hydrothermal influence during isotope stages 4 and 5 and to a lesser extent during isotope stage 1 in core KLH 093. During isotope stages 2 and 3, the hydrothermal supply of Mn was deposited elsewhere, probably because of changes in current regime or deep water oxygenation. A strong increase of the Mn/Fe ratio at the beginning of climatic stage 1 which is not accompanied by an increase of the 230Thex concentration is interpreted to be an effect of Mn remobilization and reprecipitation in the sediment.
Article
The sedimentary inventory of 230Th is often employed to distinguish between areas which receive excess sediment as the result of along-bottom transport from those which lose sediment as a result of winnowing. This process is referred to as “focusing.” A case is made here that, at least in the western equatorial Pacific, the excess 230Th in on-equator sediment is more likely scavenged from the overlying water column by the rain of organic matter than retransported along the bottom. If so, lateral variations in biological productivity may well contribute to the excess 230Th inventories elsewhere as well.