palo20265-sup-0001-supplementary

advertisement
1
2
Paleoceanography
3
Supporting Information for
4
5
Export Production Fluctuations in the Eastern Equatorial Pacific during the
Pliocene-Pleistocene: Reconstruction Using Barite Accumulation Rates
6
Zhongwu Ma1,2, Christina Ravelo2, Zhonghui Liu3, Liping Zhou1, Adina Paytan2*
7
1Laboratory
for Earth Surface Processes, Department of Geography, Peking University, Beijing 100871, China
8
9
10
2University
3Department
of California Santa Cruz, Santa Cruz CA, 95064, USA
of Earth Sciences, The University of Hong Kong, James Lee Science Building, Pokfulam Road, Hong
Kong,
11
12
13
14
15
16
17
Contents of this file
Text S1 to S3
Figures S1 to S5
18
19
20
Introduction
Marine barite crystals form in the water column in microenvironments associated with
21
decaying organic matter and their accumulation in marine sediments has been used
22
extensively as an export production proxy. A detailed description of the strengths and
23
limitations of barite as an export productivity proxy are given by Paytan and Griffith [2007]
24
and a summary of these points is included in the supplemental text below. Since one of the
25
potential limitations associated with the use of barite accumulation rates (BAR, which is the
1
26
product of weight % barite and the sediment mass accumulation rates) is the impact of major
27
constituent in the sediment (CaCO3 in this case) on the weight % of minor constituents (barite
28
in this case) on BAR, several means for calculating BAR are shown. This includes BAR
29
derived using an orbitally tunes age model, 230Th and 3He based accumulation rates (Figure
30
S1) as well as a comparison between wt% barite and wt% CaCO3 and wt% CaCO3 and BAR
31
and carbonate accumulation rates (Figure S2A). Finally we also included a plot of BAR
32
calculated on a carbonate free bases by removing the carbonate weight from the total
33
sediment used in the calculation (Figure S2B). .
34
The record of BAR derived export production over the past 4.3 Ma at ODP site 849 is
35
divided into three intervals which are illustrated in Figure S3 using 800kry, 400kyr and
36
200kyr smoothed curves; this highlights the long term trends in the record. We also show a
37
detailed statistical relationship between reconstructed export production and the climate
38
system by comparing BAR and benthic oxygen isotope using spectral analysis methods
39
(Figure S4). Finally BAR and total organic carbon accumulation rates from the same samples
40
for a selected interval at site 849 are plotted together to demonstrate that they both correspond
41
to productivity in Figure S5.
42
Text S1. Barite accumulation rates as an export productivity proxy
43
The ocean is under-saturated with respect to barite; however, authigenic barite
44
crystals form in the water column in microenvironments with decaying organic debris
45
[Arrhenius, 1959; Church, 1970; Goldberg and G.O.S, 1958]. Organisms normally
46
contain barium in their cells [Paytan and Griffith, 2007] and during the decay of
47
marine organic matter, barium is released saturating the microenvironment with
48
respect to barite and precipitating barite crystal [Ganeshram et al., 2003]. Bacteria
49
may also be involved by providing nucleation sites [Gonzalez-Muñoz et al., 2012].
50
The nutrient-like depth distribution of dissolved Ba and the distribution of particulate
51
barite in the water column [Dehairs et al., 1980; Jacquet et al., 2005; Jeandel et al.,
52
1996], as well as controlled mesocosm experiments [Ganeshram et al., 2003] and
53
analyses of Ba isotopes in seawater and particulate matter [Horner et al., 2015] all
2
54
support this mechanism of barite formation.
55
Because barite has low solubility, most of the barite that forms in the water
56
column is thought to reach the ocean sediment [Paytan and Kastner 1996a] with as
57
much as 30% of the barite “rain” to the sediment persevered in oxic sediments
58
[Dymond et al., 1992; Paytan et al., 1996b] a considerable advantage over proxies
59
such as organic carbon or opal which are not preserved as well. Once barite is buried,
60
regeneration of various Ba carrying phases within the upper few centimeters of the
61
sediment results in barite saturation in pore fluids thus the remaining barite is
62
preserved in sediments that do not undergo sulfate reduction [Paytan and Kastner
63
1996a; Paytan et al., 2007].
64
The strong correlation between marine barite and the export of organic carbon to
65
depth [Dymond et al., 1992; Eagle et al., 2003; Paytan et al., 1996b] lends itself to the
66
use of Ba and barite accumulation rates (BAR) for the reconstruction of export
67
production [Bains et al., 2000; Bonn et al., 1998; Eagle et al., 2003; Nürnberg, 1997;
68
Paytan et al., 2007; Schmitz, 1987]. BAR represents the Ba that is associated with Corg
69
flux below the euphotic zone, and BAR can directly be used to estimate export
70
production or new production [Paytan et al., 1996a; Paytan et al., 1997]. Excess Ba
71
(total Ba in the sediment normalized to Al or Ti to correct for the terrigenous
72
component of Ba input [Dymond et al., 1992] has also been using for export
73
production reconstruction. The correction or calculation of excess Ba is done to
74
account only for the so called biogenic Ba component which is equivalent to barite.
75
The correction however does not work at all sites and a direct measurement of barite
76
is as done and reported here is preferable [Averyt and Paytan, 2004; Eagle et al.,
77
2003; Gonneea and Paytan, 2006].
78
When significant sulfate reduction occurs, barite is dissolved and can no longer be
79
used as a productivity proxy. However, barite that forms under diagenetic conditions
80
or in hydrothermal environments can be identified based on crystal morphology (see
81
image below of marine barite separated from the sediments) as well as using S and/or
3
82
Sr isotopes in the barite [Paytan and Griffith, 2007 and references therein] and
83
samples used in this study are all of water column origin and have not been impacted
84
by diagenesis based on these criteria [Markovik et al., 2015].
85
86
Another process that may affect BAR calculations is sediment focusing, the
87
redistribution of recently deposited material (i.e., syndepositional redistribution) from
88
zones of sediment winnowing to zones to sediment focusing [Francois et al., 2004].
89
This could affect BAR by changing mass accumulation rates (MAR) without
90
impairing the stratigraphy. In our study we used LSR to compute MAR and calculate
91
BAR, hence we did not account for sediment focusing. However, we compare these
92
values to MAR based on thorium in the first 500 kyr and helium between 1.3 to 1.6
93
Ma (where such data are available) (e.g. there is little recognized refocusing of
94
sediment at this site according to G. Winckler’s personal communication, 2015). Our
95
BARs show generally higher values (by a factor of 2) when compared to the thorium
96
based BAR and slightly lower when compared to helium based BAR (Figure S1), this
97
would affect the amplitude but not the trends we see in the record. Our discussion
98
focuses on observed trends and does not emphasize the absolute values. Notably each
99
of the methods used for calculating sedimentation rates has its own associated
100
assumptions and limitations and additional research should be done, which is beyond
101
the scope of this work, to determine the reasons of the systematic offsets between
102
these three methods for calculating sedimentation rates at this site and in general.
4
103
An additional potential limitation of using BAR is the impact of major
104
constituents in the sediment (e.g. CaCO3) on the wt% of minor constituents (e.g.
105
barite). In other words because barite is a minor component in marine sediments, wt%
106
barite is affected by changes in abundance of other sedimentary components and
107
particularly CaCO3 which is the major sedimentary phase at site 849. To account for
108
that we use the BAR and not wt% barite to reconstruct export productivity. Using
109
mass accumulation rates should reduce the impact CaCO3 changes because these
110
changes are incorporated in the sediment mass accumulation rates used. To
111
demonstrate however that carbonate content does not exert the only control on barite
112
distribution we plot the relation between barite abundance (ppm) and CaCO3
113
abundance (wt%) (Figure S2 a) as well as the relationships between BAR and wt%
114
CaCO3 and BAR and CAR. As seen in Figure S2a while a slight negative correlation
115
is indeed observed it is quite weak and illustrates that changes in carbonate abundance
116
do not solely control the barite distribution. The correlation is weaker when BAR is
117
used as expected. Actually when we used the carbonate free BAR (BAR calculated
118
from the non-carbonate materials by deducted the carbonate from bulk sediments),
119
trends are similar to the BAR record (Figure S2 b). Few Ba/Al data we have (not
120
shown) which is immune to the carbonate dilution effect also support this conclusion.
121
The 4.3 Ma BAR based export production record in the Eastern Equatorial Pacific
122
(EEP, Site 849) fluctuates considerably on orbital and on long (million years) time
123
scales. The long term trends in the record can be divided into three distinct intervals
124
as seen when the data is plotted using 800 kyr, 400 kyr and 200 kyr smoothing. These
125
3 intervals are seen also in a Gaussian fit of the 800 kyr smoothed BAR data (Figure
126
S3).
127
Text S2. Relations between BAR and the benthic oxygen isotope record
128
To identify whether periodicities presented in our data are statistically significant,
129
we used the Singular Spectrum Analyses - MultiTaper Method (SSA-MTM) Tool kit
130
to map both benthic δ18O and the BAR spectral densities with the multi-taper method
5
131
assuming a red noise model [Ghil et al., 2002]. Data were interpolated to be evenly
132
distributed with a resolution of 10 kyr under AnalySeries software before spectral
133
analyze [Paillard et al., 1996]. This is as close as possible to the average original
134
temporal resolution, although there are intervals with resolution that is higher and
135
intervals with resolution that is lower. Because of the possibility of aliasing orbital
136
scale variability in some intervals, we used MTM evolutive spectrum analysis, carried
137
out using AnalySeries software, with a large 600 kyr time window of ~100%
138
oversampling; constant 100% pre-whitening was adopted here to diminish low
139
frequency signals. To test the coherence between reconstructed productivity and the
140
benthic oxygen isotope record, the 41 kyr and 100 kyr bands were selected for the
141
cross-spectral analysis, with a 400 kyr time window, 30% lags and 200 kyr increment.
142
This was conducted by the Arand software package [Howell et al., 2006]. Strong
143
coherence at the 80% and 95% level (Figure S4) is found during some intervals, but
144
low sampling resolution may explain the low spectral power and low coherency in
145
some of the other intervals.
146
The BAR spectral map shows some similarity to the benthic δ18O spectral map,
147
with differences that might be explained by the poor resolution of our BAR in some
148
intervals (Figure S4 a, b). With regards to the moving, and overlapping, 400-kyr
149
windows used for our analyses, there is coherency (80% level) between BAR and
150
δ18O in the obliquity (41-kyr band) in 10 out of the 15 windows analyzed between 4.3
151
Ma to ~1 Ma, and in none of the windows analyzed between 1.0 – 0 Ma. For the 100-
152
kyr band, in the interval between 1.0 – 0 Ma, there is coherency between BAR and
153
δ18O in 3 out of the 4 windows analyzed (Figure S4 c). Notably, in the intervals where
154
BAR and δ18O varied coherently (80% level), BAR (productivity) 41-kyr variability
155
was in phase, or nearly in phase, with δ18O variability between 4.3 Ma to 1.1 Ma,
156
while 100-kyr variability lagged δ18O variability between 1.1 Ma and present (Figure
157
S4 c, d).
6
158
Text S3. Comparison of BAR with other productivity records in the EEP
159
The BAR record is compared to other productivity proxies in this region (opal,
160
alkenone, organic C, etc.), as discussed in the text. In addition select samples between
161
1.8 and 2.65 Ma were analyzed for TOC on the EA-IRMS in the UCSC Stable Isotope
162
Lab, following the method described by [Krupinski et al., 2013]. The records show
163
that at this interval BAR and Corg records are similar and a general correspondence
164
between TOC and BAR is seen (Figure S5).
7
165
166
167
168
169
170
171
172
Figure S1. BAR derived from oxygen isotopes chronology (black line, used in this study),
thorium (green line) and helium (red line and red shadow) based BAR [Winckler et al., 2005;
Winckler et al., 2008]. Note Thorium based BAR is plotted with a different Y-axis (green
labeled), while Helium and oxygen isotopes based chronology BAR have the same Y-axis
(black labeled). BAR in the red line during the early Pleistocene was derived from a constant
3
HeET of 0.8 pcc STP cm-2 kyr-1 and the boundary conditions are 0.5 and 1.1 pcc STP cm-2 kyr1
respectively.
8
173
174
175
176
Figure S2 a. Carbonate abundance (wt%) and carbonate accumulation rates (CAR)
plot against barite abundance (ppm) and barite accumulation rates (BAR). BAR have
less influence by carbonate dissolution effect comparing with the barite abundance.
9
177
178
179
180
Figure S2 b. (A) BAR compare with (B) carbonate free BAR, (C) the carbonate abundance
are also shown in the figure. Bold line in the BAR and carbonate free BAR panels are 100 kyr
smoothed line. The general trends of BAR and carbonate free BAR are similar.
181
10
182
183
184
185
186
Figure S3. A 200-kyr, 400-kyr and 800-kyr averaging smoothed BAR record at site 849 are
plotted to display the 3 intervals we identify in the record. Between 4.3 and 3 Ma BAR is
high, from 3 to 1 Ma a decreasing BAR trend is seen and from 1 Ma to present BAR is
increasing. A Gaussian fit for the 800 kyr smoothed data shows these clear transitions.
187
11
188
189
190
191
192
193
Figure S4. MTM spectral analysis and the MTM evolution for (a) benthic δ18O and (b) the
BAR at site 849; (c) the coherency between benthicδ18O and the BAR; and (d) phase relation
between the records which is only shown for the windows in which the benthic 18O and the
BAR are coherent at the 80% level.
12
194
195
196
Figure S5. Comparison between BAR and organic carbon accumulation rates at ODP site 849
between 1.8 and 2.6 Ma
13
197
References
198
Arrhenius, G. (1959), Sedimentation on the ocean floor, Researches in geochemistry.
199
200
Wiley, New York, 1
Averyt, K. B., and A. Paytan (2004), A comparison of multiple proxies for export
201
production in
the equatorial
202
doi:doi:10.1029/2004PA001005
Pacific, Paleoceanography, 19(4),
1-14.
203
Bains, S., R. D. Norris, R. M. Corfield, and K. L. Faul (2000), Termination of global
204
warmth at the Palaeocene/Eocene boundary through productivity feedback,
205
Nature, 407(6801), 171-174. doi:10.1038/35025035
206
Bonn, W. J., F. X. Gingele, H. Grobe, A. Mackensen, and D. K. Futterer (1998),
207
Palaeoproductivity at the Antarctic continental margin: opal and barium records
208
for the last 400 ka, Palaeogeography, Palaeoclimatology, Palaeoecology, 139,
209
195-211. doi:10.1016/S0031-0182(97)00144-2
210
Calvo, E., C. Pelejero, L. D. Pena, I. Cacho, and G. A. Logan (2011), Eastern Equatorial
211
Pacific productivity and related-CO2 changes since the last glacial period,
212
Proceedings of the National Academy of Sciences, 108(14), 5537-5541.
213
doi:10.1073/pnas.1009761108
214
215
Church, T. M. (1970), Marine barite, Ph.D. Thesis thesis, University of California, San
Diego.
216
Dehairs, F., R. Chesselet, and J. Jedwab (1980), Discrete suspended particles of barite
217
and the barium cycle in the open ocean, Earth and Planetary Science Letters,
218
49(2), 528-550. doi:10.1016/0012-821X(80)90094-1
219
Dekens, P. S., A. C. Ravelo, and M. D. McCarthy (2007), Warm upwelling regions in
220
the
Pliocene
warm
221
doi:10.1029/2006pa001394
period,
Paleoceanography,
22(3),
PA3211.
1
222
Diester-Haass, L., K. Billups, and K. C. Emeis (2006), Late Miocene carbon isotope
223
records and marine biological productivity: Was there a (dusty) link?,
224
Paleoceanography, 21(4), PA4216. doi:10.1029/2006pa001267
225
Dymond, J., E. Suess, and M. Lyle (1992), Barium in Deep-Sea Sediment: A
226
Geochemical Proxy for Paleoproductivity, Paleoceanography, 7(2), 163-181.
227
doi:10.1029/92pa00181
228
Eagle, M., A. Paytan, K. R. Arrigo, G. van Dijken, and R. W. Murray (2003), A
229
comparison between excess barium and barite as indicators of carbon export,
230
Paleoceanography, 18(1), 21.21-21.13. doi:10.1029/2002PA000793
231
Etourneau, J., R. S. Robinson, P. Martinez, and R. Schneider (2013), Equatorial Pacific
232
peak in biological production regulated by nutrient and upwelling during the late
233
Pliocene/early
234
doi:10.5194/bg-10-5663-2013
235
Pleistocene
cooling,
Biogeosciences,
10(8),
5663-5670.
Francois, R., M. Frank, M. M. Rutgers van der Loeff, and M. P. Bacon (2004),
230
Th
236
normalization: An essential tool for interpreting sedimentary fluxes during the
237
late Quaternary, Paleoceanography, 19(1), PA1018. doi:10.1029/2003PA000939
238
Ganeshram, R. S., R. François, J. Commeau, and S. L. Brown-Leger (2003), An
239
experimental investigation of barite formation in seawater, Geochimica et
240
Cosmochimica Acta, 67(14), 2599-2605. doi:10.1016/s0016-7037(03)00164-9
241
Gartner, S., J. Chow, and R. J. Stanton Jr (1987), Late Neogene paleoceanography of
242
the eastern Caribbean, the Gulf of Mexico, and the eastern Equatorial Pacific,
243
Marine Micropaleontology, 12, 255-304. doi:10.1016/0377-8398(87)90024-7
244
245
246
Ghil, M., et al. (2002), Advanced spectral methods for climatic time series, Reviews of
Geophysics, 40(1), 1003. doi:10.1029/2000RG000092
Goldberg, E. D., and A. G.O.S (1958), Chemistry of Pacific pelagic sediments,
2
247
Geochimica et Cosmochimica Acta, 13(2–3), 153-212. doi:10.1016/0016-
248
7037(58)90046-2
249
Gonneea, M. E., and A. Paytan (2006), Phase associations of barium in marine
250
sediments,
Marine
Chemistry,
251
doi:10.1016/j.marchem.2005.12.003
100(1-2),
124-135.
252
Gonzalez-Muñoz, M. T., F. Martinez-Ruiz, F. Morcillo, J. D. Martin-Ramos, and A.
253
Paytan (2012), Precipitation of barite by marine bacteria: A possible mechanism
254
for marine barite formation, Geology, 40(8), 675-678. doi:10.1130/g33006.1
255
Horner, T. J., C. W. Kinsley, and S. G. Nielsen (2015), Barium-isotopic fractionation in
256
seawater mediated by barite cycling and oceanic circulation, Earth and Planetary
257
Science Letters, 430, 511-522. doi:10.1016/j.epsl.2015.07.027
258
259
Howell, P., N. Pisias, J. Ballance, J. Baughman, and L. Ochs (2006), ARAND TimeSeries Analysis Software, Brown University, Providence RI.
260
Jacquet, S. H. M., F. Dehairs, D. Cardinal, J. Navez, and B. Delille (2005), Barium
261
distribution across the Southern Ocean frontal system in the Crozet–Kerguelen
262
Basin, Marine Chemistry, 95(3–4), 149-162. doi:10.1016/j.marchem.2004.09.002
263
Jeandel, C., B. Dupré, G. Lebaron, C. Monnin, and J.-F. Minster (1996), Longitudinal
264
distributions of dissolved barium, silica and alkalinity in the western and southern
265
Indian Ocean, Deep Sea Research Part I: Oceanographic Research Papers, 43(1),
266
1-31. doi:10.1016/0967-0637(95)00098-4
267
Krupinski, N. B. Q., J. R. Marlon, A. Nishri, J. H. Street, and A. Paytan (2013), Climatic
268
and human controls on the late Holocene fire history of northern Israel,
269
Quaternary Research, 80(3), 396-405. doi:10.1016/j.yqres.2013.06.012
270
Lawrence, K. T., Z. Liu, and T. D. Herbert (2006), Evolution of the Eastern Tropical
271
Pacific Through Plio-Pleistocene Glaciation, Science, 312(5770), 79-83.
3
272
doi:10.1126/science.1120395
273
Liu, Z., and T. D. Herbert (2004), High-latitude influence on the eastern equatorial
274
Pacific climate in the early Pleistocene epoch, Nature, 427(6976), 720-723.
275
doi:10.1038/nature02338
276
Markovic, S., A. Paytan, and U. G. Wortmann (2015), Pleistocene sediment offloading
277
and the global sulfur cycle, Biogeosciences, 12(10), 3043-3060. doi:10.5194/bg-
278
12-3043-2015
279
Mix, A. C., et al. (Eds.) (2003), Proceedings of the Ocean Drilling Program: Initial
280
Reports,
Ocean
Drill.
Program,
281
doi:doi:10.2973/odp.proc.ir.202.2003
College
Station,
TX.
282
Nürnberg, C. C. B., G; Schlüter, M; Frank, M (1997), Barium accumulation in the
283
Atlantic sector of the Southern Ocean:results from 190,000-year records,
284
Paleoceanography, 12(0883-8305), 594-603. doi:10.1029/97PA01130
285
286
Paillard, D., L. Labeyrie, and P. Yiou (1996), Macintosh Program performs time-series
analysis, Eos Trans. AGU, 77(39), 379-379. doi:10.1029/96eo00259
287
Paytan, A., and E. M. Griffith (2007), Marine barite: Recorder of variations in ocean
288
export productivity, Deep Sea Research Part II: Topical Studies in Oceanography,
289
54(5-7), 687-705. doi:10.1016/j.dsr2.2007.01.007
290
Paytan, A., and M. Kastner (1996a), Benthic Ba fluxes in the central Equatorial Pacific,
291
implications for the oceanic Ba cycle, Earth and Planetary Science Letters,
292
142(3–4), 439-450. doi:10.1016/0012-821X(96)00120-3
293
Paytan, A., M. Kastner, and F. P. Chavez (1996b), Glacial to Interglacial Fluctuations
294
in Productivity in the Equatorial Pacific as Indicated by Marine Barite, Science,
295
274(5291), 1355-1357. doi:10.1126/science.274.5291.1355
296
Paytan, A., R. W. Murray, M. Kastner, and M. Leinen (1997), Total barium vs. amount
4
297
of barite in sediments as indicators of productivity, LPI Contribution, 921, 159-
298
160
299
Paytan, A., K. Averyt, K. Faul, E. Gray, and E. Thomas (2007), Barite accumulation,
300
ocean productivity, and Sr/Ba in barite across the Paleocene Eocene Thermal
301
Maximum, Geology, 35(12), 1139-1142. doi:10.1130/g24162a.1
302
Robinson, R. S., P. Martinez, L. D. Pena, and I. Cacho (2009), Nitrogen isotopic
303
evidence for deglacial changes in nutrient supply in the eastern equatorial Pacific,
304
Paleoceanography, 24(4), PA4213. doi:10.1029/2008pa001702
305
Schmitz, B. (1987), Barium, equatorial high productivity, and the northward wandering
306
of
the
Indian
continent,
307
doi:0.1029/PA002i001p00063
Paleoceanography,
2(1),
63-77.
308
Schneider, B., and A. Schmittner (2006), Simulating the impact of the Panamanian
309
seaway closure on ocean circulation, marine productivity and nutrient cycling,
310
Earth
311
doi:10.1016/j.epsl.2006.04.028
and
Planetary
Science
Letters,
246(3-4),
367-380.
312
Steph, S., R. Tiedemann, M. Prange, J. Groeneveld, M. Schulz, A. Timmermann, D.
313
Nürnberg, C. Rühlemann, C. Saukel, and G. H. Haug (2010), Early Pliocene
314
increase in thermohaline overturning: A precondition for the development of the
315
modern equatorial Pacific cold tongue, Paleoceanography, 25(2), PA2202.
316
doi:10.1029/2008pa001645
317
Winckler, G., R. F. Anderson, and P. Schlosser (2005), Equatorial Pacific productivity
318
and dust flux during the mid-Pleistocene climate transition, Paleoceanography,
319
20(4), PA4025. doi:10.1029/2005pa001177
320
Winckler, G., R. F. Anderson, M. Q. Fleisher, D. McGee, and N. Mahowald (2008),
321
Covariant glacial-interglacial dust fluxes in the equatorial Pacific and Antarctica,
5
322
Science, 320(5872), 93-96. doi:10.1126/science.1150595
6
Download