Hostname: page-component-8448b6f56d-dnltx Total loading time: 0 Render date: 2024-04-18T18:58:30.528Z Has data issue: false hasContentIssue false

Biogenic opal production changes during the Mid-Pleistocene Transition in the Bering Sea (IODP Expedition 323 Site U1343)

Published online by Cambridge University Press:  20 January 2017

Sunghan Kim
Affiliation:
Department of Oceanography, Pusan National University, Busan 609-735, South Korea
Kozo Takahashi
Affiliation:
Hokusei Gakuen University, Sapporo 004-8631, Japan Department of Earth and Planetary Sciences, Kyushu University, Fukuoka 812-8581, Japan
Boo-Keun Khim*
Affiliation:
Department of Oceanography, Pusan National University, Busan 609-735, South Korea
Yoshihiro Kanematsu
Affiliation:
Department of Earth and Planetary Sciences, Kyushu University, Fukuoka 812-8581, Japan
Hirofumi Asahi
Affiliation:
Department of Oceanography, Pusan National University, Busan 609-735, South Korea
Ana Christina Ravelo
Affiliation:
Ocean Sciences Department, University of California, Santa Cruz, CA 95064, USA
*
*Corresponding author. E-mail address:bkkhim@pusan.ac.kr (B.-K. Khim).

Abstract

Biogenic opal content and mass accumulation rate (MAR) at IODP Expedition 323 Site U1343 were found to fluctuate consistently, generally being high under warm conditions and low under cold conditions during the last 2.4 Ma. Continuous wavelet transform analysis of the normalized biogenic opal content indicates that export production in the Bering Sea varied predominantly at 41-ka periodicity before 1.25 Ma, and shifted to 100-ka periodicity at the onset of the Mid-Pleistocene Transition (MPT; 1.25–0.7 Ma). The 100-ka cycles dominated until the Holocene. Export production in the Bering Sea decreased markedly in the Bering Sea two times during the MPT: the first occurred at the beginning of the MPT (1.25 Ma) and the second in the middle of the MPT (0.9 Ma). These decreases coincided with both increases in the relative abundance of sea-ice diatoms and decreases in the warm-water diatom species Neodenticula seminae, indicating that reductions in export production in the Bering Sea during the MPT were associated with climate cooling. Decreases in export production in the Bering Sea during the MPT were most likely associated with the increased influence of polar/Arctic domains on the high-latitude North Pacific.

Type
Research Article
Copyright
University of Washington

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Footnotes

1 Equally-contributing author.

References

Asahi, H., Kender, S., Ikehara, M., Sakamoto, T., Ravelo, A.C., Alvarez Zarikian, C.A., and Takahashi, K. Foraminiferal oxygen isotope records at the Bering slope (IODP Exp. 323 Site U1343) provide an orbital scale age model and indicate pronounced changes during the Mid-Pleistocene Transition. American Geophysical Union, Fall Meeting, abstract #PP31B-1870. (2011). Google Scholar
Boyd, P.W., Law, C.S., Wong, C.S., Nojiri, Y., Tsuda, A., Levasseur, M., Takeda, S., Rivkin, R., Harrison, P.J., Strzepek, R., Gower, J., McKay, R.M., Abraham, E., Arychuk, M., Barwell-Clarke, J., Crawford, W., Crawford, D., Hale, M., Harada, K., Johnson, K., Kiyosawa, H., Kudo, I., Marchetti, A., Miller, W., Needoba, J., Nishioka, J., Ogawa, H., Page, J., Robert, M., Saito, H., Sastri, A., Sherry, N., Soutar, T., Sutherland, N., Taira, Y., Whitney, F., Wong, S.K.E., and Yoshimura, T. The decline and fate of an iron-induced subarctic phytoplankton bloom. Nature 428, (2004). 549553. http://dx.doi.org/10.1038/nature02437Google Scholar
Brunelle, B.G., Sigman, D.M., Cook, M.S., Keigwin, L.D., Haug, G.H., Plessen, B., Schettler, G., and Jaccard, S.L. Evidence from diatom-bound nitrogen isotopes for subarctic Pacific stratification during the last ice age and a link to North Pacific denitrification changes. Paleoceanography 22, (2007). PA1215 http://dx.doi.org/10.1029/2005PA001205Google Scholar
Caissie, B.E., Brigham-Grette, J., Lawrence, K.T., Herbert, T.D., and Cook, M.S. Last glacial maximum to Holocene sea surface conditions at Umnak Plateau, Bering Sea, as inferred from diatom, alkenone, and stable isotope records. Paleoceanography 25, (2010). PA1206 http://dx.doi.org/10.1029/2008PA001671Google Scholar
Clark, P.U., Archer, D., Pollard, D., Blum, J.D., Rial, J.A., Brovkin, V., Mix, A.C., Pisias, N.G., and Roy, M. The Middle Pleistocene Transition: characteristics, mechanisms, and implications for long-term changes in atmospheric pCO2 . Quaternary Science Reviews 25, (2006). 31503184.Google Scholar
Clement, J.L., Maslowski, W., Cooper, L.W., Grebmeier, J.M., and Walczowski, W. Ocean circulation and exchange through the northern Bering Sea—1979–2001 model results. Deep-Sea Research Part II 52, (2005). 35093540.Google Scholar
Crundwell, M., Scott, G., Naish, T., and Carter, L. Glacial–interglacial ocean climate variability from planktonic foraminifera during the Mid-Pleistocene Transition in the temperate Southwest Pacific, ODP Site 1123. Palaeogeography, Palaeoclimatology, Palaeoecology 260, (2008). 202229.CrossRefGoogle Scholar
de Garidel-Thoron, T., Rosenthal, Y., Bassinot, F., and Beaufort, L. Stable sea surface temperatures in the western Pacific warm pool over the past 1.75 million years. Nature 433, (2005). 294298.Google Scholar
DeMaster, D.J. The supply and accumulation of silica in the marine environment. Geochimica et Cosmochimica Acta 5, (1981). 17151732.Google Scholar
Expedition 323 Scientists, Bering Sea paleoceanography: Pliocene–Pleistocene paleoceanography and climate history of the Bering Sea. IODP Prel. Rept., 323. (2009). http://dx.doi.org/10.2204/iodp.pr.323.2009Google Scholar
Haug, G.H., Maslin, M.A., Sarnthein, M., Stax, R., Tiedemann, Evolution of northwest Pacific sedimentation patterns since 6 Ma (Site 882). Rea, D.K., Basov, I.A., Scholl, D.W., Allan, J.F. Proc. ODP Sci. Results 145, (1995). Ocean Drilling Program, College Station, TX. 293314.Google Scholar
Haug, G.H., Sigman, D.M., Tiedemann, R., Pedersen, T.F., and Sarnthein, M. Onset of permanent stratification in the subarctic Pacific Ocean. Nature 401, (1999). 779782.Google Scholar
Imbrie, J.Z., Imbrie-Moore, A., and Lisiecki, L.E. A phase-space model for Pleistocene ice volume. Earth and Planetary Science Letters 307, (2011). 94102.Google Scholar
Itaki, T., Uchida, M., Kim, S., Shin, H.S., Tada, R., and Khim, B.K. Late Pleistocene stratigraphy and paleoceanographic implications in northern Bering Sea slope sediments: evidence from the radiolarian species Cycladophora davisiana . Journal of Quaternary Science 24, (2009). 856865.Google Scholar
Iwasaki, S., Takahashi, K., Kanematsu, Y., Sakamoto, T., Sakai, S., Onodera, J., Asahi, H., Ravelo, A.C., Science Team of IODP Exp. 323, Paleoproductivity and paleoceanography of the last 4.3 Myrs at IODP Exp. 323 Site U1341 in the Bering Sea based on biogenic opal content. American Geophysical Union, Fall Meeting, abstract #PP31B-1865. (2011). Google Scholar
Kanematsu, Y., Takahashi, K., Kim, S., Khim, B.K., and Asahi, H. Changes in biogenic opal productivity with Milankovitch cycle during the last 1.3 Myrs at IODP Expedition 323 Sites U1341, U1343, and U1345 in the Bering Sea. Quaternary International 310, (2013). 213220.Google Scholar
Kim, S., Khim, B.K., Takahashi, K., IODP Expedition 323 Scientists, High-resolution variation of biogenic opal content in the Bering Sea (IODP Expedition 323, Site U1343) from the late Pliocene to early Pleistocene (2.2 Ma to 1.4 Ma). American Geophysical Union, Fall Meeting, abstract #PP21B-1695. (2010). Google Scholar
Kim, S., Khim, B.K., Uchida, M., and Tada, R. Millennial-scale paleoceanographic events and implication for the intermediate-water ventilation in the northern slope area of the Bering Sea during the last 71 kyrs. Global and Planetary Change 79, (2011). 8998.Google Scholar
Kitamura, A., and Kawagoe, T. Eustatic sea-level change at the Mid-Pleistocene climate transition: new evidence from the shallow-marine sediment record of Japan. Quaternary Science Reviews 25, (2006). 323335.Google Scholar
Lisiecki, L.E., and Raymo, M.E. A Pliocene–Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 20, (2005). PA1003 http://dx.doi.org/10.1029/2004PA001071Google Scholar
Liu, Z., and Herbert, T.D. High-latitude influence on the eastern equatorial Pacific climate in the early Pleistocene epoch. Nature 427, (2004). 720723.Google Scholar
Marino, M., Maiorano, P., and Lirer, F. Changes in calcareous nannofossil assemblages during the Mid-Pleistocene Revolution. Marine Micropaleontology 69, (2008). 7090.Google Scholar
Martinez-Garcia, A., Rosell-Mele, A., McClymont, E.L., Gersonde, R., and Haug, G.H. Subpolar link to the emergence of the modern equatorial Pacific cold tongue. Science 328, (2010). 15501553.Google Scholar
McClymont, E.L., and Rossel-Mele, A. Links between the onset of modern Walker circulation and the mid-Pleistocene climate transition. Geology 33, (2005). 389392.Google Scholar
McClymont, E.L., Rosell-Mele, A., Haug, G., and Lloyd, J. Expansion of subarctic water masses in the North Atlantic and Pacific oceans and implications for mid-Pleistocene ice sheet growth. Paleoceanography 23, (2008). PA4214 http://dx.doi.org/10.1029/2008PA001622Google Scholar
Mortlock, R.A., and Rroelich, P.N. A simple method for the rapid determination of opal in pelagic marine sediments. Deep-Sea Research 36, (1989). 14151426.Google Scholar
Mudelsee, M., and Raymo, M.E. Slow dynamics of the Northern Hemisphere glaciation. Paleoceanography 20, (2005). PA4022 http://dx.doi.org/10.1029/2005PA001153Google Scholar
Nakatsuka, T., Watanabe, K., Handa, N., Matsumoto, E., and Wada, E. Glacial to interglacial surface nutrient variations of Bering deep basins recorded by δ13C and δ15N of sedimentary organic matter. Paleoceanography 10, (1995). 10471061.Google Scholar
Okazaki, Y., Takahashi, K., Asahi, H., Katsuki, K., Hori, J., Yasuda, H., Sagawa, Y., and Tokuyama, H. Productivity changes in the Bering Sea during the late Quaternary. Deep-Sea Research Part II 52, (2005). 21502162.Google Scholar
Okkonen, S.R., Schmidt, G.M., Cokelet, E.D., and Stabeno, P.J. Satellite and hydrographic observations of the Bering Sea ‘Green Belt’. Deep-Sea Research Part II 51, (2004). 10331051.Google Scholar
Raymo, M.E., Oppo, D.W., and Curry, W. The mid-Pleistocene climate transition: a deep sea carbon isotopic perspective. Paleoceanography 12, (1997). 546559.Google Scholar
Rea, D.K., and Snoeckx, H. Sediment fluxes in the Gulf of Alaska: paleoceanographic record from Site 887 on the Patton-Murray seamount platform. Rea, D.K., Basov, I.A., Scholl, D.W., Allan, J.F. Proc. ODP Sci. Results 145, (1995). Ocean Drilling Program, College Station, TX. 247256.Google Scholar
Schefus, E., Sinninghe Damste, J.S.S., and Jansen, J.H.F. Forcing of tropical Atlantic sea surface temperatures during the mid-Pleistocene transition. Paleoceanography 19, (2004). PA4029 http://dx.doi.org/10.1029/2003PA000892CrossRefGoogle Scholar
Shackleton, N.J., and Hall, M.A. Stable isotope history of the Pleistocene at ODP Site 677. Becker, K., Sakai, H. et al. Proc. ODP Sci. Results 111, (1989). Ocean Drilling Program, College Station, TX. 295316.Google Scholar
Shackleton, N.J., Imbrie, J., and Pisias, N.G. The evolution of oceanic oxygen-isotope variability in the North Atlantic over the past 3 million years. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 318, (1988). 679686.Google Scholar
Snoeckx, H., Rea, D.K., Jones, C.E., and Ingram, B.L. Eolian and silica deposition in the central North Pacific: results from Sites 885/886. Rea, D.K., Basov, I.A., Scholl, D.W., Allan, J.F. Proc. ODP Sci. Results 145, (1995). Ocean Drilling Program, College Station, TX. 219230.Google Scholar
Sorokin, Y.I. Data on primary production in the Bering Sea and adjacent Northern Pacific. Journal of Plankton Research 32, (1999). 615636.Google Scholar
Springer, A.M., McRoy, P.C., and Flint, M.V. The Bering Sea Green Belt: shelf-edge processes and ecosystem production. Fisheries Oceanography 5, (1996). 205223.Google Scholar
St. John, K.E.K., and Krissek, L.A. Regional patterns of Pleistocene ice-rafted debris flux in the North Pacific. Paleoceanography 14, 5 (1999). 653662. http://dx.doi.org/10.1029/1999PA900030Google Scholar
Stabeno, P.J., Bond, N.A., Hermann, A.J., Kachel, N.B., Mordy, C.W., and Overland, J.E. Mateorology and oceanography of the northern Gulf of Alaska. Continental Shelf Research 24, (2004). 859897. http://dx.doi.org/10.1016/j.csr.2004.02.007Google Scholar
Takahashi, K., Ravelo, A.C., Zarikian, C.A., IODP Expedition 323 Scientists, IODP Expedition 323—Pliocene and Pleistocene paleoceanographic changes in the Bering Sea. Scientific Drilling 11, (2011). 413. http://dx.doi.org/10.2204/iodp.sd.11.01.2011Google Scholar
Teraishi, A., Suto, I., Onodera, J., and Takahashi, K. Diatom, silicoflagellate and ebridian biostratigraphy and paleoceanography in IODP 323 Hole U1343E at the Bering slope site. Deep-Sea Research Part II (2013). (in press) Google Scholar
Tsuda, A., Kiyosawa, H., Kuwata, A., Mochizuki, M., Shiga, N., Saito, H., Chiba, S., Imai, K., Nishioka, J., and Ono, T. Responses of diatoms to iron-enrichment (SEEDS) in the western subarctic Pacific, temporal and special comparisons. Progress in Oceanography 64, (2005). 189205. http://dx.doi.org/10.1016/j.pocean.2005.02.008Google Scholar