Quaternary Science Reviews 82 (2013) 78e92 Contents lists available at ScienceDirect Quaternary Science Reviews journal homepage: www.elsevier.com/locate/quascirev Biological and climatic consequences of a cold, stratified, high latitude ocean James D. Hays a, *, Douglas G. Martinson b, Joseph J. Morley c a Columbia University, Lamont Doherty Earth Observatory, P.O. Box 1000 Palisades, New York 10964, USA Lamont Doherty Earth Observatory, Dept of Earth and Environmental Sciences, Columbia University, USA c Lamont Doherty Earth Observatory, USA b a r t i c l e i n f o a b s t r a c t Article history: Received 16 August 2012 Received in revised form 16 September 2013 Accepted 24 September 2013 Available online 5 November 2013 The flux from deep- and shallow-living radiolarian assemblages provides evidence of a glacial, high latitude, cold ocean stratification that increased biological pump efficiency and promoted ocean carbon sequestration. Greater deep (>200 m) than shallow-living (<200 m) radiolarian assemblage flux characterizes glacial North Pacific (>45 N) sediments with the deep-living Cycladophora davisiana dominant (>24%). By contrast modern radiolarian flux consists primarily of shallow-living species (C. davisiana <10%). Clues to the cause of this unusual glacial radiolarian flux come from the presently, strongly stratified Sea of Okhotsk. Here beneath a thin nutrient depleted mixed layer radiolarian and zooplankton faunas conform to the sea’s physical stratification with lower concentrations of both in a Cold (1.5 to 1 C) Intermediate Layer (CIL) (20e125 m) and higher concentrations in waters between 200 and 500 m (Nimmergut and Abelmann, 2002). This biological stratification generates a radiolarian flux echoing that of the glacial northwest Pacific with C. davisiana 26% of total flux. Widespread C. davisiana percentages (>20%) in high latitude (>45 ) glacial sediments of both hemispheres is evidence that these oceans were capped with an Okhotsk-Like Stratification (O-LS). O-LS provides mechanisms to (1) strip nutrients from surface waters depriving the deep-ocean of preformed nutrients, increasing biological pump efficiency and (2) deepen carbon re-mineralization increasing deep-ocean alkalinity. Both may have contributed to lower glacial atmospheric CO2 concentrations. O-LS would also have amplified glacial climatic cycles by promoting the spread of high latitude sea ice in winter as occurs in the Sea of Okhotsk today, and reducing gas exchange between ocean and atmosphere in summer. Ó 2013 Elsevier Ltd. All rights reserved. Keywords: Biological pump Radiolaria North Pacific Deep sea cores Ocean carbon sequestration Flux Glacial ocean Stratification Pleistocene Holocene Sea of Okhotsk 1. Introduction Positive correlations between temperature and CO2 concentration in ice cores have led to a consensus that atmospheric CO2 variations amplified ice age climate change (Berner et al., 1979; Petit et al., 1999) but the cause of these variations are poorly understood (Sigman et al., 2010). The partial pressure of atmospheric CO2 (pCO2) is controlled by the average steady state of ocean surface water pCO2 weighted by area and gas exchange kinetics (Archer et al., 2000), which in turn varies with temperature and major nutrient concentrations. Much of the present-day ocean’s mixed layer is stripped of major nutrients by an efficient biological pump; however, large areas of the Antarctic, equatorial Pacific and North Pacific are not. Increased glacial biological pump efficiency in these areas could cause more complete nutrient utilization and * Corresponding author. Tel.: þ1 845 351 2451. E-mail address: [email protected] (J.D. Hays). 0277-3791/$ e see front matter Ó 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.quascirev.2013.09.022 consequent atmospheric CO2 drawdown. The Antarctic Ocean is especially important, for today its unused nutrients are convected to the deep-ocean as so called preformed nutrients. Thus atmospheric pCO2 depends on both surface ocean nutrient and deepocean preformed nutrient concentrations. Deeper glacial carbon re-mineralization could also have lowered atmospheric pCO2 by triggering deep-ocean alkalinity changes (Broecker and Peng, 1987; Boyle, 1988). Increased glacial ocean carbon sequestration is suggested by lower glacial than Holocene deep-ocean oxygen levels (Thompson et al., 1990; Francois et al., 1997; Gebhardt et al., 2008; Jaccard et al., 2009) and more depleted glacial d13C below 2000 m than above (Herguera, 1992; Keigwin, 1998; Curry and Oppo, 2005). Proxies for productivity and export, e.g. opal and barium flux, to subarctic North Pacific sediments do not support increased glacial relative to Holocene primary productivity (Narita et al., 2002; Kienast et al., 2004; Haug et al., 2005; Jaccard et al., 2005; Shigemitsu et al., 2007). So lower North Pacific diatom d15N (increased surface water nitrogen utilization) values in glacial relative to interglacial J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 79 Fig. 1. Recent distribution of northern he misphere permafrost (Washburn, 1980 and references therein) and percent C. davisiana in ocean floor sediments (Morley and Hays, 1983; Morley, 1983; and Pisias personal communication). Holocene high percentages of this species occur only within the Sea of Okhotsk and adjacent Pacific. sediments, have been attributed to lower upward nutrient flux caused by increased stratification, rather than higher primary productivity (Sigman et al., 1999; Galbraith et al., 2008). Glacial stratification has also been invoked in the Antarctic (Francois et al., 1997), however the biological consequences of this stratification are poorly understood in either the Antarctic or North Pacific. Although organic carbon export is often linked to primary productivity (Eppley and Peterson, 1979), it can also vary with little or no primary production change (Boyd and Newton, 1995) because the quantity of sinking organic carbon is strongly affected by consumer community structure (Michaels and Silver, 1988) that can be influenced by physical water column properties (Hargrave, 1975; Gardner et al., 1993). Glacial Ocean cooling has been suggested as a water column change that could reduce heterotrophic consumption rates and thereby increase biological pump efficiency relative to today (Matsumoto, 2007). The remains of shallow (<200 m) and deep-living (>200 m) organisms can provide information about past depths of organic carbon respiration but few members of the deep-living community leave a fossil record. Of those that do, polycystine radiolarians are the most abundant and diverse group (Kling, 1979; Gowing, 1986; Takahashi, 1991; Abelmann and Gowing, 1997; Itaki, 2003; Okazaki et al., 2004; Abelmann and Nimmergut, 2005; Tanaka and Takahashi, 2005). Radiolarians, as trophic generalists, feeding on sinking organic detritus and associated microorganisms including bacteria (Anderson, 1983; Gowing, 1986), respond to changing organic flux. Because their shells suffer less from dissolution than diatoms (Shemesh et al., 1989; Morley et al., in press) fossil radiolarians should record production changes at different levels within the water column. Shallow-living species (0e200 m) dominate radiolarians captured by plankton tows (Kling and Boltovskoy, 1995; Tanaka and Takahashi, 2008) but a deep-living species, Cycladophora davisiana, dominates Last Glacial Maximum (LGM) high latitude (>45 ) radiolarian faunas of both hemispheres (Hays et al., 1976; Morley et al., 1982; Morley, 1983; Gersonde et al., 2003). In Holocene sediments only those of the Sea of Okhotsk approach the high C. davisiana percentages found in high latitude glacial ocean sediments (Fig. 1). In the Okhotsk water column radiolarian concentrations below 200 m exceed those above, reaching a maximum between 200 and 500 m, where C. davisiana dominates (Nimmergut and Abelmann, 2002; Okazaki et al., 2004). This 80 J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 radiolarian concentration profile is well documented, but its cause remains controversial. Some attribute high C. davisiana productivity between 200 and 500 m to organic matter advected from neighboring shelves (Okazaki et al., 2003a) or transported by sea ice (Okazaki et al., 2003b). Others appeal to favorable intermediate water properties to enhance production and a possible influence from overlying stratification (Abelmann and Nimmergut, 2005). Deep-living radiolarians (>200 m) may also benefit from enhanced carbon export resulting from reduced respiration in the cold waters between 20 and 125 m (Hays and Morley, 2004). Views also diverge on the cause of high C. davisiana percentages (>20%) in high-latitude glacial sediments. A popular proposal is that they are a response to well-ventilated Intermediate Water flow (Nimmergut and Abelmann, 2002; Ohkushi et al., 2003; Abelmann and Nimmergut, 2005; Matul, 2011). Alternatively it has been argued they are a product of a cold, highly stratified water column, similar to that of the modern Sea of Okhotsk (Morley and Hays, 1983; Hays and Morley, 2004). These proposals are testable because a glacial North Pacific water column, similar to that of the modern Sea of Okhotsk, would have produced lower shallow-living flux than its Holocene counterpart while C. davisiana production, promoted solely by increased intermediate water flow, should not have influenced shallow-living radiolarian production. This paper uses a set of radiolarian species that today live dominantly either above or below 200 m to test these hypotheses. It also documents for the first time glacial-interglacial changes in the flux of shallowand deep-living radiolarian assemblages and uses this information to explore the biological and climatic consequences of glacial North Pacific stratification. 2. Selection of material Three northwest Pacific cores are selected that have nearly constant late Pleistocene sedimentation rates (Ninkovich and Robertson, 1975; Keigwin, 1995) and those rates are sufficiently high to assure good opal preservation (Broecker and Peng, 1982). They include a hydraulic piston core, raised from atop Detroit Seamount (ODP site 883D) and two piston cores from the Pacific Ocean floor (V20-122, V20-124). Holocene sediments from a Sea of 60°N 55°N Sea of Okhotsk 883D 50°N V32-161 V20-122 V20-124 S Ja ea pa of n 45°N North Pacific 40°N 140°E 150°E 160°E Fig. 2. Location of four sediment cores used in this study, Sea of Okhotsk V32-161, 48.238 N, 149.066 E, 1600 m; deep northwest Pacific cores V20-122, 46.566 N, 161.683 E, 5563 m; V20-124, 45.833 N, 154.5 E, 5534 m; and Detroit Sea Mount 883D, 51.198 N, 167.678 E, 2396 m. Okhotsk core (V32-161) are also included (Fig. 2). No evidence of dissolution was observed except in Pleistocene sediments of the Sea of Okhotsk core. The three Pacific cores are sampled from late Pleistocene (about 70 ka BP) through Holocene but only the Holocene of the Sea of Okhotsk core (V32-161) because of poor Pleistocene radiolarian preservation (Morley et al., 1991). Species abundances, the degree to which species are restricted to depths above or below 200 m and their occurrence in both Sea of Okhotsk and northwest Pacific sediments, guided the selection of six species that are grouped into Shallow (<200 m) and Deep (>200 m) Assemblages. The morphological variations of these species are illustrated in references listed in Table 1. All other species are placed in an Other Assemblage. The Shallow Assemblage consists of Spongotrochus glacialis (Popofsky), Stylochlamydium venustum (Bailey) and their probable juveniles (Spongodiscidae spp). Stratified plankton tow studies show that these species live dominantly above 50 m in the Sea of Okhotsk (Nimmergut and Abelmann, 2002) and above 100 m in the Antarctic Ocean (Abelmann and Gowing, 1997). They compose >75% of all radiolarians separated from Recent Bering Sea surface sediments, raised from <150 m (Blueford, 1983; Wang et al., 2006). Blueford (1983) refers to these species as “groups” indicating considerable morphological variation, we also recognize wide morphological variation (Table 1). In stratified plankton tows S. venustum is found mostly above 100 m and S. glacialis mostly above 250 m in the North Pacific (Kling, 1979; Tanaka and Takahashi, 2008) and other high-latitude seas including the Arctic Ocean (Hülsemann, 1963; Tibbs, 1967; Kling, 1979; Morley and Stepien, 1985; Abelmann and Gowing, 1997; Itaki et al., 2003) but can occur in lower numbers down to 3000 m (Itaki et al., 2003; Tanaka and Takahashi, 2008). S. venustum is the most abundant radiolarian in Bering Sea and northwest Pacific Holocene sediments north of 45 (Ling et al., 1971; Robertson, 1975; Wang et al., 2006). Tanaka and Takahashi (2005) showed that its flux to Holocene sediments is greater than its flux to glacial sediments in both the Bering Sea and northwest Pacific. Both S. venustum and S. glacialis are residents of polar and sub-polar waters of both hemispheres Table 1 This table presents a list of references that together illustrates the morphological variability of the species included in this study. S. venustum, S. glacialis and their probable juveniles the Spongodiscidae spp. dominate the Shallow Assemblage and are well illustrated in numerous studies of North Pacific and Antarctic radiolarians. Similarly C. davisiana and D. hirundo dominate the Deep Assemblage and are likewise widely illustrated. A. micropora and Spongurus sp. are distinctive but much less abundant and if omitted would not change the conclusions of this paper. Acanthodesmia micropora (Popofsky), Kruglikova, 1975, p. 84, Fig. 2, (19); Nimmergut and Abelmann, 2002, p. 470, pl. 2, (7); Hays and Morley, 2004, p. 601, Fig. 7, (c). Cycladophora davisiana Ehrenberg, Petrushevskaya, 1967, p. 122, Fig. 69, (IeVII); Tanaka and Takahashi, 2008, p. 69, pl. 3, (7, 8); Hays and Morley, 2004, p. 601, Fig. 7, (g, h, i); Nimmergut and Abelmann, 2002, p. 469, pl. 1, (9, 10, 11). Dictyophimus hirundo (Haeckel), Petrushevskaya, 1967, p. 114, Fig. 67, (IeV); Tanaka and Takahashi, 2008, p. 71, pl. 5 (5, 6, 7); Ling et al., 1971, p. 729, pl. 2, (8, 9, 10); Nimmergut and Abelmann, 2002, p. 471, pl. 2 (5a, 5b). Spongodiscidae spp., Tanaka and Takahashi, 2008, p. 68, pl. 2, (5, 6, 7); Okazaki et al., 2003, p. 204, pl. I (18, 20). Spongotrochus glacialis (Popofsky), Nigrini and Moore, 1979, p. S117, pl. 15, (2); Tanaka and Takahashi, 2008, p. 68, pl. 2 (1, 2); Okazaki et al., 2003b, p.204, plate 1, (21). Spongurus (?) sp. Petrushevskaya, 1968, p. 49, Fig. 26, (I); Nigrini and Moore, 1979, p. S67, pl. 8, (4); Tanaka and Takahashi, 2008, p. 69, pl. 3 (3); Okazaki et al., 2003b, p. 204, pl. 1 (23); Ling et al., 1971, p. 727, pl. 1 (6). Stylochlamydium venustum (Bailey), Boltovskoy and Riedel, 1980, p. 141, pl. 4, (3); Kruglikova, 1975, p. 84, Fig. 2, (1, 2); Tanaka and Takahashi, 2008, p. 68, pl. 2 (3); Ling et al., 1971, p.727, pl. 1 (7, 8); Nimmergut and Abelmann, 2002, p. 469, pl. 1 (1). J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 3. Analytical procedures Radiolarians, from a known weight of dried sediment, were washed through a 63 mm screen and collected on microscope slides (Moore, 1973). At least 300 individuals were counted per slide. Sediment bulk density of cores V20-122 and V20-124 was estimated by drying (65 C, 4 h), weighing and measuring the volume, corrected for shrinkage, of a set of 10 rectangular blocks, cut from the studied segments of each core. Opaline silica, measured using methods of Mortlock and Froelich (1989), is negatively correlated with bulk density in these cores (r ¼ 0.9354, level of significance, a ¼ 104). That is, the Null Hypothesis (r ¼ 0) is rejected (so, r s 0) accepting a 0.01% chance that this rejection is wrong. We determine the significance by using the bootstrap. The bootstrap works by generating synthetic noise series that have the same spectral coloring as the opal data. To do that, we preserve the lowest order univariate (mean and variance) and bivariate (acvf d autocovariance) statistical moments of the opal series by computing its power spectrum (formally, the power spectral density, PSD) by Fourier transform of the acvf (the zero frequency of the PSD is the mean, integral of PSD the variance, and PSD itself is the acvf in the frequency domain). We then transform back to the time domain via the inverse transform using random phase and correlate each of these noise series to the bulk density series. We did this for 104 synthetic noise series. A histogram of r-values provides the probability of achieving the r-value we did achieve with the actual sampled opal series. The probability of getting that r-value (0.9354) by chance is 1 in 104. The bootstrap method preserves the important statistical moments including the autocovariance accounting for the effective degrees of freedom in the series. This relationship between opal concentration and bulk density is used to estimate bulk density where direct measurements are lacking. The bulk densities of sediments from Site 883D are from Rea et al. (1993). A 14C date of 5.4 ka BP at 3 cm controls the top age of Detroit Seamount core 883D and dates of 7.8, 9.3 and 11.2 ka BP at 14, 24 and 44 cm respectively control Holocene rates (Kiefer et al., 2001). Holocene rates interpolated between these 14C dates range between 4.5 and 10.5 cm/1000 years averaging 7.1 cm/1000 years. Late Pleistocene time control is provided by 14C dates of, 19.4 and 23.5 ka BP at 107 and 144 cm depths respectively, yielding bulk accumulation rate between these depths of 9.0 cm/1000 years. Older time control is provided by the Last Occurrence (LO) of L. nipponica Nakaseko sakai Morley and Nigrini at 310 cm with an age of 50 ka BP (Spencer-Cervato et al., 1993), and the MIS 4/5 boundary (75 ka BP) at 474 cm (Kiefer et al., 2001). These ages yield late Pleistocene bulk accumulation rates ranging between 6.4 and 10.5 cm/1000 years averaging 6.7 cm/1000 years similar to this core’s average Holocene rate (Fig. 3). The LO of the diatom species Proboscia curvirostris (Jouse’) Jordan and Priddle has been extensively dated in North Pacific sediments between 300 and 350 ka BP (Koizumi and Yamamoto, 2007; Barron and Gladenkov, 1995; Spencer-Cervato et al., 1993) with most dates closer to 300 ka BP. Keigwin (1995) correlated the Marine Isotope Stage (MIS) boundaries of site 883D with the orbitally tuned records of Martinson et al. (1987) and Shackleton et al. (1990). Using the resulting MIS stage boundary ages he showed that this Site has a nearly constant sedimentation rate of about 5 cm/1000 years back to the Brunhes/ Matuyama boundary (778 ka BP) (Shackleton et al., 1990; Tauxe et al., 1996). Using MIS 7/8 and 9/10 boundary ages we calculate an age of the LO of P. curvirostris in this site of 303,000 yr BP, close to the widely reported age of 300 ka BP which we will use in this study. The average sedimentation rate between the LO of P. curvirostris and the core top is about 6 cm/1000 years suggesting that the last half of the Brunhes chron at this site has a slightly higher average sedimentation rate than the earlier half. A set of sixteen ash layers, identified by their indices of refraction, occur in seven piston cores raised from the Pacific’s floor southwest of the Detroit Seamount and with each core having a nearly constant sedimentation rate (Ninkovich and Robertson, 1975). The ages of the ash layers younger than 300,000 yrs BP are 160 140 V20-122 883D V20-124 120 Age x 103 years (Kling, 1979; Morley and Stepien, 1985; Abelmann and Gowing, 1997; Tanaka and Takahashi, 2008) but similar forms have been reported from low latitudes (Renz, 1976; McMillen and Casey, 1978). The Deep Assemblage consists of four species that stratified plankton tow collections indicate have depth preferences below 200 m. The cosmopolitan C. davisiana Ehrenberg, is by far the most abundant member of this assemblage in glacial and Recent sediments. It has been found to be abundant below 150 m and most abundant between 200 and 500 m in Sea of Okhotsk stratified plankton tows (Nimmergut and Abelmann, 2002; Okazaki et al., 2004). Globally it lives primarily below 400 m (Abelmann and Gowing, 1997; Itaki, 2003; Itaki et al., 2003). Although some researchers recognize several subspecies of C. davisiana, we have chosen for this study to include all those described by Petrushevskaya (1967, Fig. 69) as C. davisiana and these have been shown to be dominantly deep dwellers (>200 m) in the Sea of Okhotsk (Nimmergut and Abelmann, 2002; Abelmann and Nimmergut, 2005). Dictyophimus hirundo (Haeckel) is cosmopolitan (Boltovskoy and Riedel, 1980) and most abundant in Okhotsk’s water column below 400 m (Nimmergut and Abelmann, 2002), and below 250 m in the subpolar North Pacific (Tanaka and Takahashi, 2008). Acanthodesmia micropora (Popofsky) was caught in deep 1061 m but rarely in shallow 258 m sediment traps in the central Sea of Okhotsk (Hays and Morley, 2004). Spongurus (?) sp. Petrushevskaya, a cosmopolitan species, occurs primarily below 400 m in Antarctic waters (Abelmann and Gowing, 1996) and predominantly between 1000 and 3000 m in the North Pacific (Tanaka and Takahashi, 2008). It has not been reported from the Sea of Okhotsk but is included here because of its deep-living Antarctic and North Pacific habitats. 81 100 80 60 40 20 0 0 100 200 300 400 500 600 700 800 Depth in cm. Fig. 3. Age models for three North Pacific cores. See Table 2 for estimated ages of volcanic ashes in V20-122, V20-124 and text for other age control. 82 J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 determined by interpolation between the core tops and the LO of P. curvirostris. For cores that do not reach the LO of P. curvirostris we use the interpolated age of the oldest ash layer for control. Two cores from this set (V20-122 and V20-124), that contain four of the 16 ash layers, are chosen to supplement data from core 883D. The similarity of the interpolated ages of each of these ash layers in the selected cores (bold faced type) and five nearby cores (Table 2) can result only if these cores have nearly constant sedimentation rates. The mean ages of the four ash layers (Table 2) together with the LO of L. nipponica sakaii (50 ka BP) (V20-122, 240 cm; V20-124, 320 cm) are used to control their Pleistocene sedimentation rates. This age control generates Pleistocene bulk sedimentation rates ranging between 5.0 and 6.2 cm/1000 yrs, averaging 5.5 cm/1000 yrs in V20-122 and between 3.8 and 7.5 cm/ 1000 yrs averaging 5.9 cm/1000 yrs in V20-124 (Fig. 3). The depth to the major deglacial Assemblage’ changes in the trigger cores of V20-122 and V20-124 is no greater than in their respective piston cores suggesting that little sediment is missing from the piston core tops but it is unlikely that these tops have zero age. Therefore Holocene Assemblage flux is calculated using zero core top ages and the midpoints of the Holocene/Pleistocene transitions of C. davisiana percentages (12.5 ka BP) (Fig. 5a and b) and by extrapolating late Pleistocene rates to the core tops. The former method yields Holocene bulk sedimentation rates of 2.4 (V20-122) and 3.2 cm/1000 yrs (V20-124) the latter 5.5 and 5.9 cm/ 1000 yrs respectively, similar to the Holocene rates of core 883D. These different estimated Holocene bulk accumulation rates alter the magnitude but not the direction of Holocene/Pleistocene Shallow and Deep Assemblage flux changes in both cores. 3.1. Analysis of flux errors FðXk Þ ¼ FðX1 ; X2 ; X3 Þ ¼ X1 X2 X3 (1) where X1 ¼ rads./gm.; X2 ¼ bulk density; X3 ¼ sed. rate. The uncertainty in the individual variables is computed as follows: Table 2 Depth to and estimated ages of, four ash layers (AL) in seven northwest Pacific cores. Cores used in this study are in boldfaced type. Ash layer ages are determined by linear interpolation, using core tops (zero age) and last occurrence (LO) of the diatom P. curvirostris at 300,000 yrs. BP (Barron and Gladenkov, 1995; Koitzumi and Yamamoto, 2007) for time control. For cores that do not reach the LO of P. curvirostris the interpolated age of the oldest ash layer is used instead. The similarity of individual ash layer ages in multiple cores is evidence these cores have nearly constant bulk accumulation rates. Ash layer (AL) Core Depth to AL base (cm.) 1 V20-122 V20-123 RC14106 V20-124 RC14106 V20-123 V20-124 RC14106 V20-119 V20-120 V20-121 V20-122 V20-123 212 215 204 44.5 42.7 45.3 44 1.3 372 287 63.6 63.8 63 0.6 315 653 539 62.6 111.7 119.8 119 490 520 770 844 132 136 137 145 150 5 8 3.1.2. Bulk density (X2) This variable is estimated by regressing grams of sediment per cm3 (bulk density) against grams of opal per grams of sediment (concentration) of ten samples. For this it is not the error in the slope we use, but rather the error in predicting the bulk density from opal concentration. This error is a function of opal concentration, so we estimate a single “representative” error in this prediction by using the error associated with an opal concentration that is two thirds of the way from the mean concentration to the highest. The highest opal concentration is large, so this estimate is conservative. 3.1.3. Sedimentation rate (X3) Here the deptheage plot of the ash layers in cores V20-122, V20124 and the 14C dates of 883D is fit with a least squares straight line (consistent with the assumption of a constant sedimentation rate as justified above). The uncertainty in the slope (the sedimentation rate) is estimated from calculating the standard error of the slope (Draper and Smith, 1988). Because of the nonlinearity in F, we make a first order linear approximation to its uncertainty (variance) by truncating a Taylor series expansion to first order. Applying the expectance operator, in the form of variance, gives: s2F z The flux (F) is a nonlinear multivariate product of 3 random variables (Xk, variables whose exact value we do not know): 2 3.1.1. Radiolarian counts (X1) Errors associated with how representative a single sample of sediment is, as well as counting errors. Three counts were made on each half of a split sample giving 6 estimates that are averaged to give a single estimate of uncertainty. Estimated age ka. Mean age 116 4 141 9 3 X 3 X vFðXi Þ vF Xj vX Cov Xi Xj vX i mi j i¼1 j¼1 (2) mj We assume that the 3 random variables are independent (the number of radiolarians is so small relative to the other sedimentary components (diatoms, clay and volcanic ash) that any relationship between X1 and the other variables is considered trivial). Thus the covariance term (Cov[]) in (2) is 0 whenever, i s j; it only survives when i ¼ j, reducing covariance to variance giving: s2F z 3 X vFðXi Þ vFðXi Þ Var½Xi vXi mi vXi mi i¼1 (3) or, expanded: s2F z½X2 ðmÞX3 ðmÞ2 Var½X1 þ ½X1 ðmÞX3 ðmÞ2 Var½X2 þ ½X1 ðmÞX2 ðmÞ2 Var½X3 (4) The uncertainties of each variable (as variance) are combined, as dictated by Formula (4) to estimate the standard error for each flux estimate yielding the mean fluxes, their standard error (to first order) and percent error of the mean flux, all in units of rads/ (cm2 ka) (Table 3). 4. Results In all three northwest Pacific cores the Deep Assemblage’s percentage of total radiolarians declines during the deglacial (19 ka BP to 11.5 ka BP) from Pleistocene values of as much as 70% to Holocene values of about 10%. This decline parallels declines of C. davisiana percentages and d18O of Uvigerina sp. in core 883D (Figs. 4a and 5a and b). Nearly simultaneously the Shallow Assemblage rises from about 10% in the late Pleistocene to between 20 and 50% in the Holocene (Figs. 4a and 5a and b). The fact that a J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 5.5 Holocene Holocene % Shallow % Deep δ O Uvigerina sp. % C. davisiana a 70 25000 5.5 Late Pleistocene Late Pleistocene b Shallow flux Deep flux δ O Uvigerina sp. 5 5 20000 60 Percent O Uvigerina sp. o/oo 883D 50 4.5 40 4 4.5 15000 4 10000 18 30 883D 20 3.5 5000 3.5 Flux (radiolarians /cm2 1000 years) 80 83 10 0 3 0 10 20 30 40 50 60 0 3 70 0 10 20 30 40 50 60 70 3 Age x 10 year 3 Age x 10 year Fig. 4. Core 883D, percentages of Deep and Shallow Assemblages and C. davisiana together with d18O of Uvigerina sp. (a) and Deep and Shallow Assemblage flux (b). The Deep Assemblage constitutes greater than 50% of late Pleistocene total radiolarian flux while the Shallow Assemblage constitutes less than 10%. A reversal of Assemblage dominance occurs mostly during the deglacial, indicated by vertical lines between 19 and 11.5 ka BP in this figure and Figs. 5 and 6. The Deep Assemblage and C. davisiana percentages follow variations of d18O of Uvigerina sp. 70 70 a Percernt 60 Late Pleistocene Shallow Assemblage % Deep Assemblage % C. davisiana % Holocene Late Pleistocene b V20-122 60 V20-124 50 50 40 40 30 30 20 20 10 10 Percent Holocene 0 0 0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70 3 Age x 10 years Fig. 5. Deep and Shallow Assemblage and C. davisiana percentages of total radiolarian flux for cores V20-122 (a) and V20-124 (b) showing reversals of dominance mostly during the deglacial. deep radiolarian assemblage (>200 m) constitutes between 40 and 70% of the glacial radiolarian fauna in these three cores is noteworthy for nowhere in the modern ocean, or in Holocene sediments, have such high percentages of deep-living species been Table 3 Mean flux (rads/(cm2 ka) of Deep and Shallow Assemblages their standard error and the percent error of the mean flux. Three counts were made on each of two halves of a split sample taken from MIS-2 sediments. 833D V20-122 V20-124 Shallow flux 3389 298 (9%) 2768 197 (7%) 2917 304 (10%) Deep flux 12,082 1059 (9%) 9864 681 (7%) 10,397 1083 (10%) reported. This remarkable phenomenon and its cause is the focus of this paper. The major flux trends of these Assemblages and their magnitudes are similar in all three cores (Figs. 4b and 6a and b) and include a rise of Deep Assemblage flux from 70 ka BP to 19 ka BP, followed by its abrupt decrease during the deglacial. This Deep Assemblage flux is accompanied by low late Pleistocene Shallow Assemblage flux followed by an abrupt increase during the deglacial. This latter increase starting about 14 ka BP in 883D, a little earlier in V20-122 but significantly later in V20-124, lags the decrease of the Deep Assemblage. The later rise of Shallow Assemblage flux in Pacific waters near the Sea of Okhotsk (V20124) relative to this rise in open Pacific waters to the east is similar 84 J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 25000 25000 Late Pleistocene Flux (radiolarians cm2, 1000 years) a Late Pleistocene b Shallow flux (A) Shallow flux (B) Deep flux (A) V20-122 20000 Holocene V20-124 20000 Deep flux (B) 15000 15000 10000 10000 5000 5000 0 0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70 Flux (radiolarians /cm2, 1000 years) Holocene 0 3 Age x 10 years Fig. 6. Holocene and late Pleistocene Shallow and Deep Assemblage flux show similar trends in V20-122 (a) and V20-124 (b). Holocene flux is calculated (1) by using zero core top ages solid lines (A) and (2) extrapolating Pleistocene rates to the core tops, dashed lines (B). to that reported for opal flux in these areas (Kohfeld and Chase, 2011). Calculating the Holocene Deep and Shallow Assemblage flux of V20-122 and V20-124 by: (1) assuming zero core top ages (Shallow flux A, Deep Flux A, Fig. 6a and b) or (2) extrapolating Pleistocene sedimentation rates through the Holocene (Shallow flux B, Deep flux B, Fig. 6a and b), alters the magnitude but not the direction of these deglacial trends. The lower Holocene flux of V20-124 than V20-122 (Fig. 6) is caused by lower radiolarian concentrations not lower sedimentation rates. The ranges of Total Assemblage flux in these three cores are similar to the ranges reported by Tanaka and Takahashi (2005) from piston core ES (49 450 N, 168 190 E). C. davisiana is the major component of the Deep Assemblage and its variations dominate temporal changes of this Assemblage. The time series of the other Deep Assemblage species correlate positively with each other and with C. davisiana indicating a common response of all Deep Assemblage species to environmental change (Table 4). Deep and Shallow Assemblage flux are negatively Table 4 Positive correlations between the time series of all of deep-living species indicate a common response to environmental change. The positive correlations between Shallow and Other Assemblages suggests Shallow living species may dominate the Other Assemblage. The Shallow and Deep Assemblages are negatively correlated. Core Species and assemblages Correlation coefficient r p 883D C. davisiana e D. hirundo C. davisiana e A. micropora C. davisiana e Spongurus sp. Deep and Shallow Shallow and Other C. davisiana e D. hirundo C. davisiana e A. micropora C. davisiana e Spongurus sp. Deep and Shallow Shallow and Other C. davisiana e D. hirundo C. davisiana e A. microporav20-122 C. davisiana e Spongurus sp. Deep and Shallow Shallow and Other 0.7466 0.6734 0.3848 0.7318 0.2721 0.2171 0.5891 0.4139 0.7091 0.7797 0.7673 0.6205 0.5643 0.1134 0.6311 <0.0001 <0.0001 0.0053 <0.0001 0.0534 <0.152 <0.0001 <0.0047 <0.0001 <0.0001 <0.0001 <0.0001 0.0002 0.4862 <0.0001 V20-122 V20-124 correlated while a generally positive correlation between the Other and Shallow Assemblages suggests shallow-living species are important in the Other Assemblage (Table 4). These radiolarian Assemblage fluxes show no response to the inferred productivity and d15N maximum that peak around 13 ka BP during the Bolling-Allerod (Keigwin et al., 1992; Ternois et al., 2001; Sato et al., 2002; Crusius et al., 2004; Seki et al., 2004; Brunelle et al., 2007; Kohfeld and Chase, 2011). Higher resolution sampling of these radiolarian Assemblages during the deglacial is needed to more thoroughly investigate these relationships. To compare late Pleistocene and Holocene Assemblage flux and Assemblage radiolarians/gm. among Pacific and Sea of Okhotsk cores, average Shallow, Deep, Other and Total Assemblage flux and their radiolarians/gm. are calculated for MIS-2 (17.5e24.1 ka BP) and Holocene (0e11.5 ka BP) sediments (Table 5). Holocene Assemblage fluxes for V20-122 and V20-124 are calculated using both the low and high bulk accumulation rates discussed above. The Assemblage flux and their concentrations are more consistent among cores during MIS-2 than during the Holocene. This may be due to fewer Holocene than MIS-2 samples, core top age uncertainties or the glacial northwest Pacific was geographically more uniform than its Holocene counterpart. When using the higher Holocene bulk sedimentation rates for V20-122 and V20-124 Shallow Assemblage radiolarians/gm and flux increase between MIS-2 and Holocene by from four times in V20-124 to over 10 times in V20-122, Other Assemblage flux also increases between MIS-2 and Holocene in all three cores while average Holocene Deep Assemblage flux falls to less than half its MIS-2 values in all three cores (Table 5). These large changes of average Shallow and Deep Assemblage flux between MIS-2 and Holocene that occur in opposite directions in all three cores cannot be a result of sedimentation rate errors or changes of lithic component accumulations. Rather they must be caused by changes in the flux to the sea floor of the species components of the Shallow and Deep Assemblages. Average Deep Assemblage flux accounts for more than 45% of MIS-2 Total flux in the three Pacific cores but only 10% of Holocene Total flux while the Shallow Assemblage constitutes more than 18% of their Total Holocene flux but less than 10% of MIS-2 total flux. A scatter plot of these relationships (Fig. 7) produces a cluster of MIS- J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 85 Table 5 Average Shallow, Deep, Other and Total Assemblage flux (radiolarians/cm2 1000 years) and radiolarians/gm. of sediment (in bold italicized numerals) for Holocene (0e 11.5 ka BP) and MIS-2 (17.5e24.1 ka BP) sediments of northwest Pacific and Holocene Sea of Okhotsk cores (V32-161). Holocene flux in cores V20-122 and V20-124 is calculated using bulk sedimentation rates based on zero core top ages (numbers in parentheses) and also using late Pleistocene bulk accumulation rates (numbers in boldfaced type). Core Holocene Shallow Deep Other Total Shallow Deep- Other Total V20-122 (6063) 14,404 3119 11,578 2839 (3066) 7089 1036 1693 357 (2452) 5823 1346 4190 958 (2003) 4631 664 4564 965 (17,849) 42,391 9679 15,773 4512 (10,819) 25,020 3615 6057 1315 (26,364) 62,615 14,144 31,541 8309 (15,888) 36,741 5315 12,315 2637 1068 176 13,589 2236 20,388 3351 35,045 5762 1505 335 1631 220 11,785 2637 10,964 1490 13,238 3000 10,419 1421 26,528 5972 23,016 3132 883D V20-124 V32-161 MIS 2 2 values well separated from a looser cluster of Holocene values. The Sea of Okhotsk Holocene values plot between but closer to the MIS-2 cluster. Similar results occur regardless of whether the high or low Holocene sedimentation rates, discussed above, are used for cores V20-122 and V20-124. These data provide evidence that the water column concentration profiles of the Deep and Shallow Assemblages in the Sea of Okhotsk during the Holocene were more similar to those of the northwest Pacific during MIS-2 than to those of the northwest Pacific during the Holocene. It should be noted that while Okhotsk Holocene Shallow Assemblage flux is similar to that of open Pacific MIS-2 Shallow Assemblage flux, Okhotsk Holocene Deep, Other and Total Assemblage flux are significantly less than their open Pacific MIS-2 counterparts (Table 5) making the Sea of Okhotsk’s Deep Assemblage flux intermediate between the northwest Pacific’s Holocene and MIS-2 Deep Assemblage flux. This is surprising considering the present high productivity of the modern Sea of Okhotsk and the lower productivity that probably characterized the glacial North Pacific. 40 Shallow Assemblage flux as % of total flux 883D 35 Holocene MIS-2 30 C. davisiana dominates Deep Assemblage flux, in the three northwest Pacific cores, in both MIS-2 (average 63%, range 53e63%) and Holocene (average 64%, range 53e75%) sediments. The decline of C. davisana percentages of total flux between MIS-2 and Holocene is caused by both a reduction of C. davisiana flux and an increase of Shallow and Other Assemblage flux. Similarly higher C. davisiana percentages in Sea of Okhotsk Holocene than northwest Pacific Holocene sediments, are also caused by higher C. davisiana flux and lower Shallow and Other Assemblage flux in the former than the latter (Table 5). From these data the following is inferred: (1) Similar major trends of Deep and Shallow radiolarian Assemblage flux in three northwest Pacific cores indicate that they are robust and have regional significance. (2) Deep-living radiolarian species dominated total radiolarian flux to Lateglacial northwest Pacific sediments and were more dominant in Sea of Okhotsk Holocene than in northwest Pacific Holocene flux. (3) These data are consistent with the hypothesis that the water column concentration profiles of the Deep and Shallow Assemblages in the glacial northwest Pacific were more similar to those of the Holocene Sea of Okhotsk than to those of the Holocene northwest Pacific. (4) The major shoaling of northwest Pacific radiolarian production between MIS-2 and Holocene is an important biological response to deglacial oceanographic changes. 25 V20-122 5. The Sea of Okhotsk, a window on the high latitude glacial ocean 20 V20-124 V32-161 15 10 V20-124 5 883D V20-122 0 0 10 20 30 40 50 Deep Assemblage flux as % of total flux Fig. 7. Average Shallow and Deep Assemblage flux for MIS-2 and Holocene expressed as percentages of average Holocene and MIS-2 total radiolarian flux. Holocene and MIS-2 northwest Pacific values are clearly separated with Sea of Okhotsk Holocene values plotting near northwest Pacific MIS-2 values suggesting similarities between the environments of these seas. The Sea of Okhotsk’s biological stratification conforms to its physical stratification with lower concentrations of both radiolarians and zooplankton in a Cold (1.5 to 1 C) Intermediate Layer (CIL) (20e125 m) and higher concentrations of both in a subsurface maximum (Fig. 8, III) between 200 and 500 m (Gorbatenko, 1996; Nimmergut and Abelmann, 2002). The physical stratification is caused by extreme winter cooling (mean eastern Siberian January temperature 45 C (Jones et al., 1986)) and sea ice formation (Rogachev, 2000; Ogi et al., 2001). Little upward buoyancy flux can be induced in a column of near freezing seawater (Rooth, 1982; Winton, 1997; Sigman et al., 2004) so it is prone to stratify as its surface freshens (Warren, 1983; Winton, 1997). The springesummer pooling of sea ice melt water (51 cm/yr), precipitation and runoff (37 cm/yr) (Rogachev, 2000) produces a thin (10e20 m) low salinity mixed layer. Fall-winter mixing is limited to about 100 m 86 J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 Fig. 8. Annual temperature profile for Bering (panel I) and Okhotsk Seas (panel II). Panel III shows correspondence between low ocean temperatures and low radiolarian and zooplankton concentrations (a) Polycystine radiolarians, station LV28-3 (AugusteSeptember)(Nimmergut and Abelmann, 2002), (b) zooplankton, Kuril Kamchatka Trench (July) (Vinogradov, 1970), (c) nekton, Okhotsk Sea summer from Gorbatenko (1996) and (d) temperature (summer), CIL ¼ Cold Intermediate Layer, OUIW ¼ Okhotsk Upper Intermediate Water, OLIW ¼ Okhotsk Lower Intermediate Water, after Hays and Morley (2004). forming a permanent Dichothermal or Cold Intermediate Layer (CIL) between about 20 and 125 m with near zero to subzero temperatures (Moroshkin, 1968; Kitani, 1973) (Fig. 8, panel II). Winter sea ice formation and brine rejection in polynyas at the sea’s northern edge produces dense water that sinks to between150 and 450 m, mixing with Okhotsk Lower Intermediate Water (OLIW), of Pacific origin, to form Okhotsk Upper Intermediate Water (OUIW) (Fig. 8 panel III) (Gladyshev, 1998; Martin et al., 1998). The low radiolarian concentrations above 200 m have been attributed to strong seasonal contrasts, low surface water salinities and sea ice cover (Nimmergut and Abelmann, 2002; Okazaki et al., 2003b; Tanaka and Takahashi, 2005). It is likely that cold CIL temperatures also limit radiolarian production within the CIL, as has been proposed for Okhotsk zooplankton (Vinogradov, 1970), for metabolic and reproductive rates decline exponentially with falling temperature (Arrhenius, 1915; Nelson, 2004). Low numbers of heterotrophs, with low respiration rates, living within the CIL should also enhance export through it, benefiting underlying zooplankton and radiolarians, including C. davisiana (Fig. 8, Panel III). By contrast a warmer subsurface temperature minimum in the central Bering Sea (Fig. 8, Panel I) is associated with low C. davisiana percentages in underlying Holocene sediments (Fig. 1). The CIL’s minimum temperature (Tmin) is inversely related to the stability of the stratification for stability reduces summer mixing preserving low winter temperatures (Kitani, 1973). C. davisiana percentages, in the summer water column between 200 and 500 m, J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 are inversely related to the CIL’s Tmin (Fig. 9) suggesting that C. davisiana is more sensitive to CIL temperatures and water column stability than other radiolarians living in that depth range. C. davisiana percentages in Okhotsk’s Holocene sediments are also inversely related to the CIL’s Tmin (Fig. 9) connecting these percentages to the CIL’s low temperatures and stable physical stratification. Okhotsk’s cold subsurface water, while probably important, is not sufficient to produce high sedimentary C. davisiana percentages, for water as cold or colder than the Okhotsk CIL underlies the mixed layer in Antarctic waters today but C. davisiana is generally less than 5% of Antarctic Holocene radiolarians (Hays et al., 1976; Gersonde et al., 2003). The thicker Antarctic mixed layer may be an important difference because in the Okhotsk Sea Shallow Assemblage species are largely restricted to its upper 50 m (Nimmergut and Abelmann, 2002) while in the open North Pacific, where mixed layers are thicker, these same species range through the upper 100e200 m (Kling, 1979; Tanaka and Takahashi, 2008). Thin mixed layers may play an important biological role in cold stratified oceans for they not only limit the space within which shallow-living species thrive, they also enhance nutrient utilization (Sverdrup, 1953) and carbon export, the latter because thinner mixed layers increase the probability of organic particles sinking through their lower boundaries (Kerr and Kuiper, 1997). A summer/fall (Sept., Oct., Nov.) C. davisiana flux maxima accounts for most of C. davisiana’s annual flux to a central Okhotsk Sea (1060 m) sediment trap (Hays and Morley, 2004) and coincides in this trap with an annual organic carbon flux maximum of 7 mg/m2/ day (Honda et al., 1997). Similar North Pacific and Bering Sea coincident summer/fall C. davisiana and organic carbon flux maxima (Takahashi, 1995), suggest C davisiana production today is controlled by seasonal biological rhythms of the North Pacific Ocean and its marginal seas, i.e. organic carbon export, with high bacterial numbers, during the late summer heterotrophic phase of % C. davisiana vs Okhotsk CIL Tmin 35 % C. davisiana core tops % C. davisiana 200-500m y = 16.735 - 4.9484x r = 0.7612 y = 16.668 - 6.2396x r = 0.65443 30 % C. davisiana 25 20 15 10 5 0 -2.00 -1.00 0.00 1.00 2.00 3.00 CIL Tmin C° Fig. 9. Sea of Okhotsk C. davisiana percentages versus Tmin of CIL, in summer water column (200e500 m) (circles), error bars on water column data points represent a 15% uncertainty that all specimens counted were living (Abelmann and Nimmergut, 2005), Tmin from Biebow et al. (2003). C. davisiana percentages in Holocene sediments (diamonds) after Morley and Hays (1983), Tmin from CTD and XBT stations each within 80 nautical miles (mean distance of 22 nautical miles) of associated core site (NODC data base). 87 the plankton cycle (Sorokin and Sorokin, 1999; Nimmergut and Abelmann, 2002) not factors peculiar to the Sea of Okhotsk. The high annual organic carbon export to this central Okhotsk Sea trap of 1.7 g/m2/yr, combined with an organic to inorganic carbon ratio of 3.4, and a Ca/Si ratio of 0.1 led Honjo and Manganini (1996) to suggest that this sea is a strong CO2 sink. The biological utilization of nutrients as they are made available and the consequent stripping of nutrients from the summer mixed layer (Sorokin, 2002) indicates this sea has an efficient biological pump. In summary, the greater radiolarian flux from below than above 200 m in the Sea of Okhotsk is unusual in modern seas but similar to that of the glacial northwest Pacific. Both have higher Deep than Shallow Assemblage flux and C. davisiana percentages >20%. These data suggest that the glacial northwest Pacific had a physicale biological stratification similar to that of the Holocene Sea of Okhotsk. It should be noted that C. davisiana percentages, of up to 18%, occur in Recent sediments under upwelling areas (Robson, 1983; Welling et al., 1992; Jacot Des Combes and Abelmann, 2007) and other unstratified seas (Itaki, 2003). The cause of these high percentages is not known and needs to be investigated. 6. The glacial ocean viewed through the Sea of Okhotsk window Physical factors must have constrained Shallow Assemblage production in the northwest Pacific during glacial relative to Holocene times because the organic matter that fueled higher Deep Assemblage production was first available to, but unused by, the Shallow Assemblage. If well-ventilated intermediate water (Nimmergut and Abelmann, 2002) or intermediate water charged with organic matter (Okazaki et al., 2003a) benefited C. davisiana it should not have been detrimental to overlying shallow-living species. Also if sea ice transported organic matter (Okazaki et al., 2003b) benefited either it should have promoted shallow as well as deep-living species production. The glacial northwest Pacific’s Shallow and Deep Assemblage flux can best be explained by the presence of a thin mixed layer and an underlying CIL, similar to that of the Sea of Okhotsk today. Such a water structure would have inhibited Shallow and promoted Deep Assemblage production. The disappearance of this Okhotsk-Like Stratification (O-LS) during the deglacial, with resulting mixed layer thickening and warming of the CIL, could explain the decline of the Deep Assemblage, the rise of the Shallow Assemblage and the resulting fall of C. davisiana percentages. The fact that Sea of Okhotsk Deep Assemblage flux (Table 5) and Deep and Shallow Assemblage flux as a percent of total flux (Fig. 7) are intermediate between those of the northwest Pacific’s Holocene and MIS-2 suggests that northwest Pacific MIS-2 conditions were more severe than Holocene Sea of Okhotsk conditions e.g. colder CIL and more efficient biological pump. The analogy between the glacial northwest Pacific and Holocene Okhotsk Sea is marred by the fact that the Holocene Sea of Okhotsk was highly productive while various proxies suggest the glacial northwest Pacific was less so (Narita et al., 2002; Kienast et al., 2004; Seki et al., 2004; Jaccard et al., 2005; Okazaki et al., 2005; Galbraith et al., 2008). Possible causes of lower glacial than Holocene North Pacific productivity have been much discussed (for a review see Kohfeld and Chase, 2011). A popular suggestion calls for increased glacial stratification to reduce nutrient input to the mixed layer (Sarnthein et al., 2004; Sigman et al., 2004; Jaccard et al., 2005; Brunelle et al., 2007, 2010). Yet the Sea of Okhotsk’s strong stratification does not prevent Ekman pumping of nutrient rich thermocline waters into its thin mixed layer causing high seasonal productivity that fully utilizes nutrients as they are provided (Sorokin, 2002). So lower North Pacific diatom d15N values in glacial 88 J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 Fig. 10. LGM C. davisiana percentages suggest Okhotsk-Like Stratification existed at that time north of the present position of the Pacific Polar Front (Belkin et al., 2002). The geographical continuity between permafrost and O-LS suggests both were responses to cold winter air. Glacial permafrost distribution (Svendsen et al., 2004; Washburn, 1980 and references there in) and C. davisiana percentages; Atlantic data (Morley, 1983), Pacific data (Sachs, 1973; Robertson, 1975; Morley and Hays, 1983; Morley and Robinson, 1986; Morley et al., 1982; Ohkushi et al., 2003; Tanaka and Takahashi, 2005; N. Pisias personal communication and Morley personal communication). relative to interglacial sediments (Sigman et al., 1999; Galbraith et al., 2008) may not have been caused only by stratification limiting upward nutrient flux. Alternatively the glacial relative to Holocene supply of silica to the glacial North Pacific and Sea of Okhotsk thermocline may have been reduced (Gorbarenko et al., 2004; Seki et al., 2004; Okazaki et al., 2005; Kohfeld and Chase, 2011). Today mixing in the Kurile Island region brings nutrient rich water from the deep Pacific to and above North Pacific Intermediate Water (NPIW) contributing to Sea of Okhotsk and North Pacific productivity (Tsunogai, 2002; Sarmiento et al., 2004). Because this mixing occurs well below the surface, nutrient ratios are not reset by sea surface biological processes and retain high Si/N ratios. In glacial times the apparently nutrient depleted water mass above 2000 m (Herguera et al., 1992; Keigwin, 1998; Matsumoto et al., 2002) was probably ventilated in the far North Pacific (Duplessy et al., 1988; Keigwin, 1998) or Bering Sea (Matsumoto et al., 2002; Ohkushi et al., 2003; Gorbarenko et al., 2010; Horikawa et al., 2010). Because cooling the North Pacific today would further stabilize it (Warren, 1983), it is likely that in glacial time such ventilation occurred through brine rejection in high latitude polynyas as in the Sea of Okhotsk today (Gorbarenko et al., 2010). During ventilation sea surface biological activity would have reset upwelled nutrient ratios, depleting silica relative to nitrate as occurs in the Antarctic today (Sarmiento et al., 2004). Glacial North Pacific Intermediate water depleted of silica in a similar way could have lowered glacial relative to Holocene opal flux in the northwest Pacific and Sea of Okhotsk (Okazaki et al., 2005; Brunelle et al., 2007; Kohfeld and Chase, 2011). The deglacial initiation of deep mixing in the northwest Pacific should have caused a simultaneous Holocene increase in diatom production throughout the North J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 Pacific not just in the northwest Pacific and Sea of Okhotsk, yet this may not have been the case for the northeast Pacific (Kohfeld and Chase, 2011). Because the northwest Pacific and Sea of Okhotsk are beneficiaries of nutrients supplied directly by upwelled North Pacific Deep Water while the northeast Pacific receives nutrients via the NPIW, the latter’s history may be different (Tsunogai, 2002). The strong dissolution of diatom and radiolarian tests in the glacial sediments of core V32-161 indicates the deep waters of this sea were more under saturated in silica in glacial than Holocene times (Van Cappellen and Qui, 1997; Rickert et al., 2002). A similar degree of dissolution does not occur in northwest Pacific glacial sediments. In any event it is possible that the glacial northwest Pacific and Holocene Sea of Okhotsk could have had similar physical biological stratifications but differing silica supply mechanisms and consequent opal flux values. Cold winter air masses and resultant extensive sea ice cause Okhotsk’s physical stratification today. There is ample evidence of air colder than today’s eastern Siberian winter air across northern continents at the LGM (Atkinson et al., 1987; Cuffey et al., 1995; Isarin et al., 1998; Severinghaus et al., 1998; Isarin and Bohncke, 1999; Denton et al., 2005). North Pacific sea ice rafted detritus from this time, a consequence of cold winter air, extends to about 45 N (Conolly and Ewing, 1970; Kent et al., 1971). Conditions therefore were in place for O-LS to form and the C. davisiana percentages north of the 20% line in Fig. 10 are evidence that it did. This line follows the trend of today’s North Pacific Polar Front that separates water with a summer subsurface minimum to the north from water without or with a weakly developed subsurface minimum to the south (Belkin et al., 2002). For O-LS to develop north of the 20% line today would only require a thinning of the mixed layer and cooling the subsurface minimum (CIL). While summer temperatures play a major role in ice sheet growth and decay (Denton et al., 2005) cold winter temperatures and their duration control the development of permafrost (Nelson and Outcalt, 1987) and O-LS. The geographical continuity of glacial permafrost and O-LS (Fig. 10) is consistent with their mutual dependence on long cold winters. Modeling experiments suggest that reduced northern summer insolation, in response to orbital variations, cooled northern continents producing vegetative albedo feed-backs that promoted further cooling and associated cold winter air masses (Harrison et al., 1995; Gallimore and Kutzbach, 1996; Ganopolski et al., 1998; Notoro and Liu, 2007, 2008). This cold winter air could have engendered permafrost on land as well as O-LS on northern seas and the climatic feedbacks that accompany them. If so, O-LS formation and destruction could have been as abrupt as changes of the seasonal winter air that controls it. Widespread high C. davisiana percentages in Antarctic Ocean LGM sediments south of the Polar Front (Hays et al., 1976; Gersonde et al., 2003) result from a combination of higher C. davisiana flux and lower flux of other radiolarians (Hays et al., 1976) likely signaling the presence of O-LS in this region as well. Radiolarian assemblage data therefore supports inferences of North Pacific and Antarctic stratification based on increased glacial relative to interglacial surface water nitrate utilization (lower d15N) (Francois et al., 1997; Sigman et al., 1999; Brunelle et al., 2007, 2010; Galbraith et al., 2008). Phosphate (Elderfield and Rickaby, 2000) and silica data (De La Rocha et al., 1998) however suggest that the glacial Antarctic was not highly stratified. Low nitrate concentrations also characterize Okhotsk’s thin mixed layer (Sapozhnikov et al., 1999; Arzhanova and Naletova, 1999; Sorokin, 2002), so high latitude glacial O-LS obviates the need for iron fertilization (Martin, 1990) to lower glacial, relative to Holocene, surface water nitrate concentrations. Relative to the Holocene this glacial stratification would have deepened organic carbon re-mineralization by radiolarians 89 and other plankton (Fig. 8, panel III). This potential for O-LS to reduce shallow carbon re-mineralization combined with a general glacial ocean cooling (Matsumoto, 2007) provides a mechanism for high latitude oceans to efficiently transfer carbon to the deep ocean. Glacial high latitude oceans capped with O-LS had the potential to amplify glacial cycles by: 1) promoting the spread of sea ice, as occurs in the Sea of Okhotsk today where the CIL insulates sea ice from underlying warmer water allowing its spread to the shores of Hokkaido (45 N) (Bulgakov, 1965), a low-latitude extreme for modern sea ice, and a latitude similar to that of the southern limit of glacial North Pacific sea ice rafted detritus (Conolly and Ewing, 1970; Kent et al., 1971). Glacial sea ice expansion combined with strong summer stratification (Fig. 8, panel II) could have reduced glacial CO2 transfer from ocean to atmosphere (Stephens and Keeling, 2000) relative to today; 2) increasing high latitude biological pump efficiency through full utilization of surface water nutrients, depriving the deep-ocean of preformed nutrients and lowering atmospheric CO2 concentrations (Sigman et al., 2010); 3) triggering ocean alkalinity changes by deepening carbon remineralization further lowering atmospheric pCO2 (Broecker and Peng, 1987; Boyle, 1988). The deglacial transition from dominantly deep to dominantly shallow radiolarian production may reflect the change from deep to shallow organic carbon remineralization envisaged by Boyle (1988). If so it should lead alkalinity induced changes in atmospheric pCO2 by about 2500 years (Broecker and Peng, 1987). This model of the high latitude glacial ocean makes several testable predictions. The causal relationship between O-LS and sea ice predicts a consistent relationship between maximum winter sea ice extent and high sedimentary C. davisiana percentages, although their low latitude limits need not be the same. Opal flux peaks in Southern Ocean cores (Anderson et al., 2002) should be accompanied by shallow-living radiolarian flux peaks signaling reduced Southern Ocean stratification. The deglacial decline of C. davisiana percentages should coincide with or lead the rise of atmospheric pCO2. 7. Conclusions 1. The deepening of northwest Pacific glacial relative to Holocene radiolarian production records an important change of this oceans consumer community structure. A similar preference for deep-living radiolarians and zooplankton occurs in today’s Sea of Okhotsk where cold ocean stratification (O-LS) reduces shallow-relative to deep-living radiolarian production giving rise to the dominance of the deep-living C. davisiana. C. davisiana percentages (>20%) in high latitude (>45 ) LGM sediments of both hemispheres suggest these seas were capped with O-LS. 2. The physical and biological structure of glacial high latitude ocean stratification presented here has several potential climatic feedbacks; a) it promotes winter sea ice spread by insulating sea ice from warmer underlying water; b) its thin mixed layer fully utilizes surface water nutrients enhancing biological pump efficiency, and low respiration rates within the CIL deepen carbon re-mineralization, both of which could aid CO2 transfer from atmosphere to ocean in summer; c) the summer stratification may have reduced CO2 return to the atmosphere; 3. O-LS and its biological consequences rest on two fundamental properties of Earth’s climate system; a) the non-linearity of seawater’s equation of state that promotes cold ocean stratification; b) the exponential relationship between temperature and heterotrophic metabolic and reproductive rates that limit the rate of organic carbon re-mineralization within cold near surface waters (CILs). 90 J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 Acknowledgements We are grateful to Aaron Putnam, Dorothy Peteet, O. Roger Anderson, Michael Sarnthein, Taro Takahashi, Maureen Raymo and Wallace Broecker for reading the manuscript and offering useful suggestions. We are especially grateful to Ray Sambrotto, and Arnold Gordon for helpful discussions during manuscript preparation and for useful comments on multiple drafts of this paper. We are also grateful for the constructive suggestions of four anonymous reviewers. We appreciate the expert assistance of Maureen McAuliffe Anders during final manuscript preparation and are grateful to Nick Pisias for unpublished C. davisiana data. We thank the Ocean Drilling Program and its successor IODP for collecting and archiving the drill core used. The piston cores are archived at the Lamont-Doherty Earth Observatory of Columbia University whose core repository is supported by funds from National Science Foundation Grant OCE09-62010. This work was supported with funds through the Cooperative Institute for Climate Applications and Research (CICAR) under award number NA030AR4320179 from the National Oceanic and Atmospheric Administration, U.S. Department of Commerce. The statements, findings, conclusions, and recommendations are those of the authors and do not necessarily reflect the views of the National Oceanic and Atmospheric Administration or the Department of Commerce. This paper bears Lamont Doherty Earth Observatory contribution number 7735. References Abelmann, A., Gowing, M.M., 1996. Horizontal and vertical distribution pattern of living radiolarians along a transect from the Southern Ocean to the South Atlantic subtropical region. Deep-Sea Res. Part I 43, 361e382. Abelmann, A., Gowing, M.M., 1997. Spatial distribution pattern of living polycystine radiolarian taxa e baseline study for paleoenvironmental reconstructions in the Southern Ocean (Atlantic sector). Mar. Micropaleontol. 30, 3e28. Abelmann, A., Nimmergut, A., 2005. Radiolarians in the Sea of Okhotsk and their ecological implication for paleoenvironmental reconstructions. Deep-Sea Res. Part II 52, 2302e2331. Anderson, O.R., 1983. Radiolaria. Springer-Verlag, New York, p. 355. Anderson, R.F., Chase, Z., Fleisher, M.Q., Sachs, J., 2002. The southern Ocean’s biological pump during the last glacial maximum. Deep-Sea Res. Part II 49, 1909e 1938. Archer, D.E., Eshel, G., Winguth, A., Broecker, W., Pierrehumbert, R., Tobis, M., Jacob, R., 2000. Atmospheric pCO2 sensitivity to the biological pump in the ocean. Glob. Biogeochem. Cycles 14, 1219e1230. Arrhenius, S., 1915. Quantitative Laws in Biological Chemistry. G. Bell and Sons Ltd., London, p. 164. Arzhanova, N.V., Naletova, I.A., 1999. Hydrochemical structure, mesoscale eddies, and primary production in the northern part of the Sea of Okhotsk. Okeanology 39, 741e749. Atkinson, T.C., Briffa, K.R., Coope, G.R., 1987. Seasonal temperature in Britain during the last 22,000 years, reconstructed using beetle remains. Nature 325, 587e592. Barron, J.A., Gladenkov, Y.A., 1995. Early MioceneePleistocene diatom stratigraphy of leg 145. In: Rea, D.K., Basov, I.A., Scholl, D.W., Allan, J.F. (Eds.), Proc. Ocean Drilling Prog. Sci. Res. 145, pp. 3e19. Belkin, I., Krishfield, R., Honjo, S., 2002. Decadal variability of the North Pacific Polar Front: subsurface warming versus surface cooling. Geophys. Res. Lett. 29, 65-1e 65-4. Berner, W.H., Stouffer, B., Oeschger, H., 1979. Past atmospheric composition and climate, gas parameters measured on ice cores. Nature 275, 53e55. Biebow, N., Kulinich, R., Baranov, B.V., 2003. Cruise Reports: RV Akademik M.A. Lavrentyev Cruise 29, Leg 1 and Leg 2. GEOMAR Rept. 110, p. 190. Blueford, J.R., 1983. Distribution of quaternary radiolaria in the Navarin Basin geologic province, Bering Sea. Deep-Sea Res. Part A 30, 763e781. Boltovskoy, D., Riedel, W.R., 1980. Polycystine radiolaria from the southwestern Atlantic Ocean plankton. Rev. Esp. Micropaleontol. 12, 99e146. Boyd, P.W., Newton, P., 1995. Evidence of the potential influence of planktonic community structure on the interannual variability of particulate organic carbon flux. Deep-Sea Res. Part I 42, 619e639. Boyle, E.A., 1988. Vertical oceanic nutrient fractionation and glacial/interglacial CO2 cycles. Nature 331, 55e56. Broecker, W.S., Peng, T.-H., 1982. Tracers in the Sea. Eldigio Press, Lamont-Doherty Geological Observatory, Columbia University, Palisades, N. Y., p. 690 Broecker, W.S., Peng, T.-H., 1987. The role of CaCO3 compensation in the glacial to interglacial atmospheric CO2 change. Glob. Biogeochem. Cycles 1, 15e29. Brunelle, B.G., Sigman, D.M., Cook, M.S., Keigwin, L.D., Haug, G.H., Plessen, B., Schettler, G., Jaccard, S.L., 2007. 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, PA1215. http://dx.doi.org/ 10.1029/2005PA001205. Brunelle, B.G., Sigman, D.M., Jaccard, S.L., Keigwin, L.D., Plessen, B., Schettler, G., Cook, M.S., Haug, G.H., 2010. Glacial/interglacial changes in nutrient supply and stratification in the western subarctic North Pacific since the penultimate glacial maximum. Quat. Sci. Rev. 29, 2579e2590. Bulgakov, N.P., 1965. The extreme winter boundary of ice in far eastern seas ¼ Predel’naya zimniaya granitsa l’dov v dal’nevostochnykh moryakh. Nav. Oceanogr. Off. Transl. 292 (13), 66e76 (in Russian). Conolly, J.R., Ewing, M., 1970. Ice-rafted detritus in northwest Pacific deep-sea sediments. In: Hays, J.D. (Ed.), Geological Investigations of the North Pacific, Geol. Soc. Am. Mem. 126, Boulder, Colorado, pp. 219e232. Crusius, J., Pedersen, T.F., Kienast, S., Keigwin, L., Labeyrie, L., 2004. Influence of northwest Pacific productivity on North Pacific intermediate water oxygen concentrations during the Boiling-Allerød interval (14.7e12.9 ka). Geology 32, 633e636. Cuffey, K.M., Clow, G.D., Alley, R.B., Stuiver, M., Waddington, E.D., Saltus, R.W., 1995. Large Arctic temperature change at the Wisconsin-Holocene transition. Science 270, 455e458. Curry, W.B., Oppo, D.W., 2005. Glacial water mass geometry and the distribution of delta 13C of total CO2 in the western Atlantic Ocean. Paleoceanography 20, PA1017. http://dx.doi.org/10.1029/2004PA001021. De La Rocha, C.L., Brzezinski, M.A., DeNiro, M.J., Shemesh, A., 1998. Silicon-isotope composition of diatoms as an indicator of past oceanic change. Nature 395, 680e683. Denton, G.H., Alley, R.B., Comer, G.C., Broecker, W.S., 2005. The role of seasonality in abrupt climate change. Quat. Sci. Rev. 24, 1159e1182. Draper, N.R., Smith, H., 1988. Applied Regression Analysis, third ed. Wiley-Interscience, p. 736. Duplessy, J.C., Arnold, M., Shackleton, N.J., Kallel, N., Labeyrie, L., Juillet-Leclerc, A., 1988. Changes in the rate of ventilation of intermediate and deep-water masses in the Pacific Ocean during the last deglaciation. Chem. Geol. 70, 108. Elderfield, H., Rickaby, R., 2000. Oceanic Cd/P ratio and nutrient utilization in the glacial Southern Ocean. Nature 405, 305e310. Eppley, R.W., Peterson, B.J., 1979. Particulate organic matter flux and planktonic new production in the deep ocean. Nature 282, 677e680. Francois, R., Altabet, M.A., Yu, E.-F., Sigman, D.M., Bacon, M.P., Frank, M., Bohrmann, G., Bareille, G., Labeyrie, L.D., 1997. Contribution of Southern Ocean surface-water stratification to low atmospheric CO2 concentrations during the last glacial period. Nature 389, 929e935. Galbraith, E.D., Kienast, M., Jaccard, S.L., Pedersen, T.F., Brunelle, B.G., Sigman, D.M., Kiefer, T., 2008. Consistent relationship between global climate and surface nitrate utilization in the western subarctic Pacific throughout the last 500 ka. Paleoceanography 23, PA2212. http://dx.doi.org/10.1029/2007PA001518. Gallimore, R.G., Kutzbach, J.E., 1996. Role of orbitally induced changes in tundra area in the onset of glaciation. Nature 381, 503e505. Ganopolski, A., Kubatzki, C., Clausen, M., Brovkin, V., Petoukhov, V., 1998. The influence of vegetationeatmosphereeocean interaction on climate during the Mid-Holocene. Science 280, 1916e1919. Gardner, W.D., Walsh, I.D., Richardson, M.J., 1993. Biophysical forcing of particle production and distribution during a spring bloom in the North Atlantic. DeepSea Res. Part II 40, 171e195. Gebhardt, H., Sarnthein, M., Grootes, P.M., Kiefer, T., Kuehn, H., Schmieder, F., Röhl, U., 2008. Paleonutrient and productivity records from the subarctic North Pacific for Pleistocene glacial terminations I to V. Paleoceanography 23, PA4212. http://dx.doi.org/10.1029/2007PA001513. Gersonde, R., Abelmann, A., Brathauer, U., Becquey, S., Bianchi, C., Cortese, G., Grobe, H., Kuhn, G., Niebler, H.-S., Segl, M., Sieger, R., Zielinski, U., Fuetterer, D.K., 2003. Last glacial sea surface temperatures and sea-ice extent in the Southern Ocean (AtlanticeIndian sector): a multiproxy approach. Paleoceanography 18, PA1061. http://dx.doi.org/10.1029/2002PA000809. Gladyshev, S., 1998. Estimation of dense cold water formation on the northern shelf of the Sea of Okhotsk. Russ. Meteorol. Hydrol. 4, 53e59. Gorbarenko, S.A., Southon, J., Keigwin, L., Cherepanova, M.V., Gvozdeva, I.G., 2004. Late PleistoceneeHolocene oceanographic variability in the Okhotsk Sea: geochemical, lithological and paleontological evidence. Palaeogeogr. Palaeoclimatol. Palaeoecol. 209, 281e301. Gorbarenko, S.A., Psheneva, O. Yu, Artemova, A.V., Matul, A.G., Tiedemann, R., Nürnberg, D., 2010. Paleoenvironment changes in the NW Okhotsk Sea for the last 18 ka determined with micropaleontological, geochemical, and lithological data. Deep-Sea Res. Part I 57, 797e811. Gorbatenko, K.M., 1996. Seasonal aspects of the vertical distribution of zooplankton in the Sea of Okhotsk. Izvestya Ticookeanskovo Nauchno-Isledovatelskvo Ribocoziastvenovo Centra 119, 88e119 (in Russian). Gowing, M.M., 1986. Trophic biology of phaeodarian radiolarians and flux of living radiolarians in the upper 2000 m of the North Pacific central gyre. Deep-Sea Res. Part A 33, 655e674. Hargrave, B.T., 1975. The importance of total and mixed layer depth in the supply of organic material to bottom communities. Symp. Biol. Hung. 15, 157e165. Harrison, S.P., Kutzbach, J.E., Prentice, I.C., Behling, P.J., Sykes, M.T., 1995. The response of the Northern Hemisphere extratropical climate and vegetation to orbitally induced changes in insolation during the last interglaciation. Quat. Res. 43, 174e184. J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 Haug, G.H., Ganopolski, A., Sigman, D.M., Rosell-Mele, A., Swann, G.E.A., Tiedemann, R., Jaccard, S.L., Bollmann, J., Maslin, M.A., Leng, M.J., Eglinton, G., 2005. North Pacific seasonality and the glaciation of North America 2.7 million years ago. Nature 433, 821e825. Hays, J.D., Lozano, J.A., Shackleton, N.J., Irving, G., 1976. Reconstructions of the Atlantic and western Indian Ocean sectors of the 18,000 B.P. Antarctic Ocean. In: Cline, R.M., Hays, J.D. (Eds.), Investigations of Late Quaternary Paleoceanography and Paleoclimatology, Geol. Soc. Am. Mem. 145, Boulder, Colorado, pp. 337e372. Hays, J.D., Morley, J.J., 2004. The Sea of Okhotsk: a window on the ice age ocean. Deep-Sea Res. Part I 51, 593e618. Herguera, J.C., 1992. Deep-sea benthic foraminifera and biogenic opal: glacial to postglacial productivity changes in the western equatorial Pacific. Mar. Micropaleontol. 19, 79e98. Herguera, J.C., Jansen, C., Berger, W.H., 1992. Evidence of a bathyl front at 2000 m depth in the glacial Pacific, based on a depth transect on Ontong Java Plateau. Paleoceanography 7, 273e288. Honda, M.C., Kusakabe, M., Nakabayashi, S., Manganini, S.J., Honjo, S., 1997. Change in pCO2 through biological activity in the marginal seas of the western North Pacific: the efficiency of the biological pump estimated by a sediment trap experiment. J. Oceanogr. 53, 645e662. Honjo, S., Manganini, S.J., 1996. Dichothermal Layer and Biological Export Production in the Sea of Okhotsk, International Workshop on the Okhotsk Sea and Arctic; the Physics and Biogeochemistry Implied to the Global Cycles (Influence of Sea Ice on Climate and Marine Ecosystems). Japan Marine Science and Technology Center, Yokosuka, Japan, pp. 103e110. Horikawa, K., Asahara, Y., Yamamoto, K., Okazaki, Y., 2010. Intermediate water formation in the Bering Sea during glacial periods: evidence from neodymium isotope ratios. Geology 38, 435e438. Hülsemann, K., 1963. Radiolaria in Plankton from the Arctic Drifting Station T-3, Including the Description of Three New Species. Arctic Institute of North America Technical Papers 13, pp. 1e52. Isarin, R.F.B., Bohncke, S.J.P., 1999. Mean july temperatures during the Younger Dryas in northwestern and central Europe as inferred from climate indicator plant species. Quat. Res. 51, 158e173. Isarin, R.F.B., Renssen, H., Vandenberghe, J., 1998. The impact of the North Atlantic Ocean on the Younger Dryas climate of northwestern and central Europe. J. Quat. Sci. 13, 447e453. Itaki, T., 2003. Depth-related radiolarian assemblage in the water-column and surface sediments of the Japan Sea. Mar. Micropaleontol. 47, 253e270. Itaki, T., Ito, M., Narita, H., Ahagon, N., Sakai, H., 2003. Depth distribution of radiolarians from the Chukchi and Beaufort Seas, western Arctic. Deep-Sea Res. Part I 50, 1507e1522. Jaccard, S.L., Haug, G.H., Sigman, D.M., Pedersen, T.F., Thierstein, H.R., Röhl, U., 2005. Glacial/interglacial changes in subarctic North Pacific stratification. Science 308, 1003e1006. Jaccard, S.L., Galbraith, E.D., Sigman, D.M., Haug, G.H., Francois, R., Pedersen, T.F., Dulski, P., Thierstein, H.R., 2009. Subarctic Pacific evidence for a glacial deepening of the oceanic respired carbon pool. Earth Planet. Sci. Lett. 277, 156e165. Jacot Des Combes, H., Abelmann, A., 2007. A 350-ky radiolarian record off Lüderitz, Namibiaeevidence for changes in the upwelling regime. Mar. Micropaleontol. 62, 194e210. Jones, P.D., Raper, S.C.B., Bradley, R.S., Diaz, H.F., Kelly, P.M., Wigley, T.M.L., 1986. Northern hemisphere surface air temperature variations: 1851e1984. J. Clim. 25, 161e179. Keigwin, L., Jones, G.A., Froelich, P.N., 1992. A 15,000 year paleoenvironmental record from Meiji Seamount, far northwestern Pacific. Earth Planet. Sci. Lett. 111, 425e440. Keigwin, L.D., 1995. Stable isotope stratigraphy and chronology of he upper Quaternary section of site 883 Detroit Seamount. In: Rea, D.K., Basov, I.A., Scholl, D.W., Allan, J.F. (Eds.), Proc. Ocean Drilling Prog. Sci. Res. 145, pp. 257e264. Keigwin, L.D., 1998. Glacial-age hydrography of the far northwest Pacific Ocean. Paleoceanography 13, 323e339. Kent, D.V., Opdyke, N.D., Ewing, M., 1971. Climate change in the North Pacific using ice-rafted detritus as a climatic indicator. Geol. Soc. Am. Bull. 82, 2741e2754. Kerr, R.C., Kuiper, G.S., 1997. Particle settling through a diffusive-type thermohaline staircase in the ocean. Deep-Sea Res. Part I 44, 399e412. Kiefer, T., Sarnthein, M., Erlenkeuser, H., Grootes, P.M., Roberts, A.P., 2001. North Pacific response to millennial-scale changes in ocean circulation over the last 60 ka. Paleoceanography 16, 179e189. Kienast, S.S., Hendy, I.L., Crusius, J., Pedersen, T.F., Calvert, S.E., 2004. Export production in the subarctic North Pacific over the last 800 kyrs: no evidence for iron fertilization? J. Oceanogr. 60, 189e203. Kitani, K., 1973. An oceanographic study of the Okhotsk Sea e particularly in regard to cold waters. Bull. Far Seas Fish. Res. Lab. 9, 45e76. Kling, S.A., 1979. Vertical distribution of polycystine radiolarians in the central North Pacific. Mar. Micropaleontol. 4, 295e318. Kling, S.A., Boltovskoy, D., 1995. Radiolarian vertical distribution patterns across the Southern California current. Deep-Sea Res. Part I 42, 191e231. Kohfeld, K., Chase, Z., 2011. Controls on deglacial changes in biogenic fluxes in the North Pacific Ocean. Quat. Sci. Rev. 30, 3350e3363. Koizumi, I., Yamamoto, H., 2007. Paleohydrography of the Kuroshio-Kuroshio extension off Sanriku coast based on fossil diatoms. JAMSTEC Rep. Res. Develop. 5, 1e8. 91 Kruglikova, S.B., 1975. Radiolarians in the surface layer of sediments in the Sea of Okhotsk. Oceanology 15 (1), 82e86. Ling, H.Y., Stadum, C.J., Welch, M.L., 1971. Polycystine radiolaria from Bering Sea surface sediments. In: Farinacci, A. (Ed.), Proceedings of the 2nd Planktonic Conference, Rome, 1970, pp. 705e729. Tecnoscienza, Rome. Martin, J.H., 1990. Glacialeinterglacial CO2 change: the iron hypothesis. Paleoceanography 5, 1e13. Martin, S., Drucker, R., Yamashita, K., 1998. The production of ice and dense shelf water in the Okhotsk Sea polynyas. J. Geophys. Res. 103, 27771e27782. Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore Jr., T.C., Shackleton, N.J., 1987. Age dating and the orbital theory of the Ice Ages: development of a high resolution 0e300,000-year chronostratigraphy. Quat. Res. 1, 1e29. Matsumoto, K., Oba, T., Lynch-Stieglitz, J., Yamamoto, H., 2002. Interior hydrography and circulation of the glacial deep Pacific. Quat. Sci. Rev. 21, 1693e1704. Matsumoto, K., 2007. Biology-mediated temperature control on atmospheric pCO2 and ocean biogeochemistry. Geophys. Res. Lett. 34. http://dx.doi.org/10.1029/ 2007GL031301. Matul, A.G., 2011. The recent and quaternary distribution of the radiolarian species Cycladophora davisiana: a biostratigraphic and paleoceanographic tool. Oceanology 51, 335e346. McMillen, K.J., Casey, R.E., 1978. Distribution of living polycystine radiolarians in the Gulf of Mexico and Caribbean Sea, and comparison with the sedimentary record. Mar. Micropaleontol. 3, 121e145. Michaels, A.F., Silver, M.W., 1988. Primary production, sinking fluxes and the microbial food web. Deep-Sea Res. Part A 35, 473e490. Moore, T.C.J., 1973. Method of randomly distributing grains for microscopic examination. J. Sediment. Petrol 43, 904e906. Morley, J.J., 1983. Identification of density-stratified waters in the late Pleistocene North Atlantic: a faunal derivation. Quat. Res. 20, 374e386. Morley, J.J., Hays, J.D., 1983. Oceanographic conditions associated with the high abundances of the radiolarian Cycladophora davisiana. Earth Planet. Sci. Lett. 66, 63e72. Morley, J.J., Robinson, S.W., 1986. Improved method for correlating late Pleistocene/ Holocene records from the Bering Sea: application of a biosiliceous/geochemical stratigraphy. Deep-Sea Res. 33, 1203e1211. Morley, J.J., Stepien, J.C., 1985. Antarctic Radiolaria in late winter/early spring Weddell Sea waters. Micropaleontology 31, 365e371. Morley, J.J., Hays, J.D., Robertson, J.H., 1982. Stratigraphic framework for the late Pleistocene in the northwest Pacific Ocean. Deep-Sea Res. Part A 29, 1485e1499. Morley, J.J., Heusser, L.E., Shackleton, N.J., 1991. Late Pleistocene/Holocene radiolarian and pollen records from sediments in the Sea of Okhotsk. Paleoceanography 6, 121e131. Morley, J.J., Shemesh, A., Abelmann, A., 2013. Laboratory analysis of dissolution effects on Southern Ocean polycystine Radiolaria. Mar. Micropaleontol. (in press). Moroshkin, K.V., 1968. Water Masses of the Sea of Okhotsk. Joint Publications Research Service, Washington, D.C, p. 98. Mortlock, R.A., Froelich, P.N., 1989. A simple method for the rapid determination of biogenic opal in pelagic marine sediments. Deep-Sea Res. Part A 36, 1415e1426. Narita, H., Sato, M., Tsunogai, S., Murayama, M., Ikehara, M., Nakatsuka, T., Wakatsuchi, M., Harada, N., Ujiié, Y., 2002. Biogenic opal indicating less productive northwestern North Pacific during the glacial ages. Geophys. Res. Lett. 29, 22-1e22-4. http://dx.doi.org/10.1029/2001GL014320. Nelson, F.E., Outcalt, S.I., 1987. A computational method for prediction and regionalization of permafrost. Arct. Alp. Res. 19, 279e288. Nelson, P., 2004. Biological Physics: Energy, Information, Life. W. H. Freeman, New York, p. 598. Nigrini, C.A., Moore, T.C., 1979. A Guide to Modern Radiolaria. In: Cushman Foundation for Foraminiferal Res. Spec. Publ. 16, 342 pp. Nimmergut, A., Abelmann, A., 2002. Spatial and seasonal changes of radiolarian standing stocks in the Sea of Okhotsk. Deep-Sea Res. Part I 49, 463e493. Ninkovich, D., Robertson, J.H., 1975. Volcanogenic effects on rates of deposition of sediments in northwest Pacific Ocean. Earth Planet. Sci. Lett. 27, 127e136. Notoro, M., Liu, Z., 2007. Potential impact of the Eurasian boreal forest on North Pacific climate variability. J. Clim. 20, 981e992. Notoro, M., Liu, Z., 2008. Statistical and dynamical assessment of vegetation feedbacks on climate over the boreal forest. Clim. Dyn. 31, 691e712. Ogi, M., Tachibana, Y., Nishio, F., Danchenkov, M.A., 2001. Does the fresh water supply from the Amur River flowing into the Sea of Okhotsk affect sea ice formation? J. Meteorol. Soc. Jpn. 79, 123e129. Ohkushi, K., Itaki, T., Nemoto, N., 2003. Last GlacialeHolocene change in intermediate-water ventilation in the Northwestern Pacific. Quat. Sci. Rev. 22, 1477e1484. Okazaki, Y., Takahashi, K., Nakatsuka, T., Honda, M.C., 2003a. The production scheme of Cycladophora davisiana (Radiolaria) in the Okhotsk Sea and the northwestern North Pacific: implications for the palaeoceanographic conditions during the glacials in the high latitude oceans. Geophys. Res. Lett. 30. http:// dx.doi.org/10.1029/2003GL018070. Okazaki, Y., Takahashi, K., Yoshitani, H., Nakatsuka, T., Ikehara, M., Wakatsuchi, M., 2003b. Radiolarians under the seasonally sea-ice covered conditions in the Okhotsk Sea: flux and their implications for palaeoceanography. Mar. Micropaleontol. 49, 195e230. Okazaki, Y., Takahashi, K., Itaki, T., Kawasaki, Y., 2004. Comparison of radiolarian vertical distributions in the Okhotsk Sea near Kuril islands and the northwestern North Pacific off Hokkaido Island. Mar. Micropaleontol. 51, 257e284. 92 J.D. Hays et al. / Quaternary Science Reviews 82 (2013) 78e92 Okazaki, Y., Takahashi, K., Katsuki, K., Ono, A., Hori, J., Sakamoto, T., Uchida, M., Shibata, Y., Ikehara, M., Aoki, K., 2005. Late Quaternary paleoceanographic changes in the southwestern Okhotsk Sea: evidence from geochemical, radiolarian, and diatom records. Deep-Sea Res. Part II 52, 2332e2350. Petit, J.R., Jouzel, J., Raynaud, D., Barkov, N.I., Barnola, J.-M., Basile, I., Bender, M.L., Chappellaz, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyokov, V.M., Legrand, M., Lipenkov, V.Y., Lorius, C., Pepin, L., Ritz, C., Saltzman, E., Stievenard, M., 1999. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature 399, 429e436. Petrushevskaya, M., 1967. Radiolarian of orders Spumellaria and Nassellaria of the Antarctic region. In: Andriyashev, A.P., Ushakov, P.V. (Eds.), Biological Report of the Soviet Antarctic Expedition (1955e1958), vol. 3. Academy of Sciences of the USSR, Israel Program for Scientific Translations, Jerusalem, pp. 2e186. Rea, D.K., Basov, I.A., Janecek, T.R., Palmer-Julson, A., et al., 1993. Proc. Ocean Drilling Prog., Init. Repts. Leg 145. College Station, TX. http://dx.doi.org/10.2973/ odp.proc.ir.145.1993. Renz, G.W., 1976. The distribution and ecology of Radiolaria in the central Pacific: plankton and surface sediments. Bull. Scripps Inst. Oceanogr. 22, 1e267. Rickert, D., Schluter, M., Wallmann, K., 2002. Dissolution kinetics of biogenic silica from the water column to the sediments. Geochim. Cosomochim. Acta 66, 439e455. Robertson, J.H., 1975. Glacial to Interglacial Oceanographic Changes in the Northwest Pacific, Including a Continuous Record of the Last 400,000 Years. Ph.D. thesis. Columbia University, New York, p. 355. Robson, S., 1983. The distribution of recent Radiolaria in the surficial sediments of the continental margin off northern Namibia. J. Micropalaeontol. 2, 31e38. Rogachev, K.A., 2000. Recent variability in the Pacific western subarctic boundary currents and Sea of Okhotsk. Prog. Oceanogr. 47, 299e336. Rooth, C., 1982. Hydrology and ocean circulation. Prog. Oceanogr. 11, 131e149. Sachs, H.M., 1973. Quantitative Radiolarian-based Paleo-oceanography in Late Pleistocene Subarcitc Pacific Sediments. Ph.D. thesis. Brown University, Providence, Rhode Island, p. 208. Sapozhnikov, V.V., Gruzevich, A.K., Arzhanova, N.V., Naletova, I.A., Zubarevich, V.L., Sapozhnikov, M.V., 1999. Principal features of spatial distribution of organic and inorganic nutrient compounds in the Sea of Okhotsk. Okeanology 39, 198e204. Sarmiento, J.L., Gruber, N., Brzezinski, M.A., Dunne, J.P., 2004. High-Latitude controls of thermocline nutrients and low latitude biological productivity. Nature 427, 56e60. Sarnthein, M., Gebhardt, H., Kiefer, T., Kucera, M., Cook, M., Erlenkeuser, H., 2004. Mid Holocene origin of the sea surface salinity low in the Subarctic North Pacific. Quat. Sci. Rev. 23, 2089e2099. Sato, M.M., Narita, H., Tsugonai, S., 2002. Barium increasing prior to opal during the last termination of glacial ages in the Okhotsk Sea sediments. J. Oceanogr. 58, 461e467. Seki, O., Ikehara, M., Kawamura, K., Nakatsuka, T., Ohnishi, K., Wakatsuchi, M., Narita, H., Sakamoto, T., 2004. Reconstruction of paleoproductivity in the Sea of Okhotsk over the last 30 ka. Paleoceanography 19, PA1016. http://dx.doi.org/ 10.1029/2002PA000808. Severinghaus, J., Sowers, T., Brook, E.J., Alley, R.B., Bender, M.L., 1998. Timing of abrupt climate change at the end of the Younger Dryas interval from thermally fractionated gases in polar ice. Nature 391, 141e146. Shackleton, N.J., Berger, A., Peltier, W.R., 1990. An alternative astronomical calibration of the lower Pleistocene timescale based on ODP site 677. Trans. Royal Soc. Edinb.: Earth Sci. 81, 251e261. Shemesh, A., Burckle, L.H., Froelich, P.N., 1989. Dissolution and preservation of Antarctic diatoms and the effect on sediment thanatocoenoses. Quat. Res. 31, 288e308. Shigemitsu, M., Narita, H., Watanabe, Y.W., Harada, N., Tsunogai, S., 2007. Ba, Si, U, Al, Sc, La, Th, C, and 13C/12C in a sediment core in the western subarctic Pacific as proxies of past biological production. Mar. Chem. 106, 442e455. Sigman, D.M., Altabet, M.A., Francois, R., McCorkle, D.C., Gaillard, J.-F., 1999. The isotopic composition of diatom-bound nitrogen in Southern Ocean sediments. Paleoceanography 14, 118e134. Sigman, D.M., Jaccard, S.L., Haug, G.H., 2004. Polar ocean stratification in a cold climate. Nature 428, 59e63. Sigman, D.M., Hain, M.P., Haug, G.H., 2010. The polar ocean and glacial cycles in atmospheric CO2 concentration. Nature 466, 47e55. Sorokin, Y.I., 2002. Dynamics of inorganic phosphorus in pelagic communities of the Sea of Okhotsk. J. Plankton Res. 24, 1253e1263. Sorokin, Y.I., Sorokin, P.Y., 1999. Production in the Sea of Okhotsk. J. Plankton Res. 21, 201e230. Spencer-Cervato, C., Lazarus, D.B., Beckmann, J.-P., von Salis Perch-Nielsen, K., Biolzi, M., 1993. New calibration of Neogene radiolarian events in the North Pacific. Mar. Micropaleontol. 21, 261e293. Stephens, B.B., Keeling, R.F., 2000. The influence of Antarctic sea ice on glaciale interglacial CO2 variations. Nature 404, 171e174. Svendsen, J.I., Alexanderson, H., Astakhov, V.I., Demidov, I., Dowdeswell, J.A., Funder, S., Gataullin, V., Henriksen, M., Hjort, C., Houmark-Nielsen, M., Hubberten, H.W., Ingolfsson, O., Jacobsson, M., Kjaer, K.H., Larsen, E., Lokrantz, H., Lunkka, J.P., Lysa, A., Mangerud, J., Matiouchkov, A., Murray, A., Moller, P., Niessen, F., Nikolskaya, O., Polhyak, L., Saarnisto, M., Siegert, C., Siegert, M., Spielhagen, R.F., Stein, R., 2004. Late Quaternary ice sheet history of northern Europe. Quat. Sci. Rev. 23, 1229e1271. Sverdrup, H.U., 1953. On conditions for the vernal blooming of phytoplankton. J. Cons. Int. Explor. Mer. 18, 287e295. Takahashi, K., 1991. Radiolaria: flux, ecology, and taxonomy in the Pacific and Atlantic. In: Honjo, S. (Ed.), Ocean Biocoenosis Series No. 3. Woods Hole Oceanographic Institution, Woods Hole, Massachusetts, p. 303. Takahashi, K., 1995. Opal particle flux in the subarctic Pacific and Bering Sea and sidocoenosis preservation hypothesis. In: Tsunogai, S., Iseki, K., Koike, I., Oba, T. (Eds.), Global Fluxes of Carbon and its Related Substances in the Coastal Seae OceaneAtmosphere System, Proceedings of the 1994 Sapporo IGBP Symposium. M & J International, Yokohama, Japan, pp. 458e466. Tanaka, S., Takahashi, K., 2005. Late Quaternary paleoceanographic changes in the Bering Sea and the western subarctic Pacific based on radiolarian assemblages. Deep-Sea Res. Part II 52, 2131e2149. Tanaka, S., Takahashi, K., 2008. Detailed vertical distribution of radiolarian assemblage (0e3000 m, fifteen layers) in the central subarctic Pacific, June 2006. Mem. Fac. Sci. Kyushu Univ. Ser. D, Earth Planet. Sci. Lett. 32, 49e72. Tauxe, L., Herbert, T., Shackleton, N.J., Kok, Y.S., 1996. Astronomical calibration of the Matuyama-Brunhes boundary: consequences for magnetic remanence acquisition in marine carbonates and the Asian loess sequences. Earth Planet. Sci. Lett. 140, 133e146. Ternois, Y., Kawamura, K., Keigwin, L.D., Ohkouchi, N., Nakatsuka, T., 2001. A biomarker approach for assessing marine and terrigenous inputs to the sediments of Sea of Okhotsk for the last 27,000 years. Geochim. Cosmo-chem. Acta 65, 791e802. Thompson, J., Wallace, H.E., Colley, S., Toole, J., 1990. Authigenic uranium in Atlantic sediments of the last glacial stageea diagenetic phenomenon. Earth Planet. Sci. Lett. 98, 222e232. Tibbs, J.F., 1967. On some planktonic protozoa taken from the track of drift station ARLIS I, 1960e1961. Arctic Inst. N. Am. 20, 247e254. Tsunogai, S., 2002. The western North Pacific playing a key role in global biogeochemical cycles. J. Oceanogr. 58, 245e257. Van Cappellen, P., Qui, L., 1997. Biogenic silica dissolution in sediments of the Southern Ocean. Deep-Sea Res. II 44, 1129e1149. Vinogradov, M.E., 1970. Vertical Distribution of the Oceanic Zooplankton. Translated from Russian Izdatel’stvo “Nauka” Moskva 1968 by Israel Program for Scientific Translations, p. 339. Wang, R., Xiao, W., Li, Q., Chen, R., 2006. Polycystine radiolarians in surface sediments from the Bering Sea Green Belt area and their ecological implication for paleoenvironmental reconstructions. Mar. Micropaleontol. 59, 135e152. Washburn, A.L., 1980. Geocryology. John Wiley and Sons, New York, p. 406. Warren, B.A., 1983. Why is no deep water formed in the North Pacific? J. Marine Res. 41, 327e347. Welling, L.A., Pisias, N.G., Roelofs, A.K., 1992. Radiolarian Microfauna in the Northern California Current System: Indicators of Multiple Processes Controlling Productivity. In: Geol. Soc. London Spec. Pub. 64, pp. 177e195. Winton, M., 1997. The effect of cold climate upon North Atlantic deep water formation in a simple oceaneatmosphere model. J. Clim. 10, 37e51.
© Copyright 2025 Paperzz