Oceanography The Official Magazine of
The Oceanography Society
Volume 37 Issue 2

View Issue TOC
Volume 37, No. 2
Pages 46 - 59

OpenAccess

The Southern Ocean Hub for Nutrients, Micronutrients, and Their Isotopes in the Global Ocean

By Gregory F. de Souza  and Adele K. Morrison 
Jump to
Article Abstract Full text Citation References Copyright & Usage
Article Abstract

The sustenance of marine primary productivity depends on the supply of macro- and micronutrients to photosynthesizers in the ocean’s sunlit surface. Without supply from the deep, sinking particles would deplete the upper ocean of these vital elements within decades. Over the last 20 years, it has been recognized that the Southern Ocean, where nutrient-rich deep waters are brought to the surface and the water masses that fill much of the upper ocean are formed, plays a pivotal role in replenishing upper-ocean nutrients. Photosynthesizers that grow and take up nutrients within the Southern Ocean circulation “hub” thus have an outsize influence on global-​scale distributions of macronutrients and many micronutrients. The GEOTRACES program has contributed observations of the concentration and stable isotope composition of “nutrient-​type” metals like zinc, cadmium, and nickel, within the Southern Ocean and beyond it, that are driving a sea change in our understanding of their marine cycles. Simultaneously, our understanding of Southern Ocean circulation has been refined, with recognition of the importance of longitudinal variability and subtropical overturning. Here, we aim to bring together these two strands of progress, review insights gained into marine micronutrient cycling, and consider the questions that remain to be resolved.

FACING PAGE. Photo of the Mertz Polynya (~67°S, 145°E) taken on the Swiss Polar Institute’s Antarctic Circumnavigation Expedition (2016–2017). Researchers were exploring physical, chemical, and biological aspects of the ocean, atmosphere, and terrestrial sites on islands around Antarctica as well as on the continent. Photo credit: G.F. de Souza. > High res figure
Full Text

INTRODUCTION

The physiological requirements of marine photosynthesizers dictate their reliance on a slew of nutrients that they must draw from seawater. In addition to the macronutrients nitrogen, phosphorus, and (in some cases) silicon, marine phytoplankton require metal micronutrients in order to photosynthesize, efficiently fix carbon, protect themselves from oxidative stress, or access certain macronutrient pools (Fraústo da Silva and Williams, 2001; Morel et al., 2014). The nutritional status and ecological composition of marine phytoplankton communities thus depend on what might be termed the “nutrientscape” of the sunlit surface ocean in which they live—i.e., the distributions and relative abundances of vital dissolved macro- and micronutrients (e.g., Dutkiewicz et al., 2009). Simultaneously, these distributions are shaped by how surface biological uptake and downward particulate export of nutrients interact with ocean circulation to cycle and transport them through the global ocean (e.g., Sarmiento et al., 2007; Sunda, 2012).

A process key to continued surface-ocean primary productivity is the replenishment of nutrients, lost from the upper ocean through the sinking of biogenic particles, by the upwelling of deep, nutrient-rich waters, which happens primarily in the Southern Ocean surrounding Antarctica (Toggweiler, 1994; Marshall and Speer, 2012; Talley, 2013). In addition, mode waters formed at the northern edge of the Southern Ocean ventilate the thermocline of the low-​latitude ocean (Morrison et al., 2022), and the abundance and stoichiometry of nutrients in these waters provide a boundary condition for the supply of nutrients to low-​latitude ecosystems by coastal or equatorial upwelling (Sarmiento et al., 2004).

As a body of work beginning two decades ago has shown, the large-scale distributions of dissolved macronutrients—and the systematic differences between these distributions—are largely determined by how biogeochemical cycling of nutrients in the surface Southern Ocean, and its interaction with physical processes, modulates the nutrient content of waters between the zones of deep-water upwelling and thermocline ventilation (Sarmiento et al., 2004, 2007; Weber and Deutsch, 2010; Holzer and Primeau, 2013; Holzer et al., 2014). Since ground-breaking studies in the 1970s, it has also been known that vertical profiles of the concentrations of the metals zinc (Zn), cadmium (Cd), and nickel (Ni) in seawater mimic those of the major nutrients (Figure 1; Boyle et al., 1976; Sclater et al., 1976; Bruland et al., 1978). By vastly expanding data coverage and providing basin-scale sections of the abundance and stable isotope composition of these micronutrient metals, the GEOTRACES program has allowed a reevaluation of the mechanisms responsible for this striking similarity.

 

FIGURE 1. Profiles of the macronutrients (upper row) and metal micronutrients (lower row) from the subtropical North Pacific (32.7°N, 145°W) are plotted as reported by Bruland (1980). > High res figure

 

Micronutrient Mimics

Zinc, Cd, and Ni are only a few of the micronutrients that phytoplankton need; we focus on them here for reasons both conceptual and practical. Conceptually, the similarities and differences between their elemental and isotopic distributions nicely illustrate how Southern Ocean processes can shape global nutrient distributions. Practically, they each have a stable isotope system that provides an additional constraint on their cycling, and there are sufficient data to characterize their behavior in the Southern Ocean. We thus only very briefly consider the vital micronutrients iron and manganese, whose global distributions are not affected by Southern Ocean processes (see Box 1), and do not discuss the micronutrients cobalt, copper, and selenium, which are not controlled by the Southern Ocean or for which Southern Ocean data coverage remains too sparse.

 

> High res box

 

Zinc, Cd, and Ni all display what have classically been called “nutrient-type” distributions (Bruland, 1983): that is, their dissolved concentrations are at a minimum in the surface mixed layer and increase both downward in the water column and in deep waters from the North Atlantic to the North Pacific—thus mimicking the distributions of the macronutrients. This similarity can be seen in the vertical profiles of Figure 1, which also reveal differences between the three metals: the distribution of dissolved Cd most closely resembles those of nitrate (NO3) and phosphate (PO4), exhibiting the same increase through the thermocline and mid-depth maximum as these major nutrients (Boyle et al., 1976). Zinc, on the other hand, bears more similarity to silicic acid (Si(OH)4), with low concentrations extending from the surface ocean deeper into the thermocline, and reaching a maximum closer to the base of the water column (Bruland et al., 1978). A particularity of the Ni distribution (Sclater et al., 1976) is that, unlike the macronutrients or Zn and Cd, its concentration in the surface ocean never decreases below about ~20% of deep-water values, even in the nutrient-poor subtropical gyres (in contrast to <1% for Zn and Cd).

These three elements also have differing roles in phytoplankton biochemistry. Zinc has the most diverse set of biochemical roles in photosynthesizers, including as the metal center in the enzyme carbonic anhydrase, important for the efficient fixation of carbon (Morel et al., 1994, 2014). When ambient Zn concentrations are low, many phytoplankton can substitute cobalt or Cd for it (Lee and Morel, 1995; Sunda and Huntsman, 1995). Such “cambialistic” substitution for Zn may be the primary reason that Cd behaves as a micronutrient in the sea, because its only known biochemical role is in a Cd-bearing carbonic anhydrase in some marine diatoms (Lane et al., 2005). In phytoplankton, Ni is mainly used in nitrogen metabolism or to protect from oxidative stress (Ragsdale, 2009): it is associated with the metallo­enzyme urease, which catalyzes the metabolism of urea to ammonia; NiFe hydrogenase, which prevents the inhibition of nitrogen fixation by O2 and H2 in diazotrophs; and a Ni-bearing superoxide dismutase primarily found in cyanobacteria, which gets rid of harmful reactive oxygen species.

At the time when the first analyses of these micronutrients in sea­water were made, the marine biogeochemical paradigm interpreted major nutrient distributions in an essentially one-dimensional fashion (e.g., Broecker and Peng, 1982). In that view, concentrations of silicic acid increase deeper in the water column than the other two macronutrients (Figure 1) because siliceous hard parts dissolve more slowly than bacteria decompose organic matter—so that Si is released from sinking particles at greater depths than NO3 and PO4. The similarities between the distributions of Zn, Cd, and Ni and the macronutrients were analogously interpreted as resulting from directly coupled cycling: for instance, that profiles of Zn and Si look similar because Zn is incorporated into the siliceous frustules of diatoms (Bruland et al., 1978). But we now know that the marine distributions of dissolved nutrients are produced by a complex set of three-dimensional interactions between biologically driven fluxes and physical transport and mixing processes. On both these fronts, Southern Ocean processes play a key role at the scale of the global ocean.

 

THE SOUTHERN OCEAN CIRCULATION HUB

The current paradigm for explaining Southern Ocean nutrient resupply emphasizes a longitudinally averaged, two-dimensional framework with the meridional overturning circulation as the primary driver of upward nutrient transport (e.g., Sarmiento et al., 2004; Morrison et al., 2015). The upwelling limb of the overturning circulation is largely controlled by the westerly winds over the Southern Ocean, which drive northward Ekman transport in the surface layer. South of the latitude where the westerly winds are strongest (~50°S), the Ekman transport is divergent and thus draws up nutrient-rich deep waters from below, along the sloping density layers of the Southern Ocean. Upon reaching the surface mixed layer, one portion of these upwelled waters flows northward in the “upper” overturning cell, while the other portion flows south toward Antarctica in the “lower” overturning cell. It is the upper overturning cell that dominates the supply of nutrients for global primary production, due to its northward surface flow (Marinov et al., 2006; Primeau et al., 2013). The upwelling waters of the lower overturning cell have limited residence time when they outcrop at the surface around Antarctica, as they are rapidly returned to the abyss, influencing deep water nutrient distributions (Sarmiento et al., 2007).

Recent work has highlighted the unique three-dimensional structure of Southern Ocean upwelling. Rather than being longitudinally homogeneous, the upwelling is intensified in a handful of eddy hotspots on the eastern side of large bathymetric features (Figure 2a; e.g., Foppert et al., 2017; Tamsitt et al., 2017). The localization of the upwelling is particularly prominent at ~1,000 m depth (Tamsitt et al., 2017; Yung et al., 2022), but localization is also evident at the base of the mixed layer (Viglione and Thompson, 2016). Modeling studies suggest that the net circumpolar upwelling transport of nutrients along isopycnals is similarly dominated by eddies at a small number of localized hotspots (Dufour et al., 2015). In addition to this intensified upwelling transport at eddy hotspots, there may also be significant larger-scale inter-basin differences in upwelling transport of deep waters into the mixed layer (Viglione and Thompson, 2016; Prend et al., 2022). However, an open question remains regarding what impact any longitudinal variations in nutrient delivery to the mixed layer have on the distribution of nutrients and phytoplankton ecology across the Southern Ocean. The distribution of biogeochemical properties in the surface Southern Ocean is strongly guided by the Antarctic Circumpolar Current (ACC), whose flow is organized into a series of jets aligned with strong density fronts (i.e., sharp changes in the physical structure of the water column on either side of the jet). Nutrient concentrations in the surface Southern Ocean are generally homogeneous between these fronts, with the strongest concentration gradients coinciding with the physical front (Pollard et al., 2002). It may be that, despite inhomogeneous upwelling, the strong flow of the ACC, as well as lateral mixing by eddies, smooths out longitudinal (along-front) gradients in nutrient concentrations (e.g., Morrison et al., 2022).

 

FIGURE 2. The three-dimensional structure of the physical processes impacting nutrients in the Southern Ocean. (a) The inhomogeneous spatial distribution of simulated upwelling of deep waters at 1,000 m depth. Reproduced from Tamsitt et al. (2017) (b) Processes (enumerated) contributing to the export of nutrients northwards into the global thermocline. Blue colors show potential density, and orange colors show potential vorticity on a density surface in the interior. Low potential vorticity is an indicator of recently ventilated mode water. Reproduced from Morrison et al. (2022). > High res figure

 

The mode waters of the global thermocline are replenished and enriched by subduction on the northern edge of the Antarctic Circumpolar Current (Sallée et al., 2010). The subducting waters are a mixture of both northward-flowing Southern Ocean surface waters and subtropical waters flowing southward in western boundary currents (Figure 2b). While many nutrients exhibit negligible concentrations in the subtropical source waters, it is still necessary to consider contributions from both northern and southern source waters to understand nutrient distributions in the mode waters (Fernández Castro et al., 2022). Many conceptual frameworks and schematics of mode water formation focus exclusively on the role of the two-​dimensional longitudinally averaged overturning circulation (e.g., Sarmiento et al., 2004; Marinov et al., 2006). However, like the upwelling, the transfer of nutrients from the surface of the Southern Ocean northward into the global thermocline is a complex, three-dimensional process (Morrison et al., 2022). Subduction of waters from the mixed layer into the interior occurs in localized hotspots, with substantial inter-basin differences (Sallée et al., 2010). After subduction, water mass properties are rapidly homogenized across the mode waters by eddy mixing and gyre circulation (e.g., Gupta et al., 2022). Eddy-driven mixing along isopycnals may also contribute substantially to the northward flux of nutrients into the mode waters, depending on the nutrient gradient along isopycnals (Fripiat et al., 2021).

 

THE SOUTHERN OCEAN NUTRIENT HUB

The role of the Southern Ocean in both the upwelling of deep waters and the ventilation of the low-latitude thermocline makes it an important “hub” of the global overturning circulation. As a result, the physical and biogeochemical properties of abyssal, intermediate, and upper-ocean water masses formed here influence tracer distributions at a near-global scale. When it comes to the distributions of macro- and micronutrients, what is particularly important is how the uptake of nutrients by Southern Ocean ecosystems modifies the properties of nutrient-​rich waters upwelled to the surface, before they are subducted into the ocean interior as mode and intermediate waters.

This is shown most clearly by the effect of Southern Ocean nutrient uptake on macronutrients (Sarmiento et al., 2004, 2007). Primary production in the high-latitude Southern Ocean is dominated by diatoms—phytoplankton with opaline cell walls that thrive in dynamic, highly seasonal and competitive environments (Margalef, 1978). Probably as the combined result of limitation by light and Fe (see Box 1) and adaptation to Si-rich waters fed by deep-water upwelling, Southern Ocean diatoms are heavily silicified (Baines et al., 2010). They thus draw down Si much more strongly than the other major nutrients NO3 and PO4, so that as surface waters spiral northward across the fronts of the ACC, Si is depleted more quickly (i.e., further south) than nitrate or phosphate (Figure 3). Surface ocean Si concentrations are generally depleted north of the Polar Front, the southernmost of the ACC fronts north of the upwelling zone (Pollard et al., 2002).

 

FIGURE 3. Latitudinal variation in macro- and micronutrient concentrations and isotopic compositions in the surface Southern Ocean. The upper four rows show data from the GEOTRACES GIPY04/GIPY05 transect mostly along the zero meridian (Abouchami et al., 2011; Fripiat et al., 2011; Zhao et al., 2014; Wyatt et al., 2014). No δ15N-NO3 data are available for this transect; see Sigman et al. (1999) for a Pacific-sector cross-frontal δ15N-NO3 section. The bottom row shows data from the surface ocean (<25 m) from GEOTRACES GIPY06 transect along ~140°W. These data are unpublished and reproduced with kind permission of Andrew Bowie. PF and SAF denote approximate positions of the Polar Front and Subantarctic Front. Error bars on isotopic data represent uncertainty as reported by the authors. > High res figure

 

The mode and intermediate waters formed at the northern edge of the ACC inherit this surface signal and are thus relatively poor in Si—a characteristic that they impart to the low-latitude thermocline. Thus, the fact that Si concentrations in the upper ocean are low relative to nitrate and phosphate (Figure 1), and increase deeper in the thermocline, has less to do with the fact that dissolution of (Si-bearing) opal takes place deeper than the remineralization of (N- and P-bearing) organic matter; rather, it stems from the fact that high-latitude diatoms have more efficiently stripped Si out of the source regions of the waters that ventilate the thermocline (Sarmiento et al., 2004, 2007; Holzer et al., 2014).

 

Macronutrient Isotopes and the Global Reach of Southern Ocean Uptake

This southerly view of global marine nutrient cycling emerged shortly before the first GEOTRACES expeditions during the International Polar Year 2007–2008, and has been substantiated by macronutrient stable isotope data that emerged partly from those early Southern Ocean transects. Phosphorus has only one stable isotope, but the stable isotope compositions of seawater nitrate and silicic acid are expressed using the δ notation, which represents the deviation (in parts per thousand) of the isotope ratio 15N/14N or 30Si/28Si from that of a standard, for example:


 

An increase in the δ15N of seawater nitrate (δ15N-NO3) or the δ30Si of silicic acid thus reflects an enrichment in the heavier isotopes of these elements. Biological uptake of nitrate and silicic acid is associated with isotope fractionation (Wada and Hattori, 1978; De La Rocha et al., 1997), with preferential uptake of the lighter isotopes leaving the residual dissolved pool isotopically heavy (i.e., with high δ15N-NO3 or δ30Si). Thus, in the Southern Ocean, as the concentrations of nitrate and silicic acid decrease across the fronts of the ACC, δ15N-NO3 and δ30Si increase (Figure 3; Sigman et al., 1999; Varela et al., 2004). These isotopic fingerprints of biological uptake are inherited by mode and intermediate waters when they form at the northern edge of the ACC—thus making nitrate and silicic acid in the thermocline isotopically heavy relative to the deep ocean (Sigman et al., 2000; Fripiat et al., 2011, 2023).

It is the presence of these water masses that causes δ15N-NO3 and δ30Si to increase at intermediate depths (Figure 4) in most of the mid- and low-latitude ocean; indeed, the thermocline distributions of δ15N-NO3 and δ30Si trace the equatorward spreading of mode and intermediate waters within the subtropical gyre circulation (de Souza et al., 2012a; Rafter et al., 2013; Grasse et al., 2020). In the Pacific, the biogeochemical impact of southern-sourced mode and intermediate waters is mostly restricted to the Southern Hemisphere (e.g., Sarmiento et al., 2004), but in the Atlantic Ocean, elevated δ15N-NO3 and δ30Si signals extend all the way to the North Atlantic and Arctic Oceans. This is because the upper limb of the overturning circulation transports southern-sourced waters into the Northern Hemisphere, eventually feeding the formation of North Atlantic Deep Water (Talley, 2013). Nitrate and silicic acid in North Atlantic Deep Water, whose constituent water masses form in the subpolar North Atlantic and Arctic, thus bear the telltale elevated δ15N-NO3 and δ30Si values that result from biological uptake at the other end of the globe, around Antarctica (de Souza et al., 2012b, 2015; Brzezinski and Jones, 2015; Holzer and Brzezinski, 2015; Marconi et al., 2015; Varela et al., 2016).

 

FIGURE 4. Profiles of the isotopic composition of macronutrients (upper row) and micronutrients (lower row) in the southern (sub)tropics. Micronutrient isotope data are all from the same station (GR15) occupied during GEOTRACES section GP19 in the southwest Pacific (Takano et al., 2017; Sieber et al., 2019; see also Figure 5). Macronutrient isotope data are from de Souza et al. (2012b) and Rafter et al. (2013). Error bars show uncertainty as reported by the authors. > High res figure

 

From Macronutrients to Micronutrients

The Southern Ocean control on large-scale nutrient cycling discussed above is generalizable to the micronutrients: if Southern Ocean ecosystems tend to take up a nutrient in excess of the major nutrients nitrate or phosphate, its large-scale distribution will be skewed toward the deep and abyssal ocean, because efficient export to depth in the Southern Ocean will reduce its concentrations in the thermocline and, ultimately, in deep waters formed at high northern latitudes. By providing datasets of consistently co-sampled micronutrient metals and macronutrients, the GEOTRACES program has enabled the systematic study of their relationships in the surface Southern Ocean, while basin-scale sections of metal concentrations and isotopes provide larger-scale constraints. Next, we review the progress made in understanding the role of the Southern Ocean in the marine Zn, Cd, and Ni cycles in the GEOTRACES era.

 

Zinc and Its Isotopes

Given what we know about how highly silicified Southern Ocean diatoms govern the marine Si distribution, it might be thought that the similarity between Zn and Si comes about because these diatoms also incorporate a lot of Zn into their siliceous frustules (Bruland et al., 1978). But because of its importance in the biochemistry of these eukaryotes, most of their cellular Zn is situated in their organic matter, with at most a few percent in the frustule (Ellwood and Hunter, 2000; Twining et al., 2014; although see the recent results of Grun et al., 2023). Why, then, does the marine Zn distribution not look more like that of NO3 and PO4, which are released back to seawater when organic matter is remineralized? The answer lies in the fact that phytoplankton trace metal quotas exhibit considerable variability at both the phenotypic and physiological levels (e.g., Ho et al., 2003; Twining and Baines, 2013). So, on the one hand, eukaryotic phytoplankton have generally higher cellular Zn quotas than prokaryotes (Saito et al., 2003), and diatoms have even higher Zn quotas than other co-existing eukaryotes. On the other hand, laboratory cultures have shown that the Zn quotas of diatoms and other eukaryotic phytoplankton increase when they are growth-​limited by Fe (Sunda and Huntsman, 2000) or when ambient seawater has higher concentrations of the free Zn ion (see Box 2; Sunda and Huntsman, 1992).

 

> High res box

 

So, when deep-water upwelling in the Southern Ocean supplies plenty of Zn to Fe-limited diatom communities, the ecosystem response is to take up large amounts of Zn: diatoms growing in Zn-rich waters south of the Antarctic Polar Front have Zn:P ratios that can be more than 10 times those of low-​latitude phytoplankton (Twining and Baines, 2013). As a result, Zn drawdown across the fronts of the ACC is as strong as that of Si, with Zn concentrations decreasing by a factor of 3 across the Antarctic Polar Front and becoming essentially depleted, together with Si, in the formation regions of mode and intermediate waters further north (Figure 3, Ellwood, 2004, 2008; Zhao et al., 2014; Janssen et al., 2020). Thus, although there is no biochemical coupling between their uptakes, the fact that Southern Ocean ecosystems strip both Si and Zn from surface waters means that Zn is “trapped” within the Southern Ocean like Si, rather than being exported into the low-latitude thermocline like NO3 and PO4 (Vance et al., 2017; de Souza et al., 2018). While this leading-order role of Southern Ocean uptake in governing the marine Zn distribution is clear from numerous studies (e.g., Roshan et al., 2018; Weber et al., 2018; Middag et al., 2019), it has been proposed that additional processes involving the reversible sorption of Zn to sinking particles must work north of the Southern Ocean, in order to maintain the similarity between the Zn and Si distributions even in regions not dominated by Southern Ocean water masses, such as the North Pacific (John and Conway, 2014; Weber et al., 2018; Zheng et al., 2021; Sieber et al., 2023a).

The most intriguing aspect of marine Zn is its stable isotope composition δ66Zn, which has a seawater distribution unique among the nutrient isotope systems. Even though most phytoplankton have been shown to preferentially take up the light isotopes of Zn in laboratory culture (e.g., John et al., 2007; Samanta et al., 2018; Köbberich and Vance, 2019), seawater δ66Zn is low in the thermocline and surface ocean, rather than high as for the macronutrients (Figure 4; Conway and John, 2014; Takano et al., 2017; John et al., 2018). This is certainly not the result of Southern Ocean isotope fractionation: here, as surface Zn concentrations drop precipitously toward the north, δ66Zn values show a minimal but just-resolvable increase (Zhao et al., 2014; Wang et al., 2019; Sieber et al., 2020). But this slight signal of biological fractionation appears to be lost north of the Subantarctic Front (Figure 3; Zhao et al., 2014; Ellwood et al., 2020; Sieber et al., 2020), perhaps because of the admixture of isotopically light zinc carried southward in subtropical thermocline waters (Figure 2b; Takano et al., 2017; Samanta et al., 2017). What exact processes are responsible for low subtropical δ66Zn is currently under debate, with two contrasting hypotheses suggested: (1) the removal of isotopically heavy Zn sorbed to sinking particles (John and Conway, 2014; Weber et al., 2018; Sieber et al., 2023a), which requires chelation by natural organic ligands (see Box 2) to prefer isotopically light Zn, at odds with laboratory studies of organic chelators (e.g., Ban et al., 2002; Marković et al., 2017); or (2) the addition of isotopically light Zn from external sources that may be anthropogenic (Lemaitre et al., 2020; Liao et al., 2020), although analyses of open-marine aerosols (Dong et al., 2013; Packman et al., 2022; Zhang et al., 2024) have thus far not revealed Zn isotope compositions that can explain the lowest δ66Zn values observed in thermocline waters. Whether either, both, or neither of these hypotheses explains the thermocline δ66Zn distribution remains the subject of active research.

 

Cadmium and Its Isotopes

Dissolved and particulate marine data consistently show that, of all the trace metals, Cd behaves most similarly to the macronutrients, being cycled tightly together with P (e.g., Abouchami et al., 2011; Twining et al., 2015; Yang et al., 2018; Middag et al., 2018; Ohnemus et al., 2019; Cloete et al., 2021). Its isotopic behavior, too, is most like the macronutrient systems δ15N-NO3 and δ30Si. Biological uptake fractionates the isotopes of Cd, and as the concentrations of Cd decrease across the fronts of the ACC, it becomes increasingly enriched in the heavy isotopes (i.e., δ114Cd increases; Figure 3; Abouchami et al., 2011; Xue et al., 2013). This similarity to the macronutrient stable isotope systems extends to much of the global ocean: as with δ15N-NO3 and δ30Si, Cd subducted in the mode and intermediate waters is isotopically heavy (Figure 4; Abouchami et al., 2014; Sieber et al., 2019a), and this elevated δ114Cd signal can be traced through the subtropical thermocline (Conway and John, 2015; Xie et al., 2017; George et al., 2019; Sieber et al., 2019b, 2023b) all the way into North Atlantic Deep Water (Abouchami et al., 2014; Conway and John, 2015).

But given its limited biochemical role, why should phytoplankton uptake of Cd so exactly mimic phosphate? In fact, in detail it does not: Southern Ocean surface transects show that Cd is taken up more strongly relative to PO4, with Cd reaching low levels north of the Subantarctic Front, while PO4 remains undepleted (Figure 3; Ellwood, 2008; Abouchami et al., 2011; Baars et al., 2014). This is consistent with two independent sets of observations. First, particulate data show that biogenic particles in the Southern Ocean are highly enriched in Cd relative to P (Bourne et al., 2018; Twining and Baines, 2013). Second, cultures of eukaryotic phytoplankton and incubations of natural assemblages have shown that they will take up more Cd under conditions found in the surface Southern Ocean: under growth limitation by Fe (Cullen et al., 2003), at high concentrations of the free Cd ion (see Box 2; Lee et al., 1995; Sunda and Huntsman, 1998), or at low concentrations of free Mn or Zn (see Box 1; Cullen et al., 1999; Sunda and Huntsman, 2000). So, Fe- and/or Mn-limited Southern Ocean diatom communities would be expected to take up more Cd even south of the Polar Front, and Cd uptake should increase even more strongly when Zn is reduced to low levels to its north, as is indeed observed (Figure 3; Sunda and Huntsman, 2000; Ellwood, 2008).

The stronger Southern Ocean uptake of Cd has also helped to clarify a feature of its distribution that long perplexed oceanographers: the low Cd:PO4 ratio of North Atlantic Deep Water that results in a slight nonlinearity (known affectionately as “the kink”) in the Cd–PO4 relationship in the Atlantic Ocean (Boyle, 1988; Frew and Hunter, 1992; de Baar et al., 1994). In the context of the Southern Ocean nutrient hub as it works for Zn and Si, and the influence of uptake on the Cd distribution in the Southern Ocean surface, it has become clear that the relative Cd-poverty of North Atlantic waters comes about because Southern Ocean phytoplankton reduce Cd concentrations in the upper limb of the overturning circulation slightly more than they do PO4 (Baars et al., 2014; Quay et al., 2015; Middag et al., 2018; Roshan and DeVries, 2021), such that Cd is depleted relative to PO4 in the thermocline (Figure 5).

 

FIGURE 5. Meridional sections of phosphate, silicic acid, and Cd in the southern Pacific from GEOTRACES section GP19. Macronutrient data are from the GEOTRACES IDP2021 (GEOTRACES Intermediate Data Product Group, 2021) and reproduced with kind permission of Toshitaka Gamo and Hajime Obata. Cadmium data are from Sieber et al. (2019). > High res figure

 

Nickel and Its Isotopes

Perhaps because it is never drawn down to very low concentrations in the surface ocean, nickel has until recently received less attention than Zn and Cd, and thus concentration and especially isotopic data are sparser for this metal. Nonetheless, it is clear that across the fronts of the ACC, Ni concentrations decrease much less strongly than those of Zn and Cd. Surface Ni concentrations decrease by just 20% across the Polar Front, and remain above 3 nmol kg–1—~50% of the highest surface Southern Ocean concentrations—even north of the Subantarctic Front, where both Zn and Cd are depleted (Figure 3; Ellwood, 2008; Janssen et al., 2020). In the oligotrophic subtropics, where all inorganic macro- and micronutrients are at very low concentrations, Ni is present at around 2 nmol kg–1, 20% of deep-ocean levels (Figure 1; Middag et al., 2020). It was proposed that this concentration might represent a pool that is not bioavailable (Price and Morel, 1991; Mackey et al., 2002), but recently, nutrient amendments have shown that both eukaryotic and prokaryotic phytoplankton can draw down this pool of sea­water Ni significantly when provided with sufficient other nutrients (John et al., 2022). Thus, based on biogeochemical modeling, it has been suggested that the non-zero subtropical Ni minimum instead represents a residual pool of Ni “left over” once other nutrients have been depleted, due to the low Ni requirements of Southern Ocean ecosystems (John et al., 2022).

Certainly, a smaller drawdown of Ni in the Southern Ocean is consistent with its limited role in eukaryote biochemistry. Its only known use in these organisms is in urease, which phytoplankton obligately need when their only nitrogen source is urea (Price and Morel, 1991; Egleston and Morel, 2008), but which may be less important in the nitrate-rich Southern Ocean, where urea is only seasonally an important nitrogen source, mostly in the Subantarctic (e.g., Joubert et al., 2011). Furthermore, sparse seasonally resolved observations also suggest only limited Ni uptake by Southern Ocean ecosystems, with summertime mixed-layer Ni concentrations only ~10% lower than in winter both south and north of the Subantarctic Front (Ellwood, 2008; Cloete et al., 2019). This makes it plausible that the marine Ni distribution, like that of Zn and Cd, is also driven by the stoichiometric requirements of diatom-dominated ecosystems in the Southern Ocean.

At the same time, the stable isotope composition of dissolved Ni (δ60Ni) shows that the subtropics cannot just be a passive receptacle for the Ni leftovers of the high latitudes. Values of δ60Ni show no resolvable variability in the Southern Ocean (Cameron and Vance, 2014; Wang et al., 2019; Archer et al., 2020) or in the subpolar North Atlantic (Lemaitre et al., 2022). In the subtropical ocean, however, δ60Ni increases through the upper thermocline to maxima in the surface (Figure 4; Takano et al., 2017; Archer et al., 2020; Yang et al., 2021; Lemaitre et al., 2022). Lemaitre et al. (2022) argue that this isotopic divide results from Ni cycling processes particular to the oligotrophic subtropical ocean, in which the dearth of inorganic nitrogen sources and the domination of the phytoplankton community by prokaryotes should both tend to increase relative Ni demand (see earlier section on Micronutrient Mimics).

This poses an interesting question that remains unresolved: is the marine Ni distribution driven by low-latitude Ni demand by prokaryotes, or instead by the limited Ni requirement of high-​latitude eukaryote-dominated ecosystems? In the latter case, the global Ni distribution is controlled from the south much like those of the macronutrients as well as Zn and Cd (John et al., 2022); in the former, the reduction of Ni concentrations in the Subantarctic, and with it the global Ni distribution, is at least partially driven by subtropical Ni drawdown, and the contribution of subtropical waters to Southern Ocean mode waters (Figure 2b; Fernández Castro et al., 2022). This would make the biogeochemical controls on the marine Ni cycle unique among the elements
considered here.

 

PERSPECTIVES

Data from the first two decades of GEOTRACES show that the Southern Ocean control on global macronutrient distributions is generalizable to numerous micronutrient metals. Simultaneously, physical oceanographic research over the last decade has increasingly revealed the zonal asymmetry of physical Southern Ocean processes key to biogeochemical tracer transport (Figure 2; Morrison et al., 2022; Gray, 2024)—for instance, that the pathways of upwelling and subduction in the Southern Ocean are concentrated in discrete hotspots in the longitudinal direction (e.g., Sallée et al., 2010; Tamsitt et al., 2017).

It remains unknown to what extent this heterogeneity in the physical system impacts the distribution of biogeochemical tracers. Southern Ocean mode waters exhibit co-variation between biogeochemical and physical properties that is consistent between ocean basins (Bushinsky and Cerovečki, 2023). Observations also show that surface-ocean nutrient concentrations are generally homogeneous in the longitudinal direction between Southern Ocean fronts (Pollard et al., 2002), with distributions instead dominated by strong variations in the latitudinal direction (Figure 3). This suggests that the time­scales of lateral eddy-driven mixing (also known as isopycnal diffusion) and transport by the ACC in the along-front direction may be fast enough to smooth out any variability arising from localized hotspots of upwelling and subduction. However, research on the three-​dimensional structure of physical–biogeochemical interactions in the Southern Ocean is in its infancy (Gray, 2024). More research is needed to robustly determine whether a two-dimensional (depth–latitude), longitudinally averaged framework is sufficient to understand the cycling of nutrients and micronutrients in the Southern Ocean.

A further recent development is broader recognition of the role that subtropical thermocline waters play in setting the (biogeochemical) properties of Southern Ocean mode waters (Iudicone et al., 2011; Morrison et al., 2022; Fernández Castro et al., 2022). Progress in our understanding of how this contribution impacts macro- and micronutrients may be made by considering biogeochemical edge cases such as the metal chromium (Cr), which has no known biological role (Fraústo da Silva and Williams, 2001) but displays a muted gradient across the fronts of the ACC that could result from mixing of subtropical and subpolar waters (Rickli et al., 2019). South of the ACC, the influence of biogeochemical cycling in the gyres of the Ross and Weddell Seas on the vertical—​and, subsequently, larger-​scale—​redistribution of macro- and micronutrients (e.g., MacGilchrist et al., 2019; Sieber et al., 2020) deserves more focused study.

When it comes to the micronutrients Zn, Cd, and Ni specifically, GEOTRACES data have not only clarified the role of the Southern Ocean hub in their marine cycles, but also raised questions that remain to be answered in the years to come.

For Zn, it is the isotopic distribution that remains most debated: will we be able to resolve whether the low δ66Zn of the low-latitude thermocline is a result of internal biogeochemical cycling or external input? Doing so will require challenging work, in the lab and in the natural environment, to robustly quantify the direction and magnitude of the isotope effect associated with binding to natural seawater ligands, provide observational constraints on the distribution coefficients of Zn sorption to different kinds of marine particles, and better characterize the isotopic composition of atmospheric sources of Zn to the ocean.

The large-scale systematics of Cd and its isotopes are less contentious, but one open question pertains to the role of the tropical ocean, including tropical oxygen-minimum zones, in modifying the Cd–PO4 relationship determined by the Southern Ocean. There appears to be a small preferential loss of seawater Cd in the shallow tropical subsurface (Ohnemus et al., 2017; Guinoiseau et al., 2019; de Souza et al., 2022; Sieber et al., 2023b), but whether this is driven by biology or redox conditions (or both, or neither) is currently unclear; a useful first step would be a more detailed characterization of the associated Cd-rich particulates (Ohnemus et al., 2019). The exact controls on the δ114Cd systematics of the thermocline also remain to be elucidated (Xie et al., 2019), and here isotope-​enabled models of ocean biogeochemistry may prove to be useful tools.

Open questions abound when it comes to the marine cycle of nickel. Is its cycle driven primarily by low-​latitude prokaryote-​dominated ecosystems (Lemaitre et al., 2022) or by high-​latitude Southern Ocean uptake as for the Zn and Cd (John et al., 2022)? And is the Southern Ocean Ni gradient driven primarily by regional uptake, or rather by the influence of Ni-poor subtropical waters transported southwards into the Subantarctic? Do prokaryotic and eukaryotic phytoplankton fractionate the isotopes of Ni differently during uptake? What is the affinity of Ni for sorption to various types of marine particles? As work proceeds at sea, in the clean lab, and on computing clusters, answers to these questions and more will emerge.

 

Acknowledgments

We would like to acknowledge Jorge Sarmiento and the stimulating interdisciplinary group of students and postdocs he brought together, which was the inspiration for this collaboration. We thank Andrew Bowie and the IMAS/ACE CRC trace metal team for the Ni data shown in Figure 3, and Toshitaka Gamo and Hajime Obata for the macronutrient data in Figure 5. GFdS is supported by ETH Zurich. AKM is supported by the Australian Research Council Australian Centre for Excellence in Antarctic Science (SR200100008). We declare no conflicts of interest. The international GEOTRACES program is possible in part thanks to the support from the US National Science Foundation (Grant OCE-2140395) to the Scientific Committee on Oceanic Research (SCOR).
Citation

de Souza, G.F., and A.K. Morrison. 2024. The Southern Ocean hub for nutrients, micronutrients, and their isotopes in the global ocean. Oceanography 37(2):46–59, https://doi.org/10.5670/oceanog.2024.414.

References

Abouchami, W., S.J.G. Galer, H.J.W. de Baar, A.-C. Alderkamp, R. Middag, P. Laan, H. Feldmann, and M.O. Andreae. 2011. Modulation of Southern Ocean cadmium isotope signature by ocean circulation and primary productivity. Earth and Planetary Science Letters 305:83–91, https://doi.org/10.1016/​j.epsl.2011.02.044.

Abouchami, W., S.J.G. Galer, H.J.W. de Baar, R. Middag, D. Vance, Y. Zhao, M. Klunder, K. Mezger, H. Feldmann, and M.O. Andreae. 2014. Biogeochemical cycling of cadmium isotopes in the Southern Ocean along the Zero Meridian. Geochimica et Cosmochimica Acta 127:348–367, https://doi.org/10.1016/j.gca.2013.10.022.

Archer, C., D. Vance, A. Milne, and M.C. Lohan. 2020. The oceanic biogeochemistry of nickel and its isotopes: New data from the South Atlantic and the Southern Ocean biogeochemical divide. Earth and Planetary Science Letters 535:116118, https://doi.org/​10.1016/j.epsl.2020.116118.

Baars, O., and P.L. Croot. 2011. The speciation of dissolved zinc in the Atlantic sector of the Southern Ocean. Deep Sea Research Part II 58(25–26):2,720–2,732, https://doi.org/​10.1016/j.dsr2.2011.02.003.

Baars, O., W. Abouchami, S.J.G. Galer, M. Boye, and P.L. Croot. 2014. Dissolved cadmium in the Southern Ocean: Distribution, speciation, and relation to phosphate. Limnology and Oceanography 59(2):385–399, https://doi.org/​10.4319/lo.2014.59.2.0385.

Baines, S.B., B.S. Twining, M.A. Brzezinski, D.M. Nelson, and N.S. Fisher. 2010. Causes and biogeochemical implications of regional differences in silicification of marine diatoms. Global Biogeochemical Cycles 24(4), https://doi.org/​10.1029/2010gb003856.

Balaguer, J., F. Koch, C. Hassler, and S. Trimborn. 2022. Iron and manganese co-limit the growth of two phytoplankton groups dominant at two locations of the Drake Passage. Communications Biology 5(1):207, https://doi.org/10.1038/s42003-​022-03148-8.

Balaguer, J., S. Thoms, and S. Trimborn. 2023. The physiological response of an Antarctic key phytoplankton species to low iron and manganese concentrations. Limnology and Oceanography 68(9):2,153–2,166, https://doi.org/10.1002/lno.12412.

Ban, Y., M. Aida, M. Nomura, and Y. Fujii. 2002. Zinc isotope separation by ligand exchange chromatography using cation exchange resin. Journal of Ion Exchange 13:46–52, https://doi.org/10.5182/jaie.13.46.

Boiteau, R.M., C.P. Till, A. Ruacho, R.M. Bundy, N.J. Hawco, A.M. McKenna, K.A. Barbeau, K.W. Bruland, M.A. Saito, and D.J. Repeta. 2016. Structural characterization of natural nickel and copper binding ligands along the US GEOTRACES Eastern Pacific Zonal Transect. Frontiers in Marine Science 3:243, https://doi.org/10.3389/fmars.2016.00243.

Bourne, H.L., J.K.B. Bishop, P.J. Lam, and D.C. Ohnemus. 2018. Global spatial and temporal variation of Cd:P in euphotic zone particulates. Global Biogeochemical Cycles 32:1,123–1,142, https://doi.org/10.1029/2017GB005842.

Boyd, P.W., A.J. Watson, C.S. Law, E.R. Abraham, T. Trull, R. Murdoch, D.C.E. Bakker, A.R. Bowie, K.O. Buesseler, H. Chang, and others. 2000. A mesoscale phytoplankton bloom in the polar Southern Ocean stimulated by iron fertilization. Nature 407(6805):695–702, https://doi.org/​10.1038/35037500.

Boyle, E.A., F. Sclater, and J.M. Edmond. 1976. On the marine geochemistry of cadmium. Nature 263(5572):42–44, https://doi.org/​10.1038/263042a0.

Boyle, E.A. 1988. Cadmium: Chemical tracer of deepwater paleoceanography. Paleoceanography 3(4):471–489, https://doi.org/​10.1029/PA003i004p00471.

Broecker, W.S., and T.H. Peng. 1982. Tracers in the Sea. Eldigio Press/Lamont-Doherty Geological Observatory, Palisades, NY, 690 pp.

Browning, T.J., E.P. Achterberg, A. Engel, and E. Mawji. 2021. Manganese co-limitation of phytoplankton growth and major nutrient drawdown in the Southern Ocean. Nature Communications 12(1):884, https://doi.org/10.1038/s41467-021-21122-6.

Bruland, K.W., G.A. Knauer, and J.H. Martin. 1978. Zinc in north-east Pacific water. Nature 271:741, https://doi.org/10.1038/271741a0.

Bruland, K.W. 1980. Oceanographic distributions of cadmium, zinc, nickel and copper in the North Pacific. Earth and Planetary Science Letters 47:176–198, https://doi.org/​10.1016/​0012-​821X(80)90035-7.

Bruland, K.W. 1983. Trace elements in sea water. Pp. 157–220 in Chemical Oceanography. J.P. Riley and R. Chester, eds, Academic Press, London.

Bruland, K.W. 1989. Complexation of zinc by natural organic ligands in the central North Pacific. Limnology and Oceanography 34(2):269–285, https://doi.org/10.4319/lo.1989.34.2.0269.

Bruland, K.W. 1992. Complexation of cadmium by natural organic ligands in the central North Pacific. Limnology and Oceanography 37(5):1,008–1,017, https://doi.org/10.4319/lo.1992.37.5.1008.

Brzezinski, M.A., and J.L. Jones. 2015. Coupling of the distribution of silicon isotopes to the meridional overturning circulation of the North Atlantic Ocean. Deep Sea Research Part II 116:79–88, https://doi.org/​10.1016/​j.dsr2.2014.11.015.

Bushinsky, S.M., and I. Cerovečki. 2023. Subantarctic Mode Water biogeochemical formation properties and interannual variability. AGU Advances 4:e2022AV000722, https://doi.org/​10.1029/2022AV000722.

Cameron, V., and D. Vance. 2014. Heavy nickel isotope compositions in rivers and the oceans. Geochimica et Cosmochimica Acta 128:195–211, https://doi.org/10.1016/j.gca.2013.12.007.

Cloete, R., J.C. Loock, T. Mtshali, S. Fietz, and A.N. Roychoudhury. 2019. Winter and summer distributions of copper, zinc and nickel along the International GEOTRACES Section GIPY05: Insights into deep winter mixing. Chemical Geology 511:342–357, https://doi.org/10.1016/​j.chemgeo.2018.10.023.

Cloete, R., J.C. Loock, N. van Horsten, S. Fietz, T.N. Mtshali, H. Planquette, and A.N. Roychoudhury. 2021. Winter biogeochemical cycling of dissolved and particulate cadmium in the Indian sector of the Southern Ocean (GEOTRACES GIpr07 transect). Frontiers in Marine Science 8:656321, https://doi.org/​10.3389/fmars.2021.656321.

Coale, K.H., K.S. Johnson, F.P. Chavez, K.O. Buesseler, R.T. Barber, M.A. Brzezinski, W.P. Cochlan, F.J. Millero, P.G. Falkowski, J.E. Bauer, and others. 2004. Southern Ocean Iron Enrichment Experiment: Carbon cycling in high- and low-Si waters. Science 304(5669):408–414, https://doi.org/​10.1126/​science.1089778.

Conway, T.M., and S.G. John. 2014. The biogeochemical cycling of zinc and zinc isotopes in the North Atlantic Ocean. Global Biogeochemical Cycles 28(10):1,111–1,128, https://doi.org/10.1002/2014GB004862.

Conway, T.M., and S.G. John. 2015. Biogeochemical cycling of cadmium isotopes along a high-​resolution section through the North Atlantic Ocean. Geochimica et Cosmochimica Acta 148(0):269–283, https://doi.org/10.1016/​j.gca.2014.09.032.

Cullen, J.T., T.W. Lane, F.M.M. Morel, and R.M. Sherrell. 1999. Modulation of cadmium uptake in phytoplankton by seawater CO2 concentration. Nature 402(6758):165–167, https://doi.org/10.1038/46007.

Cullen, J.T., Z. Chase, K.H. Coale, S.E. Fitzwater, and R.M. Sherrell. 2003. Effect of iron limitation on the cadmium to phosphorus ratio of natural phytoplankton assemblages from the Southern Ocean. Limnology and Oceanography 48(3):1,079–1,087, https://doi.org/10.4319/lo.2003.48.3.1079.

de Baar, H.J.W., P.M. Saager, R.F. Nolting, and J. van der Meer. 1994. Cadmium versus phosphate in the world ocean. Marine Chemistry 46(3):261–281, https://doi.org/​10.1016/0304-4203(94)90082-5.

De La Rocha, C.L., M.A. Brzezinski, and M.J. DeNiro. 1997. Fractionation of silicon isotopes by marine diatoms during biogenic silica formation. Geochimica et Cosmochimica Acta 61(23):5,051–5,056, https://doi.org/10.1016/S0016-7037(97)00300-1.

de Souza, G.F., B.C. Reynolds, G.C. Johnson, J.L. Bullister, and B. Bourdon. 2012a. Silicon stable isotope distribution traces Southern Ocean export of Si to the eastern South Pacific thermocline. Biogeosciences 9(11):4,199–4,213, https://doi.org/​10.5194/bg-9-4199-2012.

de Souza, G.F., B.C. Reynolds, J. Rickli, M. Frank, M.A. Saito, L.J.A. Gerringa, and B. Bourdon. 2012b. Southern Ocean control of silicon stable isotope distribution in the deep Atlantic Ocean. Global Biogeochemical Cycles 26(2), https://doi.org/​10.1029/2011gb004141.

de Souza, G.F., R.D. Slater, M.P. Hain, M.A. Brzezinski, and J.L. Sarmiento. 2015. Distal and proximal controls on the silicon stable isotope signature of North Atlantic Deep Water. Earth and Planetary Science Letters 432:342–353, https://doi.org/​10.1016/​j.epsl.2015.10.025.

de Souza, G.F., S. Khatiwala, M.P. Hain, S.H. Little, and D. Vance. 2018. On the origin of the marine zinc-silicon correlation. Earth and Planetary Science Letters 492:22–34, https://doi.org/10.1016/​j.epsl.2018.03.050.

de Souza, G.F., D. Vance, M. Sieber, T.M. Conway, and S.H. Little. 2022. Re-assessing the influence of particle-hosted sulphide precipitation on the marine cadmium cycle. Geochimica et Cosmochimica Acta 322:274–296, https://doi.org/10.1016/j.gca.​2022.02.009.

Dong, S., D.J. Weiss, S. Strekopytov, K. Kreissig, Y. Sun, A.R. Baker, and P. Formenti. 2013. Stable isotope ratio measurements of Cu and Zn in mineral dust (bulk and size fractions) from the Taklimakan Desert and the Sahel and in aerosols from the eastern tropical North Atlantic Ocean. Talanta 114:103–109, https://doi.org/10.1016/​j.talanta.2013.03.062.

Dufour, C.O., S.M. Griffies, G.F. de Souza, I. Frenger, A.K. Morrison, J.B. Palter, J.L. Sarmiento, E.D. Galbraith, J.P. Dunne, W.G. Anderson, and others. 2015. Role of mesoscale eddies in cross-​frontal transport of heat and biogeochemical tracers in the Southern Ocean. Journal of Physical Oceanography 45:3,057–3,081, https://doi.org/​10.1175/JPO-D-14-0240.1.

Dutkiewicz, S., M.J. Follows, and J.G. Bragg. 2009. Modeling the coupling of ocean ecology and biogeochemistry. Global Biogeochemical Cycles 23(4), https://doi.org/10.1029/2008gb003405.

Egleston, E.S., and F.M.M. Morel. 2008. Nickel limitation and zinc toxicity in a urea-grown diatom. Limnology and Oceanography 53(6):2,462–2,471, https://doi.org/10.4319/lo.2008.53.6.2462.

Ellwood, M.J., and K.A. Hunter. 2000. The incorporation of zinc and iron into the frustule of the marine diatom Thalassiosira pseudonana. Limnology and Oceanography 45(7):1,517–1,524, https://doi.org/​10.4319/lo.2000.45.7.1517.

Ellwood, M.J. 2004. Zinc and cadmium speciation in Subantarctic waters east of New Zealand. Marine Chemistry 87:37–58, https://doi.org/10.1016/​j.marchem.2004.01.005.

Ellwood, M.J. 2008. Wintertime trace metal (Zn, Cu, Ni, Cd, Pb and Co) and nutrient distributions in the Subantarctic Zone between 40–52°S; 155–160°E. Marine Chemistry 112(1-2):107–117, https://doi.org/​10.1016/j.marchem.2008.07.008.

Ellwood, M.J., R. Strzepek, X. Chen, T.W. Trull, and P.W. Boyd. 2020. Some observations on the biogeochemical cycling of zinc in the Australian sector of the Southern Ocean: A dedication to Keith Hunter. Marine and Freshwater Research 71(3):355–373, https://doi.org/10.1071/MF19200.

Fernández Castro, B., M. Mazloff, R.G. Williams, A.C. Naveira Garabato. 2022. Subtropical contribution to sub-Antarctic mode waters. Geophysical Research Letters 49(11):e2021GL097560, https://doi.org/​10.1029/2021GL097560.

Foppert, A., K.A. Donohue, D.R. Watts, and K.L. Tracey. 2017. Eddy heat flux across the Antarctic Circumpolar Current estimated from sea surface height standard deviation. Journal of Geophysical Research: Oceans 122:6,947–6,964. https://doi.org/10.1002/2017jc012837.

Fraústo da Silva, J.J.R., and R.J.P. Williams. 2001. The Biological Chemistry of the Elements, 2nd ed. Oxford University Press, 600 pp.

Frew, R.D., and K.A. Hunter. 1992. Influence of Southern Ocean waters on the cadmium-​phosphate properties of the global ocean. Nature 360:144–146, https://doi.org/​10.1038/​360144a0.

Fripiat, F., A.-J. Cavagna, F. Dehairs, S. Speich, L. André, and D. Cardinal. 2011. Silicon pool dynamics and biogenic silica export in the Southern Ocean, inferred from Si-isotopes. Ocean Science 7:533–547, https://doi.org/​10.5194/os-7-533-2011.

Fripiat, F., A. Martínez-García, D. Marconi, S.E. Fawcett, S.H. Kopf, V.H. Luu, P.A. Rafter, R. Zhang, D.M. Sigman, and G.H. Haug. 2021. Nitrogen isotopic constraints on nutrient transport to the upper ocean. Nature Geoscience 14(11):855–861, https://doi.org/10.1038/s41561-021-00836-8.

Fripiat, F., D.M. Sigman, A. Martínez-García, D. Marconi, X.E. Ai, A. Auderset, S.E. Fawcett, S. Moretti, A.S. Studer, and G.H. Haug. 2023. The impact of incomplete nutrient consumption in the Southern Ocean on global mean ocean nitrate δ15N. Global Biogeochemical Cycles 37(2):e2022GB007442, https://doi.org/​10.1029/2022GB007442.

George, E., C.H. Stirling, M. Gault-Ringold, M.J. Ellwood, and R. Middag. 2019. Marine biogeochemical cycling of cadmium and cadmium isotopes in the extreme nutrient-depleted subtropical gyre of the South West Pacific Ocean. Earth and Planetary Science Letters 514:84–95, https://doi.org/10.1016/j.epsl.2019.02.031.

GEOTRACES Intermediate Data Product Group. 2021. The GEOTRACES Intermediate Data Product 2021 (IDP2021). NERC EDS British Oceanographic Data Centre NOC, https://doi.org/10/g55p.

Gervais, F., U. Riebesell, and M.Y. Gorbunov. 2002. Changes in primary productivity and chlorophyll a in response to iron fertilization in the Southern Polar Frontal Zone. Limnology and Oceanography 47(5):1,324–1,335, https://doi.org/​10.4319/lo.2002.47.5.1324.

Grasse, P., I. Closset, J.L. Jones, S. Geilert, and M.A. Brzezinski. 2020. Controls on dissolved silicon isotopes along the US GEOTRACES Eastern Pacific Zonal Transect (GP16). Global Biogeochemical Cycles 34(9):e2020GB006538, https://doi.org/​10.1029/2020gb006538.

Gray, A.R. 2024. The four-dimensional carbon cycle of the Southern Ocean. Annual Review of Marine Science 16:163–190, https://doi.org/10.1146/annurev-marine-041923-104057.

Grun, R., M. Samanta, and M.J. Ellwood. 2023. Variability in zinc:phosphorous and zinc:silicon ratios and zinc isotope fractionation in Southern Ocean diatoms: Observations from laboratory and field experiments. Marine Chemistry 257:104330, https://doi.org/10.1016/j.marchem.2023.104330.

Guinoiseau, D., S.J.G. Galer, W. Abouchami, M. Frank, E.P. Achterberg, and G.H. Haug. 2019. Importance of cadmium sulfides for biogeochemical cycling of Cd and its isotopes in oxygen deficient zones—​a case study of the Angola Basin. Global Biogeochemical Cycles 33(12):1,746–1,763, https://doi.org/​10.1029/2019GB006323.

Gupta, M., R.G. Williams, J.M. Lauderdale, O. Jahn, C. Hill, S. Dutkiewicz, M.J. Follows. 2022. A nutrient relay sustains subtropical ocean productivity. Proceedings of the National Academy of Sciences of the United States of America 119:e2206504119, https://doi.org/10.1073/pnas.2206504119.

Ho, T.-Y., A. Quigg, Z.V. Finkel, A.J. Milligan, K. Wyman, P.G. Falkowski, and F.M.M. Morel. 2003. The elemental composition of some marine phytoplankton. Journal of Phycology 39(6):1,145–1,159, https://doi.org/​10.1111/j.0022-3646.2003.03-090.x.

Holzer, M., and F.W. Primeau. 2013. Global teleconnections in the oceanic phosphorus cycle: Patterns, paths, and timescales. Journal of Geophysical Research: Oceans 118(4):1,775–1,796, https://doi.org/​10.1002/​jgrc.20072.

Holzer, M., F.W. Primeau, T. DeVries, and R. Matear. 2014. The Southern Ocean silicon trap: Data-constrained estimates of regenerated silicic acid, trapping efficiencies, and global transport paths. Journal of Geophysical Research: Oceans 119:1–19, https://doi.org/10.1002/2013jc009356.

Holzer, M., and M.A. Brzezinski. 2015. Controls on the silicon isotope distribution in the ocean: New diagnostics from a data-constrained model. Global Biogeochemical Cycles 29(3):267–287, https://doi.org/​10.1002/​2014GB004967.

Iudicone, D., K.B. Rodgers, I. Stendardo, O. Aumont, G. Madec, L. Bopp, O. Mangoni, and M. Ribera d’Alcala. 2011. Water masses as a unifying framework for understanding the Southern Ocean carbon cycle. Biogeosciences 8:1,031–1,052, https://doi.org/10.5194/bg-8-1031-2011.

Janssen, D.J., M. Sieber, M.J. Ellwood, T.M. Conway, P.M. Barrett, X. Chen, G.F. de Souza, C.S. Hassler, and S.L. Jaccard. 2020. Trace metal and nutrient dynamics across broad biogeochemical gradients in the Indian and Pacific sectors of the Southern Ocean. Marine Chemistry 221:103773, https://doi.org/​10.1016/​j.marchem.2020.103773.

John, S.G., R.W. Geis, M.A. Saito, and E.A. Boyle. 2007. Zinc isotope fractionation during high-​affinity and low-affinity zinc transport by the marine diatom Thalassiosira oceanica. Limnology and Oceanography 52(6):2,710–2,714, https://doi.org/​10.4319/​lo.2007.52.6.2710.

John, S.G., and T.M. Conway. 2014. A role for scavenging in the marine biogeochemical cycling of zinc and zinc isotopes. Earth and Planetary Science Letters 394:159–167, https://doi.org/​10.1016/j.epsl.2014.02.053.

John, S.G., J. Helgoe, and E. Townsend. 2018. Biogeochemical cycling of Zn and Cd and their stable isotopes in the Eastern Tropical South Pacific. Marine Chemistry 201:256–262, https://doi.org/​10.1016/j.marchem.2017.06.001.

John, S.G., R.L. Kelly, X. Bian, F. Fu, M.I. Smith, N.T. Lanning, H. Liang, B. Pasquier, E.A. Seelen, M. Holzer, and others. 2022. The biogeochemical balance of oceanic nickel cycling. Nature Geoscience 15(11):906–912, https://doi.org/10.1038/s41561-022-01045-7.

Joubert, W.R., S.J. Thomalla, H.N. Waldron, M.I. Lucas, M. Boye, F.A.C. Le Moigne, F. Planchon, and S. Speich. 2011. Nitrogen uptake by phytoplankton in the Atlantic sector of the Southern Ocean during late austral summer. Biogeosciences 8(10):2,947–2,959, https://doi.org/​10.5194/​bg-8-2947-2011.

Kim, T., H. Obata, Y. Kondo, H. Ogawa, and T. Gamo. 2015. Distribution and speciation of dissolved zinc in the western North Pacific and its adjacent seas. Marine Chemistry 173:330–341, https://doi.org/​10.1016/j.marchem.2014.10.016.

Köbberich, M., and D. Vance. 2019. Zn isotope fractionation during uptake into marine phytoplankton: Implications for oceanic zinc isotopes. Chemical Geology 523:154–161, https://doi.org/10.1016/​j.chemgeo.2019.04.004.

Landing, W.M., and K.W. Bruland. 1987. The contrasting biogeochemistry of iron and manganese in the Pacific Ocean. Geochimica et Cosmochimica Acta 51(1):29–43, https://doi.org/​10.1016/0016-7037(87)90004-4.

Lane, T.W., M.A. Saito, G.N. George, I.J. Pickering, R.C. Prince, and F.M.M. Morel. 2005. A cadmium enzyme from a marine diatom. Nature 435:42–42, https://doi.org/10.1038/435042a.

Latour, P., K. Wuttig, P. van der Merwe, R.F. Strzepek, M. Gault-Ringold, A.T. Townsend, T.M. Holmes, M. Corkill, and A.R. Bowie. 2021. Manganese biogeochemistry in the Southern Ocean, from Tasmania to Antarctica. Limnology and Oceanography 66(6):2,547–2,562, https://doi.org/​10.1002/​lno.11772.

Lee, J.G., and F.M.M. Morel. 1995. Replacement of zinc by cadmium in marine phytoplankton. Marine Ecology Progress Series 127:305–309, https://doi.org/​10.3354/​meps127305.

Lee, J.G., S.B. Roberts, and F.M.M. Morel. 1995. Cadmium: A nutrient for the marine diatom Thalassiosira weissflogii. Limnology and Oceanography 40:1,056–1,063, https://doi.org/​10.4319/lo.1995.40.6.1056.

Lemaitre, N., G.F. de Souza, C. Archer, R.-M. Wang, H. Planquette, G. Sarthou, and D. Vance. 2020. Pervasive sources of isotopically light zinc in the North Atlantic Ocean. Earth and Planetary Science Letters 539:116216, https://doi.org/10.1016/​j.epsl.2020.116216.

Lemaitre, N., J. Du, G.F. de Souza, C. Archer, and D. Vance. 2022. The essential bioactive role of nickel in the oceans: Evidence from nickel isotopes. Earth and Planetary Science Letters 584:117513, https://doi.org/10.1016/j.epsl.2022.117513.

Liao, W.-H., S. Takano, S.-C. Yang, K.-F. Huang, Y. Sohrin, and T.-Y. Ho. 2020. Zn isotope composition in the water column of the northwestern Pacific Ocean: The importance of external sources. Global Biogeochemical Cycles 34(1):e2019GB006379, https://doi.org/10.1029/2019GB006379.

MacGilchrist, G.A., A.C. Naveira Garabato, P.J. Brown, L. Jullion, S. Bacon, D.C.E. Bakker, M. Hoppema, M.P. Meredith, and S. Torres-Valdés. 2019. Reframing the carbon cycle of the subpolar Southern Ocean. Science Advances 5:eaav6410, https://doi.org/10.1126/sciadv.aav6410.

Mackey, D.J., J.E. O’Sullivan, R.J. Watson, and G. Dal Pont. 2002. Trace metals in the Western Pacific: Temporal and spatial variability in the concentrations of Cd, Cu, Mn and Ni. Deep Sea Research Part I 49(12):2,241–2,259, https://doi.org/​10.1016/S0967-0637(02)00124-3.

Marconi, D., M.A. Weigand, P.A. Rafter, M. McIlvin, M. Forbes, K.L. Casciotti, and D.M. Sigman. 2015. Nitrate isotope distributions on the US GEOTRACES North Atlantic cross-basin section: Signals of polar nitrate sources and low latitude nitrogen cycling. Marine Chemistry 177:143–156, https://doi.org/10.1016/j.marchem.2015.06.007.

Margalef, R. 1978. Life-forms of phytoplankton as survival alternatives in an unstable environment. Oceanologica Acta 1(4):493–509.

Marinov, I., A. Gnanadesikan, J.R. Toggweiler, and J.L. Sarmiento. 2006. The Southern Ocean biogeochemical divide. Nature 441(7096):964–967, https://doi.org/10.1038/nature04883.

Marković, T., S. Manzoor, E. Humphreys-Williams, G.J.D. Kirk, R. Vilar, and D.J. Weiss. 2017. Experimental determination of zinc isotope fractionation in complexes with the phytosiderophore 2'-deoxymugeneic acid (DMA) and its structural analogues, and implications for plant uptake mechanisms. Environmental Science & Technology 51:98–107, https://doi.org/10.1021/​acs.est.6b00566.

Marshall, J., and K. Speer. 2012. Closure of the meridional overturning circulation through Southern Ocean upwelling. Nature Geoscience 5(3):171–180, https://doi.org/10.1038/ngeo1391.

Martin, J.H., R.M. Gordon, and S.E. Fitzwater. 1990. Iron in Antarctic waters. Nature 345(6271):156–158, https://doi.org/10.1038/345156a0.

Middag, R., S.M.A.C. van Heuven, K.W. Bruland, and H.J.W. de Baar. 2018. The relationship between cadmium and phosphate in the Atlantic Ocean unravelled. Earth and Planetary Science Letters 492:79–88, https://doi.org/10.1016/​j.epsl.2018.03.046.

Middag, R., H.J.W. De Baar, and K.W. Bruland. 2019. The relationships between dissolved zinc and major nutrients phosphate and silicate along the GEOTRACES GA02 transect in the west Atlantic Ocean. Global Biogeochemical Cycles 33(1):63–84, https://doi.org/10.1029/2018GB006034.

Middag, R., H.J.W. de Baar, K.W. Bruland, and S.M.A.C. van Heuven. 2020. The distribution of nickel in the West-Atlantic Ocean, its relationship with phosphate and a comparison to cadmium and zinc. Frontiers in Marine Science 7:105, https://doi.org/​10.3389/fmars.2020.00105.

Moore, C.M., M.M. Mills, K.R. Arrigo, I. Berman-Frank, L. Bopp, P. Boyd, E.D. Galbraith, R.J. Geider, C. Guieu, S.L. Jaccard, and others. 2013. Processes and patterns of oceanic nutrient limitations. Nature Geoscience 6:701–710, https://doi.org/10.1038/ngeo1765.

Morel, F.M.M., J.R. Reinfelder, S.B. Roberts, C.P. Chamberlain, J.G. Lee, and D. Yee. 1994. Zinc and carbon co-limitation of marine phytoplankton. Nature 369(6483):740–742, https://doi.org/​10.1038/369740a0.

Morel, F.M.M., A.J. Milligan, and M.A. Saito. 2014. Marine bioinorganic chemistry: The role of trace metals in the oceanic cycles of major nutrients. Pp. 123–150 in Treatise on Geochemistry (Second Edition). H.D. Holland, and K.K. Turekian, eds, Elsevier, Oxford.

Morrison, A.K., T.L. Fröhlicher, and J.L. Sarmiento. 2015. Upwelling in the Southern Ocean. Physics Today 68(1):27–32, https://doi.org/10.1063/PT.3.2654.

Morrison, A.K., D.W. Waugh, A. McC. Hogg, D.C. Jones, and R.P. Abernathey. 2022. Ventilation of the Southern Ocean pycnocline. Annual Review of Marine Science 14:405–430, https://doi.org/​10.1146/annurev-marine-010419-011012.

Ohnemus, D.C., S. Rauschenberg, G.A. Cutter, J.N. Fitzsimmons, R.M. Sherrell, and B.S. Twining. 2017. Elevated trace metal content of prokaryotic communities associated with marine oxygen deficient zones. Limnology and Oceanography 62:3–25, https://doi.org/10.1002/lno.10363.

Ohnemus, D.C., R. Torrie, and B.S. Twining. 2019. Exposing the distributions and elemental associations of scavenged particulate phases in the ocean using basin-scale multi-element data sets. Global Biogeochemical Cycles 33:725–748, https://doi.org/​10.1029/2018GB006145.

Packman, H., S.H. Little, A.R. Baker, L. Bridgestock, R.J. Chance, B.J. Coles, K. Kreissig, M. Rehkämper, and T. van de Flierdt. 2022. Tracing natural and anthropogenic sources of aerosols to the Atlantic Ocean using Zn and Cu isotopes. Chemical Geology 610:121091, https://doi.org/10.1016/​j.chemgeo.2022.121091.

Pausch, F., K. Bischof, and S. Trimborn. 2019. Iron and manganese co-limit growth of the Southern Ocean diatom Chaetoceros debilis. PLOS ONE 14(9):e0221959, https://doi.org/​10.1371/journal.pone.0221959.

Prend, C.J., A.R. Gray, L.D. Talley, S.T. Gille, F.A. Haumann, K.S. Johnson, S.C. Riser, I. Rosso, J. Sauvé, and J.L. Sarmiento. 2022. Indo-Pacific sector dominates Southern Ocean carbon outgassing. Global Biogeochemical Cycles 36(7):e2021GB007226, https://doi.org/​10.1029/2021gb007226.

Price, N.M., and F.M.M. Morel. 1991. Colimitation of phytoplankton growth by nickel and nitrogen. Limnology and Oceanography 36(6):1,071–1,077, https://doi.org/10.4319/lo.1991.36.6.1071.

Primeau, F., M. Holzer, and T. DeVries. 2013. Southern Ocean nutrient trapping and the efficiency of the biological pump. Journal of Geophysical Research: Oceans 118(5):2,547–2,564, https://doi.org/10.1002/jgrc.20181.

Pollard, R.T., M.I. Lucas, and J.F. Read. 2002. Physical controls on biogeochemical zonation in the Southern Ocean. Deep Sea Research Part II 49(16):3,289–3,305, https://doi.org/​10.1016/S0967-0645(02)00084-X.

Quay, P., J.D. Cullen, W.M. Landing, and P. Morton. 2015. Processes controlling the distributions of Cd and PO4 in the ocean. Global Biogeochemical Cycles 29:830–841, https://doi.org/​10.1002/​2014GB004998.

Rafter, P.A., P.J. DiFiore, and D.M. Sigman. 2013. Coupled nitrate nitrogen and oxygen isotopes and organic matter remineralization in the Southern and Pacific Oceans. Journal of Geophysical Research: Oceans 118(10):4,781–4,794, https://doi.org/10.1002/jgrc.20316.

Ragsdale, S.W. 2009. Nickel-based enzyme systems. Journal of Biological Chemistry 284(28):18,571–18,575, https://doi.org/​10.1074/jbc.R900020200.

Rickli, J., D.J. Janssen, C. Hassler, M.J. Ellwood, and S.L. Jaccard. 2019. Chromium biogeochemistry and stable isotope distribution in the Southern Ocean. Geochimica et Cosmochimica Acta 262:188–206, https://doi.org/10.1016/j.gca.2019.07.033.

Roshan, S., T. DeVries, J. Wu, and G. Chen. 2018. The internal cycling of zinc in the ocean. Global Biogeochemical Cycles 32:1,833–1,849, https://doi.org/​10.1029/​2018GB006045.

Roshan, S., and T. DeVries. 2021. Global contrasts between oceanic cycling of cadmium and phosphate. Global Biogeochemical Cycles 35:e2021GB006952, https://doi.org/​10.1029/​2021GB006952.

Saito, M.A., D.M. Sigman, and F.M.M. Morel. 2003. The bioinorganic chemistry of the ancient ocean: The co-evolution of cyanobacterial metal requirements and biogeochemical cycles at the Archean-Proterozoic boundary. Inorganica Chimica Acta 356:308–318, https://doi.org/10.1016/S0020-​1693(03)00442-0.

Sallée, J.-B., K. Speer, S. Rintoul, and S. Wijffels. 2010. Southern Ocean thermocline ventilation. Journal of Physical Oceanography 40(3):509–529, https://doi.org/​10.1175/​2009JPO4291.1.

Samanta, M., M.J. Ellwood, M. Sinoir, and C.S. Hassler. 2017. Dissolved zinc isotope cycling in the Tasman Sea, SW Pacific Ocean. Marine Chemistry 192:1–12, https://doi.org/10.1016/j.marchem.2017.03.004.

Samanta, M., M.J. Ellwood, and R. Strzepek. 2018. Zinc isotope fractionation by Emiliania huxleyi cultured across a range of free zinc ion concentrations. Limnology and Oceanography 63:660–671, https://doi.org/10.1002/lno.10658.

Sarmiento, J.L., N. Gruber, M.A. Brzezinski, and J.P. Dunne. 2004. High-latitude controls of thermocline nutrients and low latitude biological productivity. Nature 427(6969):56–60, https://doi.org/​10.1038/​nature02127.

Sarmiento, J.L., J. Simeon, A. Gnanadesikan, N. Gruber, R.M. Key, and R. Schlitzer. 2007. Deep ocean biogeochemistry of silicic acid and nitrate. Global Biogeochemical Cycles 21(1), https://doi.org/​10.1029/​2006GB002720.

Sclater, F.R., E. Boyle, and J.M. Edmond. 1976. On the marine geochemistry of nickel. Earth and Planetary Science Letters 31(1):119–128, https://doi.org/​10.1016/​0012-821X(76)90103-5.

Sieber, M., T.M. Conway, G.F. de Souza, H. Obata, S. Takano, Y. Sohrin, and D. Vance. 2019a. Physical and biogeochemical controls on the distribution of dissolved cadmium and its isotopes in the Southwest Pacific Ocean. Chemical Geology 511:494–509, https://doi.org/10.1016/​j.chemgeo.2018.07.021.

Sieber, M., T.M. Conway, G.F. de Souza, C.S. Hassler, M. Ellwood, and D. Vance. 2019b. High-resolution Cd isotope systematics in multiple zones of the Southern Ocean from the Antarctic Circumnavigation Expedition. Earth and Planetary Science Letters 527:115799, https://doi.org/10.1016/​j.epsl.2019.115799.

Sieber, M., T.M. Conway, G.F. de Souza, M.J. Ellwood, and D. Vance. 2020. Cycling of zinc and its isotopes across multiple zones of the Southern Ocean: Insights from the Antarctic Circumnavigation Expedition. Geochimica et Cosmochimica Acta 268:310–324, https://doi.org/​10.1016/​j.gca.2019.09.039.

Sieber, M., N.T. Lanning, X. Bian, S.-C. Yang, S. Takano, Y. Sohrin, T.S. Weber, J.N. Fitzsimmons, S.G. John, and T.M. Conway. 2023a. The importance of reversible scavenging for the marine Zn cycle evidenced by the distribution of zinc and its isotopes in the Pacific Ocean. Journal of Geophysical Research: Oceans 128(4):e2022JC019419, https://doi.org/10.1029/2022JC019419.

Sieber, M., N.T. Lanning, Z.B. Bunnell, X. Bian, S.-C. Yang, C.M. Marsay, W.M. Landing, C.S. Buck, J.N. Fitzsimmons, S.G. John, and others. 2023b. Biological, physical, and atmospheric controls on the distribution of cadmium and its isotopes in the Pacific Ocean. Global Biogeochemical Cycles 37(2):e2022GB007441, https://doi.org/​10.1029/2022GB007441.

Sigman, D.M., M.A. Altabet, D.C. McCorkle, R. François, and G. Fischer. 1999. The δ15N of nitrate in the Southern Ocean: Consumption of nitrate in surface waters. Global Biogeochemical Cycles 13(4):1,149–1,166, https://doi.org/​10.1029/​1999GB900038.

Sigman, D.M., M.A. Altabet, D.C. McCorkle, R. François, and G. Fischer. 2000. The δ15N of nitrate in the Southern Ocean: Nitrogen cycling and circulation in the ocean interior. Journal of Geophysical Research: Oceans 105(C8):19,599–19,614, https://doi.org/​10.1029/2000JC000265.

Sunda, W.G., and R.R.L. Guillard. 1976. The relationship between cupric ion activity and the toxicity of copper to phytoplankton. Journal of Marine Research 34:511–529.

Sunda, W.G., and S.A. Huntsman. 1987. Microbial oxidation of manganese in a North Carolina estuary. Limnology and Oceanography 32(3):552–564, https://doi.org/10.4319/lo.1987.32.3.0552.

Sunda, W.G., and S.A. Huntsman. 1992. Feedback interactions between zinc and phytoplankton in seawater. Limnology and Oceanography 37(1):25–40, https://doi.org/​10.4319/lo.1992.37.1.0025.

Sunda, W.G., and S.A. Huntsman. 1995. Cobalt and zinc interreplacement in marine phytoplankton: Biological and geochemical implications. Limnology and Oceanography 40:1,404–1,417, https://doi.org/10.4319/lo.1995.40.8.1404.

Sunda, W.G., and S.A. Huntsman. 1998. Control of Cd concentrations in a coastal diatom by interactions among free ionic Cd, Zn and Mn in seawater. Environmental Science & Technology 32:2,961–2,968, https://doi.org/​10.1021/es980271y.

Sunda, W.G., and S.A. Huntsman. 2000. Effect of Zn, Mn, and Fe on Cd accumulation in phytoplankton: Implications for oceanic Cd cycling. Limnology and Oceanography 45(7):1,501–1,516, https://doi.org/​10.4319/lo.2000.45.7.1501.

Sunda, W.G. 2012. Feedback interactions between trace metal nutrients and phytoplankton in the ocean. Frontiers in Microbiology 3:204, https://doi.org/​10.3389/fmicb.2012.00204.

Tagliabue, A., A. Bowie, P. Boyd, K. Buck, K. Johnson, and M. Saito. 2017. The integral role of iron in ocean biogeochemistry. Nature 543:51–59, https://doi.org/10.1038/nature21058.

Tagliabue, A., K.N. Buck, L.E. Sofen, B.S. Twining, O. Aumont, P.W. Boyd, S. Caprara, W.B. Homoky, R. Johnson, D. König, and others. 2023. Authigenic mineral phases as a driver of the upper-ocean iron cycle. Nature 620(7972):104–109, https://doi.org/​10.1038/s41586-023-06210-5.

Takano, S., M. Tanimizu, T. Hirata, K.-C. Shin, Y. Fukami, K. Suzuki, and Y. Sohrin. 2017. A simple and rapid method for isotopic analysis of nickel, copper, and zinc in seawater using chelating extraction and anion exchange. Analytica Chimica Acta 967:1–11, https://doi.org/10.1016/​j.aca.2017.03.010.

Talley, L.D. 2013. Closure of the global overturning circulation through the Indian, Pacific, and Southern Oceans: Schematics and transports. Oceanography 26(1):80–97, https://doi.org/​10.5670/oceanog.2013.07.

Tamsitt, V., H.F. Drake, A.K. Morrison, L.D. Talley, C.O. Dufour, A.R. Gray, S.M. Griffies, M.R. Mazloff, J.L. Sarmiento, J. Wang, and W. Weijer. 2017. Spiraling pathways of global deep waters to the surface of the Southern Ocean. Nature Communications 8:172, https://doi.org/10.1038/s41467-017-00197-0.

Toggweiler, J.R. 1994. The ocean’s overturning circulation. Physics Today 47(11):45–50, https://doi.org/​10.1063/1.881425.

Twining, B.S., and S.B. Baines. 2013. The trace metal composition of marine phytoplankton. Annual Review of Marine Science 5:191–215, https://doi.org/​10.1146/annurev-marine-121211-172322.

Twining, B.S., S.D. Nodder, A.L. King, D.A. Hutchins, G.R. LeCleir, J.M. DeBruyn, E.W. Maas, S. Vogt, S.W. Wilhelm, and P.W. Boyd. 2014. Differential remineralization of major and trace elements in sinking diatoms. Limnology and Oceanography 59:689–704, https://doi.org/​10.4319/lo.2014.59.3.0689.

Twining, B.S., S. Rauschenberg, P.L. Morton, and S. Vogt. 2015. Metal contents of phytoplankton and labile particulate material in the North Atlantic Ocean. Progress in Oceanography 137:261–283, https://doi.org/10.1016/j.pocean.2015.07.001.

Vance, D., S.H. Little, G.F. de Souza, S. Khatiwala, M.C. Lohan, and R. Middag. 2017. Silicon and zinc biogeochemical cycles coupled through the Southern Ocean. Nature Geoscience 10:202–206, https://doi.org/10.1038/ngeo2890.

Varela, D.E., C.J. Pride, and M.A. Brzezinski. 2004. Biological fractionation of silicon isotopes in Southern Ocean surface waters. Global Biogeochemical Cycles 18(1), https://doi.org/​10.1029/​2003GB002140.

Varela, D.E., M.A. Brzezinski, C.P. Beucher, J.L. Jones, K.E. Giesbrecht, B. Lansard, and A. Mucci. 2016. Heavy silicon isotopic composition of silicic acid and biogenic silica in Arctic waters over the Beaufort shelf and the Canada Basin. Global Biogeochemical Cycles 30(6):804–824, https://doi.org/​10.1002/2015GB005277.

Viglione, G.A., and A.F. Thompson. 2016. Lagrangian pathways of upwelling in the Southern Ocean. Journal of Geophysical Research: Oceans 121:6,295–6,309, https://doi.org/​10.1002/2016jc011773.

Wada, E., and A. Hattori. 1978. Nitrogen isotope effects in the assimilation of inorganic nitrogenous compounds by marine diatoms. Geomicrobiology Journal 1(1):85–101, https://doi.org/​10.1080/​01490457809377725.

Wang, R.-M., C. Archer, A.R. Bowie, and D. Vance. 2019. Zinc and nickel isotopes in seawater from the Indian sector of the Southern Ocean: The impact of natural iron fertilization versus Southern Ocean hydrography and biogeochemistry. Chemical Geology 511:452–464, https://doi.org/10.1016/​j.chemgeo.2018.09.010.

Weber, T.S., and C. Deutsch. 2010. Ocean nutrient ratios governed by plankton biogeography. Nature 467(7315):550–554, https://doi.org/​10.1038/nature09403.

Weber, T., S.G. John, A. Tagliabue, and T. DeVries. 2018. Biological uptake and reversible scavenging of zinc in the global ocean. Science 361:72–76, https://doi.org/10.1126/science.aap8532.

Wu, M., J.S.P. McCain, E. Rowland, R. Middag, M. Sandgren, A.E. Allen, and E.M. Bertrand. 2019. Manganese and iron deficiency in Southern Ocean Phaeocystis antarctica populations revealed through taxon-specific protein indicators. Nature Communications 10(1):3582, https://doi.org/10.1038/s41467-019-11426-z.

Wyatt, N.J., A. Milne, E.M.S. Woodward, A.P. Rees, T.J. Browning, H.A. Bouman, P.J. Worsfold, and M.C. Lohan. 2014. Biogeochemical cycling of dissolved zinc along the GEOTRACES South Atlantic transect GA10 at 40°S. Global Biogeochemical Cycles 28(1), https://doi.org/10.1002/2013gb004637.

Xie, R.C., S.J.G. Galer, W. Abouchami, M.J.A. Rijkenberg, H.J.W. de Baar, J. De Jong, and M.O. Andreae. 2017. Non-Rayleigh control of upper-ocean Cd isotope fractionation in the western South Atlantic. Earth and Planetary Science Letters 471:94–103, https://doi.org/10.1016/​j.epsl.2017.04.024.

Xie, R.C., M. Rehkämper, P. Grasse, T. van de Flierdt, M. Frank, and Z. Xue. 2019. Isotopic evidence for complex biogeochemical cycling of Cd in the eastern tropical South Pacific. Earth and Planetary Science Letters 512:134–146, https://doi.org/​10.1016/j.epsl.2019.02.001.

Xu, Y., D. Tang, Y. Shaked, and F.M.M. Morel. 2007. Zinc, cadmium, and cobalt interreplacement and relative use efficiencies in the coccolithophore Emiliania huxleyi. Limnology and Oceanography 52(5):2,294–2,305, https://doi.org/10.4319/lo.2007.52.5.2294.

Xue, Z., M. Rehkämper, T.J. Horner, W. Abouchami, R. Middag, T. van de Flierdt, and H.J.W. de Baar. 2013. Cadmium isotope variations in the Southern Ocean. Earth and Planetary Science Letters 382:161–172, https://doi.org/10.1016/​j.epsl.2013.09.014.

Yang, S.-C., J. Zhang, Y. Sohrin, and T.-Y. Ho. 2018. Cadmium cycling in the water column of the Kuroshio-Oyashio Extension region: Insights from dissolved and particulate isotopic composition. Geochimica et Cosmochimica Acta 233:66–80, https://doi.org/10.1016/j.gca.2018.05.001.

Yang, S.-C., R.L. Kelly, X. Bian, T.M. Conway, K.-F. Huang, T.-Y. Ho, J.A. Neibauer, R.G. Keil, J.W. Moffett, and S.G. John. 2021. Lack of redox cycling for nickel in the water column of the eastern tropical North Pacific oxygen deficient zone: Insight from dissolved and particulate nickel isotopes. Geochimica et Cosmochimica Acta 309:235–250, https://doi.org/10.1016/​j.gca.2021.07.004.

Yung, C.K., A.K. Morrison, and A. McC. Hogg. 2022. Topographic hotspots of Southern Ocean eddy upwelling. Frontiers in Marine Science 9:855785, https://doi.org/10.3389/fmars.2022.855785.

Zhang, X., N. Lemaitre, J.D. Rickli, T.J. Suhrhoff, R. Shelley, A. Benhra, S. Faye, M.A. Jeyid, and D. Vance. 2024. Tracing anthropogenic aerosol trace metal sources in the North Atlantic Ocean using Pb, Zn and Ni isotopes. Marine Chemistry 258:104347, https://doi.org/10.1016/​j.marchem.2023.104347.

Zhao, Y., D. Vance, W. Abouchami, and H.J.W. de Baar. 2014. Biogeochemical cycling of zinc and its isotopes in the Southern Ocean. Geochimica et Cosmochimica Acta 125:653–672, https://doi.org/​10.1016/j.gca.2013.07.045.

Zheng, L., T. Minami, S. Takano, T.-Y. Ho, and Y. Sohrin. 2021. Sectional distribution patterns of Cd, Ni, Zn, and Cu in the North Pacific Ocean: Relationships to nutrients and importance of scavenging. Global Biogeochemical Cycles 35(7):e2020GB006558, https://doi.org/10.1029/2020GB006558.

Copyright & Usage

This is an open access article made available under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution, and reproduction in any medium or format as long as users cite the materials appropriately, provide a link to the Creative Commons license, and indicate the changes that were made to the original content. Images, animations, videos, or other third-party material used in articles are included in the Creative Commons license unless indicated otherwise in a credit line to the material. If the material is not included in the article’s Creative Commons license, users will need to obtain permission directly from the license holder to reproduce the material.