Dissolved organic carbon (DOC) produced by primary production in the sunlit ocean can be physically transported to the mesopelagic zone. The majority of DOC exported to this zone is remineralized by heterotrophic microbes over a range of timescales. Capturing a deep convective mixing event is rare, as is observing how microbes respond in situ to the exported DOC. Here, we report ship and Argo float observations from hydrostation North Atlantic Aerosols and Marine Ecosystems Study (NAAMES) 2 Station 4 (N2S4; 47.46°N, 38.72°W), a retentive anticyclonic eddy in the subtropical region of the western North Atlantic. Changes in biogeochemistry and bacterioplankton responses were tracked as the water column mixed to approximately 230 m and restratified over the subsequent 3 days. Over this period, rapid changes in bacterioplankton production (BP) and cell abundance were observed throughout the water column. BP increased by 91% in the euphotic zone (0–100 m) and 55% in the upper mesopelagic zone (100–200 m), corresponding to 33% and 103% increases in cell abundance, respectively. Within the upper mesopelagic, BP upon the occupation of N2S4 (20 ± 4.7 nmol C L–1 d–1) was significantly greater than the average upper mesopelagic BP rate (2.0 ± 1.6 nmol C L–1 h–1) at other stations that had been stratified for longer periods of time. BP continued to increase to 31 ± 3.0 nmol C L–1 d–1 over the 3-day occupation of N2S4. The rapid changes in BP in the upper mesopelagic did not coincide with rapid changes in community composition, but the taxa that increased in their relative contribution included those typically observed in the epipelagic zone. We interpret the subtle but significant community structure dynamics at N2S4 to reflect how injection of labile organic matter into the upper mesopelagic zone by physical mixing supports continued growth of euphotic zone-associated bacterioplankton lineages on a timescale of days.

Dissolved organic carbon (DOC) is produced as a result of numerous food web processes (Carlson and Hansell, 2015). However, the magnitude of annual DOC production is ultimately constrained by autochthonous primary production rates. Global net primary production (NPP) produces in excess of approximately 50 Pg C y–1 (Dunne et al., 2007; Westberry et al., 2008; Silsbe et al., 2016) and is the main source of DOC production (approximately 21 Pg C y–1) in the oceans (Carlson and Hansell, 2015). The oceanic DOC pool stores approximately 662 Pg C, which is comparable to the inventory of CO2 in the atmosphere and the amount of carbon in terrestrial biomass (Hansell et al., 2009). Heterotrophic bacterioplankton rapidly recycle DOC for energy generation and biomass synthesis (Azam et al., 1983), and those living in the ocean’s surface waters are responsible for remineralizing a majority (50%–90%) of newly produced DOC (Williams, 1984; Azam et al., 1993; Ducklow, 1999; Carlson and Hansell, 2015). DOC that resists rapid microbial degradation can accumulate in surface waters, generating vertical gradients in DOC (Carlson et al., 1994; Carlson and Hansell, 2015).

Why and how DOC accumulates is not well understood, but a number of hypotheses have been proposed (see reviews in Benner and Amon, 2015; Carlson and Hansell, 2015; Dittmar, 2015). Some DOC compounds accumulate because they are intrinsically resistant to or slowly degraded by heterotrophic utilization (Hansell, 2013; Shen and Benner, 2020). Phytoplankton have been observed to directly produce recalcitrant compounds (Aluwihare et al., 1997) or precursors to recalcitrant compounds (Arakawa et al., 2017). Additionally, labile DOC can be altered to recalcitrant compounds by phototransformation (Kieber et al., 1997; Benner and Biddanda, 1998) or by heterotrophic microbes, including archaea and bacteria (Ogawa et al., 2001; Gruber et al., 2006; Jiao et al., 2010; Guerrero-Feijóo et al., 2017; Goto et al., 2020). Alternatively, the molecular diversity of organic compounds that comprise the bulk DOC pool can limit unique compounds from existing at detectable concentrations or from encountering microbes with the appropriate uptake mechanisms. Thus, the accumulated DOC pool reflects the sum of these highly diverse but individually low-concentration compounds (Kattner et al., 2011; Arrieta et al., 2015; Dittmar, 2015). DOC may also accumulate because metabolic costs may outweigh the benefits of oxidizing particular compounds (Carlson et al., 2009; Treusch et al., 2009; Giovannoni, 2017; Landry et al., 2017; Saw et al., 2020). Possibly, all of these mechanisms contribute simultaneously to the accumulation of DOC compounds in surface waters, where their functional recalcitrance establishes a mismatch between their supply and consumption by microbes (Zakem et al., 2020). However, accumulated DOC that is resistant to rapid microbial utilization at one geographical location or depth can be readily utilized by heterotrophic bacterioplankton at another (Carlson et al., 2011).

In regions like the western North Atlantic where convective mixing extends from the euphotic zone into the mesopelagic zone, particulate organic carbon and accumulated DOC in surface waters can be delivered to the deep ocean physically by the deepening of the mixed layer (e.g., mixed layer pump; Copin-Montégut and Avril, 1993; Carlson et al., 1994; Gardner et al., 1995; Dall’Olmo et al., 2016; Lacour et al., 2019). Deep convection obliterates vertical density gradients and redistributes suspended and dissolved organic matter homogeneously throughout the mixed layer. DOC concentrations become reduced in the epipelagic and enriched in the mesopelagic as a portion of the accumulated euphotic-zone DOC is mixed to depth. As the water column stratifies and the mixed layer shoals to the surface, the exported constituents (DOC and suspended particulates, including microbial biomass and communities) become isolated in the mesopelagic zone below the mixed layer (Carlson et al., 1994; Hansell and Carlson, 2001). However, only approximately 11% of the surface-accumulated DOC exported to the deep ocean persists to depths greater than 500 m (Carlson et al., 2010; Hansell et al., 2012), due to remineralization within the upper mesopelagic zone. Field and experimental studies have provided evidence for the degradation of surface-accumulated DOC by mesopelagic heterotrophic microbes following deep convection (Copin-Montégut and Avril, 1993; Carlson et al., 1994; Ducklow et al., 1995; Carlson et al., 2004; Carlson et al., 2010). Within a warm-core eddy in the Gulf Stream, bacterioplankton abundance was observed to increase throughout the surface 150 m after the water column had been mixed to 450 m and later restratified. The sub-euphotic increase in bacterial abundance was presumed to be a response to organic matter delivered to depth (Ducklow, 1986). At the Bermuda Atlantic Time-series Study (BATS) site (31.67°N, 64.17°W), a portion of the DOC exported out of the euphotic zone following winter deep convection was observed repeatedly to be removed in the upper mesopelagic on a timescale of weeks (Hansell and Carlson, 2001; Goldberg et al., 2009; Liu et al., 2022). At the same time, absolute bacterioplankton abundance in the upper mesopelagic at BATS was observed to increase following deep convection (Carlson et al., 2009), with increases in relative abundance by specific bacterioplankton taxa like SAR11 subclade II, OCS116, SAR202, and marine Actinobacteria (Morris et al., 2005; Carlson et al., 2009; Treusch et al., 2009; Vergin et al., 2013; Liu et al., 2020a). These observations suggest that after surface accumulated DOC was exported to the upper mesopelagic zone, some mesopelagic bacterioplankton were capable of responding to the exported DOC on a timescale of weeks to months (Morris et al., 2005; DeLong et al., 2006; Landry et al., 2017).

Metagenomic evidence supports the claim that some mesopelagic microbial lineages are well adapted to utilize compounds that are apparently recalcitrant to epipelagic populations (Saw et al., 2020). Indeed, mixed culture experiments conducted at BATS demonstrated multi-week increases in the relative abundance of slow-growing mesopelagic bacterioplankton (e.g., SAR202, Methylophaga, Hyphomonadaceae, and Alcanivoracaceae) after the introduction of recalcitrant carboxyl-rich alicyclic (CRAM) proxy compounds (Liu et al., 2020b). While these experiments advance understanding about what could happen in the days and weeks following deep convection, the measurements of in situ microbial responses in the hours to days following deep convection could provide insights regarding the fate of surface-accumulated DOC after it is vertically exported from the epipelagic and potentially remineralized in the mesopelagic zone.

During the second of four field cruises of the North Atlantic Aerosols and Marine Ecosystems Study (NAAMES) in May 2016, we had the unique opportunity to sample a hydrostation in the subtropical region of the western North Atlantic (47.46°N, 38.72°W) during a deep convective mixing event and follow the subsequent water column restratification over a 3-day period (Behrenfeld, 2010; Della Penna and Gaube, 2019). At this hydrostation (NAAMES 2 Station 4, hereafter N2S4), we sought to describe changes in biogeochemistry and in bacterioplankton production (BP) and community structure throughout the water column, with a particular focus on the upper mesopelagic zone (100–200 m). We hypothesized that the surface-accumulated DOC delivered by deep convective mixing into the upper mesopelagic zone would stimulate in situ heterotrophic production by specific mesopelagic-associated lineages, like SAR11 subclade II and SAR202. This stimulation would then lead to the differentiation of bacterioplankton community structure between the euphotic and mesopelagic zones. Here, we describe the observations made at N2S4 and compare them to those made in stratified water columns that were sampled during NAAMES 3 (September 2017). Three of these stratified stations (NAAMES 3 Stations 3, 3.5, and 4) were in the subtropical region of the western North Atlantic, within 1.5˚ latitude and longitude of N2S4 (Della Penna and Gaube, 2019). A fourth station (NAAMES 3 Station 6) was in the subpolar region but was included in this analysis as a representative of a well-stratified station that was also occupied for multiple days (Della Penna and Gaube, 2019).

2.1. Sampling stations

Of the four cruises of the NAAMES study, we focus on the two where specific hydrostations were occupied for more than two days. The data from the two NAAMES cruises presented here are from NAAMES 2 (May 2016) and NAAMES 3 (September 2017). Each involved a ship transect aboard the R/V Atlantis between 39°N and 56°N latitude and 38°W and 47°W longitude (Figure 1).

Figure 1.

Ship, drifter, and float tracks for NAAMES 2 and 3 stations. Points along the ship and float tracks (lines) denote where water column data were taken. Panel (a) shows the tracks for NAAMES 2 Station 4 (N2S4) superimposed on a composite map of chlorophyll a derived from data collected daily between May 23 and May 27, 2016, by the Visible Infrared Imaging Radiometer Suite sensor aboard the Suomi-NPP satellite. Dashed black contours indicate the isolines of sea-level anomaly (referring to May 25, 2016) showing that our data collection took place between the core and periphery of a mesoscale eddy. Panel (b) shows the tracks for NAAMES 3 Station 6 (top, subpolar region) and NAAMES 3 Stations 3, 3.5, and 4 (bottom, subtropical region). The approximate location for N2S4 (47.46°N, 38.72°W) is also shown on the bottom of panel (b) as reference (star). NAAMES = North Atlantic Aerosols and Marine Ecosystems Study.

Figure 1.

Ship, drifter, and float tracks for NAAMES 2 and 3 stations. Points along the ship and float tracks (lines) denote where water column data were taken. Panel (a) shows the tracks for NAAMES 2 Station 4 (N2S4) superimposed on a composite map of chlorophyll a derived from data collected daily between May 23 and May 27, 2016, by the Visible Infrared Imaging Radiometer Suite sensor aboard the Suomi-NPP satellite. Dashed black contours indicate the isolines of sea-level anomaly (referring to May 25, 2016) showing that our data collection took place between the core and periphery of a mesoscale eddy. Panel (b) shows the tracks for NAAMES 3 Station 6 (top, subpolar region) and NAAMES 3 Stations 3, 3.5, and 4 (bottom, subtropical region). The approximate location for N2S4 (47.46°N, 38.72°W) is also shown on the bottom of panel (b) as reference (star). NAAMES = North Atlantic Aerosols and Marine Ecosystems Study.

Close modal

On May 24, 2016 (NAAMES 2), the R/V Atlantis occupied a water mass (hydrostation N2S4, 47.46°N, 38.72°W) within the core of a moderately retentive anticyclonic eddy in the subtropical region of the western North Atlantic (Della Penna and Gaube, 2019). Upon occupation, Surface-Velocity-Program (SVP) drifters and a Seabird Navis BGCi float (float ID number n0647) were deployed. The drifters consisted of a spherical surface float tethered to a weighted holey sock drogue extending to 15 m. For this specific analysis, we selected only the drifters that transmitted for more than 24 hours after deployment, resulting in the trajectory observations of one drifter for N2S4 (platform ID number 300234062417670). Conductivity-temperature-depth (CTD) profiles from both the R/V Atlantis and the float revealed deep convective mixing at the time of occupation with a mixed layer of approximately 230 m. The sampling stations of the R/V Atlantis followed the drifter track during the period May 24–26 and then the float track during May 26–27. The drifter and float deployed at N2S4 were retained within the anticyclonic eddy core for 3 days before being transported to the eddy edge by surface currents. The float continued to profile the eddy for a total of 6 days (until May 30; Della Penna and Gaube, 2019; Figure 1a). To provide visually spatial context for our observations, we downloaded the maps of satellite altimetry (for May 25, 2016) from the Copernicus Marine Service Information portal, as well as chlorophyll a (Chl a; for May 23–27, 2016) from the Visible Infrared Imaging Radiometer Suite sensor aboard the Suomi-NPP satellite. A full analysis of the satellite altimetry data set for NAAMES can be found in Della Penna and Gaube (2019). Sea-level anomaly maps are available daily and are gridded on 1/4°.

In September 2017 (NAAMES 3), the R/V Atlantis sampled 3 subtropical stations within 1.5° latitude and longitude of the coordinates of N2S4, including N3S3 at 47.03°N, 40.11°W; N3S3.5 at 48.04°N, 39.24°W; and N3S4 at 48.64°N, 39.13°W (Figure 1b). N3S3 was located within the core of an anticyclonic mode-water eddy, and N3S4 was located at the periphery of a cyclonic eddy (Della Penna and Gaube, 2019). Each 24-hour occupation included a CTD cast to 200 m and a Seabird Navis BGCi float deployment (ID number n0850 at N3S3 and ID number n0849 at N3S4). Station 3.5, not associated with an eddy, was occupied for approximately 2 hours and did not include a float deployment.

On September 13, 2017, during the third NAAMES cruise, the R/V Atlantis arrived on a subpolar hydrostation (N3S6, at 53.38°N, 39.54°W) located in a quiescent subarctic region of the northwest Atlantic, which was characterized by large-scale wind-driven circulation and relatively more uniform currents (Della Penna and Gaube, 2019). This hydrostation was not associated with an eddy and was tracked and sampled via SVP drifters (platform ID numbers 300234065316880 and 300234065316710), Seabird Navis BGCi floats (float ID numbers n0846 and n0847) and ship. The drifters and floats deployed at this station were transported eastward by wind-driven circulation (Della Penna and Gaube, 2019; Figure 1b). The water column at N3S6 remained strongly stratified over a 4-day occupation (mixed layer depth ≤ 31 m).

We recognize that advective features can impact our observations. Because the focus of this study is on the biogeochemical and microbial dynamics in the upper mesopelagic zone (100–200 m) of N2S4, we limit the impact of advective features to our observations at N2S4 by including in our analyses only float and CTD profiles (and associated discrete water samples) with the same temperature, salinity, and density properties at and below 50 m (Figure S1). Anomalous profiles based on temperature-salinity diagrams were also omitted from further analysis for the NAAMES 3 stations (Figures S2 and S3).

2.2. Hydrographic data and discrete sample collection

Hydrographic ship data were collected up to 5 times each day of station occupation from Sea-Bird Scientific SBE-911+ CTD profiler sensors attached to a rosette with 24 × 10 L Niskin bottles while on station. Additional sensors included those for chlorophyll fluorescence (WET Labs ECO-AFL/FL), beam transmission (WET Labs C-Star), turbidity (WET Labs ECO), and oxygen (SBE43). These data are available through National Aeronautics and Space Administration’s Ocean Biology Distributed Active Archive Center (OB.DAAC) SeaWiFS Bio-optical Archive and Storage System (NAAMES, 2018). Data were averaged over 1-m bins for the present analysis.

Data from Seabird Navis BGCi floats were retrieved and processed by the University of Maine’s In-situ Sound & Color (MISC) Lab. Float profiles were collected up to 7 times each day of station occupation. The L2 processed data used in the present analysis are available via the University of Maine In-situ Sound & Color Lab FTP (NAAMES Floats, 2018) as well as the Argo Global Data Assembly Center (Argo, 2022). Float data were averaged over 1-m bins for the present analysis.

Following previous studies of the NAAMES stations, mixed layer depths for each CTD and float profile were determined as the depth below 5 m at which the Brunt–Väisālä buoyancy frequency (N2) was greater than its standard deviation (Graff and Behrenfeld, 2018; Morison et al., 2019). The maximum depth of the euphotic zone was determined from float data as the depth where irradiance was 1% of surface photosynthetically active radiation (PAR). The maximum euphotic zone depth for N2S4 and NAAMES 3 was estimated to be 77 and 57 m, respectively. Apparent oxygen utilization (AOU, µmol O2 L–1) is defined as the depletion of oxygen relative to saturation and was calculated for each 1-m binned CTD profile as the difference between the oxygen saturation concentration and the observed oxygen concentration. Oxygen saturation was estimated from potential temperature and salinity following Weiss (1970).

The Niskin bottles collected seawater samples at nominal depths of 5, 10, 25, 50, 75, 100, 150, and 200 m for the analysis of Chl a, phytoplankton cell abundance, nitrate + nitrite (N + N), bacterioplankton cell abundance (BA), bacterioplankton leucine incorporation rate, DOC, total dissolved amino acids (TDAA), and bacterioplankton community composition via 16 S rRNA amplicon sequencing. Replicate samples for DOC and bacterioplankton leucine incorporation were collected at all stations, while all other variables are represented by single sample analyses. These data are available through OB.DAAC SeaBASS (NAAMES, 2018), as well as the Biological and Chemical Oceanography Data Management Office (Carlson, 2020). All processed data, analyses, and code presented here are available on GitHub (Baetge, 2022).

2.2.1. Chlorophyll a, phytoplankton cell abundance, and net primary production

Chl a concentrations (µg L–1) were measured from single discrete samples by fluorometric analysis using the acidification technique and a Turner Designs 10AU digital fluorometer (Mueller et al., 2003). For each sample, 500 mL of raw seawater were filtered through a 25-mm 0.45-µm (Millipore HA) nitrocellulose filter, which was then extracted in 90% acetone for 48 hours at 0°C. Fluorescence was measured before and after acidification to determine Chl a concentration. Phytoplankton abundances (cells L–1) of single samples for cells of < 64 µm diameter were determined within hours of sample collection using a BD Influx Sorting Flow Cytometer and categorized into four major groups: Prochlorococcus, Synechococcus, picoeukaryotes, and nanoeukaryotes (see methods in Graff and Behrenfeld, 2018). Total phytoplankton cell abundance was calculated as the sum of the abundances of all the groups. NPP (µmol C L–1 d–1) was determined from 14C-sodium bicarbonate uptake incubations. Specifically, water was collected from 3 depths predawn, inoculated with 14C sodium bicarbonate, and incubated in on-deck incubators at light levels corresponding to the collection depths. Water collected at the surface was separated and incubated at 6 light levels (approximately 1%, 10%, 20%, 40%, 60%, and 100% transmission of surface light). NPP was estimated at 1-m bins through the water column to 200 m using the Photoacclimation Productivity Model (PPM) from Fox et al. (2020), which accounts for physiological adjustments in intracellular Chl a concentration in response to light and nutrient availability (Behrenfeld et al., 2016). There was strong agreement between NPP estimates from the PPM and available measurements of 24-h 14C-uptake conducted at all NAAMES stations (Fox et al., 2020).

2.2.2. BA and 3H-leucine incorporation

BA (cells L–1) was determined for unfiltered seawater samples preserved with certified American Chemical Society formalin to a final concentration of 1% (vol: vol). Within 36 hours of collection, cells from each sample were filtered onto a blackened 0.2-µm polycarbonate membrane filter, stained with DAPI (4′,6-diamidino-2-phenylindole dihydrochloride, 5 µg mL–1), mounted onto a slide with high viscosity immersion oil (Thermo Scientific Richard-Allan Scientific Resolve) and stored at –20°C until enumeration. DAPI-stained cells were enumerated manually using an Olympus BX51epifluorescence microscope with ultraviolet excitation at 1,000× magnification. Twelve fields of view were counted for each slide, and on average, 50–60 cells were counted for each field-of-view (Porter and Feig, 1980).

Bacterioplankton 3H-leucine incorporation rates (pmol 3H-Leu L–1 h–1) were estimated using a modified version of the microcentrifuge method (Smith and Azam, 1992). For each depth, a killed control (fixed immediately with 100 µL of 100% trichloroacetic acid [TCA]) and replicate 1.6-mL seawater samples were spiked with 3H-Leu (L-[4,5- 3H(N)] at 20 nM; specific activity 50.2–52.6 Ci mmol–1; Perkin Elmer, Boston, MA). After incubating in the dark for 2–3 hours at ± 2°C of in situ temperature (15.3°C–25.7°C at N2S4, 15.5°C–19.3°C at NAAMES 3 Station 3, 10.4°C–16.0°C at NAAMES 3 Station 3.5, 7.2°C–16.6°C at NAAMES 3 Station 4, and 3.9°C–11.6°C at NAAMES 3 Station 6), incubations were terminated with 100-µL cold 100% TCA. Cells were pelletized after a series of microcentrifugation steps, washed with 5% TCA, 80% ethanol (vol/vol), and resuspended in Ultima Gold scintillation cocktail as described in Ducklow et al. (2001). Radioactivity was measured using a Hidex 300 Scintillation Analyzer and was corrected for quenching using an external gamma source and a quench curve. The coefficient of variation (CV) between measurements from replicate incubations was generally 1%–15%. However, CVs for the upper mesopelagic samples increased up to 40% because of the lower incorporation rates at those depths. The range in CVs for the upper mesopelagic zone was 1%–25% for N2S4, 10%–20% for NAAMES 3 Station 3, 1%–3% for NAAMES 3 Station 4, and 12%–40% for NAAMES 3 Station 6.

2.3. Dissolved inorganic nutrients, dissolved organic carbon, and total dissolved amino acids

N + N concentrations (µmol N L–1) at each depth were determined from single discrete samples that were gravity-filtered directly from the Niskin bottles through 47-mm 0.8-µm polycarbonate (Millipore) filters and stored at –20°C in sterile 50-mL conical centrifuge tubes. Samples were analyzed using a Lachat QuickChem QC8500 automated ion analyzer (University of Rhode Island Graduate School of Oceanography Marine Science Research Facility), which measures N + N with a precision of approximately 0.3 µmol N L–1 and has a detection limit of approximately 0.8 µmol N L–1.

Replicate DOC samples at each depth were gravity-filtered directly from the Niskin bottles through pre-combusted 47-mm 0.7-µm GF/F filters (Whatman, first flushed with 100 mL of sample water) into pre-combusted 40-mL certified Environmental Protection Agency (EPA) borosilicate glass vials. Samples were acidified to a pH of ≤ 3 by adding 50-µL DOC-free 4 N HCl and stored at 14°C in a volatile organic-free environmental chamber until analysis. DOC concentrations (µmol C L–1) were measured in batches of ≤ 35 samples using the high-temperature combustion method (Carlson et al., 2010; Halewood et al., 2022) on Shimadzu TOC-V or TOC-L analyzers. Concentrations were quantified using standard solutions of glucose and low carbon ultrapure water. Samples were systematically referenced against surface and deep seawater calibrated with consensus reference material (Hansell SSR Lot#08-18), which were run every 6–8 samples. Precision of the Shimadzu analyzers was within 2% CV.

TDAA were determined from single samples that were gravity-filtered directly from the Niskin bottles through 47-mm 0.7-µm GF/F (Whatman, first flushed with 100 mL of sample water) into acid-washed 60-mL high-density polyethylene bottles and stored at –20°C. Samples were processed following Liu et al. (2020a), where samples were hydrolyzed in 6 N HCl under nitrogen for 20 hours at 110°C, neutralized using nitrogen evaporation, derivatized with ortho-phthalaldehyde, and measured using a Dionex ICS5000+ high performance liquid chromatography (HPLC) equipped with a fluorescence detector (excitation = 330 nm and emission = 418 nm). The molecular formula of each amino acid resolved by the HPLC was used to calculate its carbon concentration in µmol C L–1. Thus, TDAA represents the sum of the carbon concentrations of all the individual amino acids. DOC-normalized yields of TDAA were calculated as the percentage of total DOC measured as amino acid carbon.

2.4. Derived variables for N2S4

Net BP (nmol C L–1 d–1) was estimated for each depth by applying a leucine conversion factor of 1.5 kg C mol Leu–1 incorporated (Simon and Azam, 1989). Phytoplankton cell abundance, NPP, AOU, BA, BP, DOC, and TDAA profiles were all integrated for the euphotic and the upper mesopelagic zones using the trapezoidal rule. The euphotic zone includes samples and data collected above the maximum depth at which PAR was equivalent to 1% surface PAR, which for N2S4 was 77 m and for NAAMES 3 stations 3, 3.5, 4, and 6 was 57 m. The 1% light level ranged between 27 and 77 m (n = 28) at N2S4. The mesopelagic zone includes samples and data collected between the depth where PAR was 1% of its surface value and 200 m. The integrated stocks and rates for both depth horizons were normalized to the integrated depth interval to provide mean volumetric concentrations and rates for the respective depth zones. The upper bound for the error (E) incurred by a trapezoidal integration of a depth profile was estimated by the error bound formula:

where a is the minimum depth, b is the maximum depth, n is the number of trapezoids used in the integration approximation, and k is the upper bound of the absolute value of the second derivative of a polynomial function fitted to the depth profile.

2.5. Bacterioplankton community composition

Nucleic acid sample collection, extraction, purification and 16 S rRNA gene amplification, library preparation, sequencing, and amplicon assignment are detailed in Bolaños et al. (2021). A single 4-L sample from each depth was collected by filtration through a 0.22-µm Sterivex filter (polyethersulfone membrane, Millipore) and preserved in sucrose lysis buffer (SLB, 20-mM ethylenediaminetetraacetic acid [EDTA], 400-mM NaCl, 0.75-M sucrose, and 50-mM Tris-HCl, pH 9.0) at –80°C. Nucleic acids were extracted and purified following the phenol: isoamyl alcohol: chloroform (25:1:24) protocol of Giovannoni et al. (1996). The V1–V2 region of the 16 S rRNA gene was amplified via polymerase chain reaction (PCR) using the 27F (5′-AGAGTTTGATCNTGGCTCAG-3′) forward and 338 RPL (5′-GCWGCCWCCCGTAGGWGT-3′) reverse primers, each with respective Illumina overhang adapters (Bolaños et al., 2021). The 25-µL PCR reactions consisted of 2.5 µL (5 ng μL–1) of genomic DNA template, 2.5 µL of each primer (1 µM), 12 µL of 2× KAPA HiFi HotStart ReadyMix, and 5.5-µL PCR water. Thermocycling conditions of the PCR reactions were 3 minutes at 95°C, 25 cycles of 30 seconds at 95°C, 30 seconds at 55°C, 30 seconds at 72°C, and 5 minutes at 75°C. Libraries for each amplicon reaction product were constructed by attaching dual indices and Illumina sequencing adapters with the Nextera XT Index Kit using a second PCR amplification. Purified libraries were pooled in equimolar concentrations and sequenced using the Illumina MiSeq platform (v2, 2x250 PE lane) at the Center for Genome Research and Biocomputing (Oregon State University, Corvallis, OR, USA). Amplicon sequence data sets can be found in the National Center for Biotechnology Information Sequence Read Archive (NCBI SRA) database under the BioProject identifier PRJNA627189.

Primer sequences were removed from demultiplexed FASTQ files using the CutAdapt algorithm (Martin, 2011). All subsequent sequence analyses were performed using the R software environment (v 4.0.0). Using the package dada2 (v 1.2.0; Callahan et al., 2016), trimmed FASTQ files were quality filtered, dereplicated, and merged to create an amplicon sequence variant (ASV) table, from which potential chimeras were removed de novo. ASVs were taxonomically assigned using the SILVA database (v 123, Quast et al., 2013) and PhyloAssigner (v089, Vergin et al., 2013). SAR116 was reclassified from order Rickettsiales to order Puniceispirillales according to the latest version of SILVA (v. 138.1). Phylogenetic databases are available on GitHub (Bolaños, 2020). Using the function rarefy_even_depth in the package phyloseq (v 1.32.0; McMurdie and Holmes, 2013), sequence read counts were subsampled to the minimum sample read depth (9,702 reads) with replacement to standardize for sampling effort. The Chao1 and Shannon alpha diversity indices were estimated using the phyloseq function estimate_richness. The Shannon evenness index was calculated by dividing the Shannon diversity index by the total number of ASVs observed (Clarke, 1993). Pairwise comparisons of these alpha diversity indices were performed with t tests using the function compare_means of the package ggpubr (v 0.3.0). A one-way permutational multivariate analysis of variance (PERMANOVA) test was used to assess whether there were statistically significant differences between groups (e.g., stations, depth horizons, or time) using the function adonis in the R package vegan (v 2.5-6; Anderson, 2001). Bray–Curtis dissimilarities were calculated using the vegdist function of the package vegan (v 2.5-6; Oksanen et al., 2013). Nonmetric multidimensional scaling (NMDS) ordinations with Bray–Curtis dissimilarities were computed with the phyloseq function ordinate. Similarity percentages breakdown (SIMPER) using the simper function of the package vegan was used to identify major ASVs whose differences over time were an important contribution to the Bray–Curtis dissimilarities in the upper mesopelagic of N2S4. For all statistical analyses, p values of > 0.05, ≤ 0.05, and ≤ 0.01 are described as not significant, significantly different, or highly significantly different, respectively.

For each of the euphotic and upper mesopelagic zones, the relative abundance of a taxonomic group to the family level was calculated by summing its ASV counts across all depths in the depth horizon and then dividing the summed counts by the total counts in the depth horizon. An approximation of cell abundance for each taxonomic group was estimated using the product between the relative abundance of its sequences and the total cell abundance, normalized to its gene copy number, hereafter termed “16 S rRNA-estimated cell abundance.” Bacterial ribosomal RNA copy number information was retrieved from the ribosomal RNA operon copy number database (rrnDB; Stoddard et al., 2015). If copy number information was not available for a phylogenetic order, then the copy number for the next higher taxonomic rank was used.

3.1. NAAMES 2 Station 4 (N2S4)

The mixed layer depth in the anticyclonic eddy of N2S4 was 227 m upon occupation on May 24, 2016, and shoaled during station occupation over the next 4 days to < 26 m (average 11 ± 8 m, n = 9), with sustained thermal stratification through May 29, 2016 (observations after May 27 were from float n0647; Figure 2). On the first day of occupation of N2S4, the biological and chemical parameters N + N, Chl a, phytoplankton cell abundance, BP, AOU, and DOC were distributed homogeneously throughout the surface to 200 m, reflecting the deep mixed layer (Figure 3a–f). BA and TDAA were not distributed homogeneously through depth on the first day of occupation of N2S4 when the mixed layer was deepest (Figure 3h and i). However, BA increased at most depths over the station occupation, while TDAA displayed an increasing and decreasing trend in the euphotic and upper mesopelagic, respectively (Figure 3). Many of the measured variables displayed different trends between the euphotic and mesopelagic zones that can be more clearly discerned from depth-integrated changes, as detailed below (Figure 4).

Figure 2.

Composite time series from ship conductivity-temperature-depth (CTD) and float data at North Atlantic Aerosols and Marine Ecosystems Study (NAAMES) 2 Station 4. Vertical black lines denote when water column data were collected by the ship CTD (dotted) or by float n0647 (solid). White lines show the mixed layer depth; red lines delineate the bottom of the euphotic zone (1% of surface photosynthetically active radiation). In the background are interpolated data of (a) temperature with isobars binned to 0.1°C, (b) salinity with isohalines binned to 0.05, and (c) chlorophyll a fluorescence with isolines binned to 0.5 mg m–3.

Figure 2.

Composite time series from ship conductivity-temperature-depth (CTD) and float data at North Atlantic Aerosols and Marine Ecosystems Study (NAAMES) 2 Station 4. Vertical black lines denote when water column data were collected by the ship CTD (dotted) or by float n0647 (solid). White lines show the mixed layer depth; red lines delineate the bottom of the euphotic zone (1% of surface photosynthetically active radiation). In the background are interpolated data of (a) temperature with isobars binned to 0.1°C, (b) salinity with isohalines binned to 0.05, and (c) chlorophyll a fluorescence with isolines binned to 0.5 mg m–3.

Close modal
Figure 3.

Depth profiles of biological and biogeochemical properties during occupation of North Atlantic Aerosols and Marine Ecosystems Study 2 Station 4. Profiles over 0–200 m for (a) concentration of nitrate + nitrite from single water samples, (b) concentration of chlorophyll a from single water samples, (c) phytoplankton cell concentration from single water samples, enumerated by cell-sorting flow cytometry, (d) net bacterioplankton production from duplicate incubations, where error bars represent standard deviation of the mean rate between incubations, (e) apparent oxygen utilization derived from oxygen measurements from the conductivity-temperature-depth, (f) dissolved organic carbon from duplicate water samples, where error bars represent standard deviation of the mean concentration of the two water samples, (g) net primary production estimated using the photoacclimation productivity model, (h) bacterioplankton cell abundance from single water samples, enumerated by epifluorescence microscopy, where error bars represent standard deviation of the mean count of 10 fields of view, and (i) total dissolved amino acids from single water samples, where error bars represent standard deviation of the mean concentration from analytical replicates. Horizontal red lines indicate the maximum depth at which irradiance was 1% of surface photosynthetically active radiation (77 m).

Figure 3.

Depth profiles of biological and biogeochemical properties during occupation of North Atlantic Aerosols and Marine Ecosystems Study 2 Station 4. Profiles over 0–200 m for (a) concentration of nitrate + nitrite from single water samples, (b) concentration of chlorophyll a from single water samples, (c) phytoplankton cell concentration from single water samples, enumerated by cell-sorting flow cytometry, (d) net bacterioplankton production from duplicate incubations, where error bars represent standard deviation of the mean rate between incubations, (e) apparent oxygen utilization derived from oxygen measurements from the conductivity-temperature-depth, (f) dissolved organic carbon from duplicate water samples, where error bars represent standard deviation of the mean concentration of the two water samples, (g) net primary production estimated using the photoacclimation productivity model, (h) bacterioplankton cell abundance from single water samples, enumerated by epifluorescence microscopy, where error bars represent standard deviation of the mean count of 10 fields of view, and (i) total dissolved amino acids from single water samples, where error bars represent standard deviation of the mean concentration from analytical replicates. Horizontal red lines indicate the maximum depth at which irradiance was 1% of surface photosynthetically active radiation (77 m).

Close modal
Figure 4.

Depth-normalized integrations of biological and biogeochemical properties of North Atlantic Aerosols and Marine Ecosystems Study 2 Station 4. Stocks and rates for the euphotic (0–75 m) and the upper mesopelagic (100–200 m) zones: (a) chlorophyll a concentration, (b) phytoplankton cell concentration, (c) nitrate + nitrite, (d) apparent oxygen utilization, (e) bacterioplankton cell abundance, (f) net bacterioplankton production, (g) dissolved organic carbon, and (h) total dissolved amino acids. The volumetric values presented represent the mean value for the euphotic and upper mesopelagic zones, respectively, determined by integrating values within each depth zone and normalizing by integration depth. Error bars represent the upper bound of the error incurred by the trapezoidal rule approximation.

Figure 4.

Depth-normalized integrations of biological and biogeochemical properties of North Atlantic Aerosols and Marine Ecosystems Study 2 Station 4. Stocks and rates for the euphotic (0–75 m) and the upper mesopelagic (100–200 m) zones: (a) chlorophyll a concentration, (b) phytoplankton cell concentration, (c) nitrate + nitrite, (d) apparent oxygen utilization, (e) bacterioplankton cell abundance, (f) net bacterioplankton production, (g) dissolved organic carbon, and (h) total dissolved amino acids. The volumetric values presented represent the mean value for the euphotic and upper mesopelagic zones, respectively, determined by integrating values within each depth zone and normalizing by integration depth. Error bars represent the upper bound of the error incurred by the trapezoidal rule approximation.

Close modal

In the euphotic zone (0–75 m samples), increases in depth-integrated Chl a and phytoplankton cell abundance corresponded to decreases in N + N and AOU (Figure 4a–d). Between May 24 and May 27, euphotic zone Chl a increased from 0.50 ± 0.03 µg L–1 to 0.85 ± 0.34 µg L–1 (t test, p = 0.05), phytoplankton cell abundance increased from 1.5 × 107 ± 3.3 × 104 to 4.7 × 107 ± 3.5 × 107 cells L–1 (t test, p = 0.1; Figure 4a and b). N + N in the euphotic zone showed a decreasing trend (t test, p = 0.03; Figure 4c). Euphotic zone AOU decreased from May 24 to May 27 from 16.6 ± 0.3 to 11.3 ± 6.2 µmol O2 L–1 (t test, p < 0.0001; Figure 4d). Both BA and BP increased in the euphotic zone over time (Figure 4e and f). From May 24 to 26, BA increased from 1.2 × 109 ± 8.9 × 107 cells L–1 to 1.7 × 109 ± 4.1 × 108 cells L–1 (t test, p = 0.1; Figure 4e), while BP increased from 21.2 ± 2.5 nmol C L–1 d–1 to 52.5 ± 2.5 nmol C L–1 d–1 (t test, p < 0.0001; Figure 4f). As a result of the competing photoautotrophic and heterotrophic processes, we were not able to resolve significant changes in the bulk DOC concentrations in the euphotic zone during our station occupation (t test, p = 0.2). Bulk DOC concentrations within the euphotic zone averaged 53.4 ± 0.2 µmol C L–1. However, the contribution of TDAA to the bulk DOC pool increased from 1% to 1.4% in the euphotic zone (t test, p = 0.2; Figure 4g and h).

In the upper mesopelagic zone (100–200 m), Chl a decreased from 0.44 ± 0.05 µg L–1 to 0.27 ± 0.05 µg L–1 (t test, p = 0.02), and phytoplankton cell abundance declined from 1.5 × 107 ± 4.1 × 105 cells L–1 to 9.9 × 106 cells L–1 ± 7.1 × 105 (t test, p < 0.01; Figure 4a and b). There were small increases in upper mesopelagic AOU (1 µmol O2 L–1, t test, p < 0.0001) and N + N (0.2 µmol N L–1, t test, p = 0.2) corresponding to increases in BA (7.4 × 108 ± 1.8 × 108 cells L–1 to 1.5 × 109 ± 1.0 × 108 cells L–1, t test, p = 0.007) and BP (20.0 ± 4.7 nmol C L–1 d–1 to 31.0 ± 3.0 nmol C L–1 d–1, t test, p = 0.02; Figure 4c–f). The average mesopelagic BP rate during the well-stratified late summer (i.e., NAAMES 3 stations) was 2.1 ± 1.6 nmol C L–1 d–1. We observed that the upper mesopelagic BP rates upon occupation of N2S4 were 10-fold greater than those of the NAAMES 3 stations, averaging 20.0 ± 4.7 nmol C L–1 d–1 (n = 3) and increasing to as much as 35.2 ± 5.6 nmol C L–1 d–1 (n = 3) over the 3 days of occupation of N2S4 (mean upper mesopelagic BP after the first day of occupation was 32.0 ± 3.6 nmol C L–1 d–1, n = 18). The increase in BA in the upper mesopelagic (by 7.6 × 108 cells L–1) accounted for 63% of the total cell abundance increase throughout the water column (1.2 × 109 cells L–1; Figure 5e). Additionally, the upper mesopelagic increase in BP (by 11 nmol C L–1 d–1) accounted for 35% of the total increase in BP throughout the upper 200 m of the water column (31 nmol C L–1 d–1; Figure 4f).

Figure 5.

Depth profiles of biological and biogeochemical properties during occupation of North Atlantic Aerosols and Marine Ecosystems Study 3 stations. Profiles over 0–200 m during occupation of subtropical stations N3S3, N3S3.5, and N3S4 and subpolar station N3S6 for (a) concentration of nitrate + nitrite from single water samples, (b) concentration of chlorophyll a from single water samples, (c) phytoplankton cell concentrations from single water samples, enumerated by cell-sorting flow cytometry, (d) net primary production estimated using the photoacclimation productivity model, (e) apparent oxygen utilization derived from oxygen measurements from the conductivity-temperature-depth, (f) bacterioplankton cell abundance from single water samples, enumerated by epifluorescence microscopy, (g) net bacterioplankton production from duplicate incubations, and (h) dissolved organic carbon from duplicate water samples. Horizontal red lines indicate the maximum depth at which irradiance was 1% of surface photosynthetically active radiation (57 m). Error bars and ribbons represent the standard deviation of the mean of samples taken from multiple casts.

Figure 5.

Depth profiles of biological and biogeochemical properties during occupation of North Atlantic Aerosols and Marine Ecosystems Study 3 stations. Profiles over 0–200 m during occupation of subtropical stations N3S3, N3S3.5, and N3S4 and subpolar station N3S6 for (a) concentration of nitrate + nitrite from single water samples, (b) concentration of chlorophyll a from single water samples, (c) phytoplankton cell concentrations from single water samples, enumerated by cell-sorting flow cytometry, (d) net primary production estimated using the photoacclimation productivity model, (e) apparent oxygen utilization derived from oxygen measurements from the conductivity-temperature-depth, (f) bacterioplankton cell abundance from single water samples, enumerated by epifluorescence microscopy, (g) net bacterioplankton production from duplicate incubations, and (h) dissolved organic carbon from duplicate water samples. Horizontal red lines indicate the maximum depth at which irradiance was 1% of surface photosynthetically active radiation (57 m). Error bars and ribbons represent the standard deviation of the mean of samples taken from multiple casts.

Close modal

Bulk DOC concentrations displayed a systematic decreasing trend within the upper mesopelagic zone over the occupation of station N2S4 (t test, p = 0.04; Figure 4g). The contribution of TDAA to the bulk DOC pool in the upper mesopelagic zone decreased from 1.3% to 0.9% (t test, p = 0.1; Figure 4h).

3.2. NAAMES 3 Stations 3, 3.5, 4, and 6

In contrast to N2S4, the water masses occupied during NAAMES 3 were thermally stratified. Hydrostations N3S3, N3S3.5, and N3S4 were in the subtropical region of the western North Atlantic and were all occupied for less than 24 hours with up to two CTD casts per station. Hydrostation N3S6 in the subpolar region was occupied for four days with up to 4 CTD casts per day. N3S6 remained thermally stratified throughout the occupation with little change in its mixed layer depth (mean of 24 ± 3 m, n = 31) and little variability in its biological and biogeochemical properties (Figures 5 and S4).

All NAAMES 3 stations presented here exhibited strong vertical gradients in N + N, Chl a, phytoplankton cell abundance, NPP, AOU, BA, BP, and DOC (Figure 5). Inorganic N + N increased over depth at all stations and displayed a latitudinal trend of increasing mesopelagic concentrations northward. Euphotic zone N + N concentrations were lower at all 4 of these stations relative to N2S4, averaging 1.9 ± 2.7 µmol N L–1 compared to 5.2 ± 0.1 µmol N L–1 at N2S4. Except for N3S3, the NAAMES 3 stations had higher N + N concentrations in the upper mesopelagic, averaging 12.4 ± 4.2 µmol N L–1 compared to 5.4 ± 0.1 µmol N L–1 for N2S4 (Figures 4a and 5a).

Chl a decreased over depth at N3S4 and N3S6 and a subsurface chlorophyll maximum was apparent at 50 m for N3S4. In the euphotic zone, Chl a averaged 0.5 ± 0.2 µg L–1 for these stations, which was lower than the average concentration at N2S4 (0.8 ± 0.4 µg L–1). Chl a was also lower in the upper mesopelagic of N3S4 and N3S6 (0.02 ± 0.02 µg L–1) relative to N2S4 (0.3 ± 0.1 µg L–1; Figures 4b and 5b).

At the NAAMES 3 stations, phytoplankton cell abundance decreased with depth, with higher euphotic zone concentrations at subtropical stations N3S3, N3S3.5, N3S4 than subpolar station N3S6. Phytoplankton cell abundance averaged 2.0 × 108 ± 1.3 × 108 cells L–1 in the euphotic zone, which was higher than average the cell abundance at N2S4 (2.7 × 107 ± 1.7 × 106 cells L–1). In the upper mesopelagic, phytoplankton cell abundance averaged 1.7 × 107 ± 5.7 × 106 cells L–1, comparable to the average cell abundance at N2S4 (1.2 × 107 ± 2.6 × 106 cells L–1; Figures 4c and 5c).

NPP decreased with depth at N3S3, N3S4, N3S6, with higher surface rates at subpolar station N3S6 relative to subtropical stations N3S3 and N3S4. The euphotic zone of the NAAMES 3 stations had a comparable average (0.9 ± 0.8 µmol C L–1 d–1) and lower range (0.1–2.7 µmol C L–1 d–1) of NPP than N2S4 (average 1.1 ± 1.0 µmol C L–1 d–1, range = 0–4.2 µmol C L–1 d–1; Figures 4d and 5d).

AOU was variable between all NAAMES 3 stations but decreased over depth at all stations, with an average in the euphotic zone of 14.5 ± 12.3 µmol O2 L–1 comparable to that at N2S4 (14.1 ± 3.4 µmol O2 L–1). Average and median AOU values for the upper mesopelagic zone at the NAAMES 3 stations were 47.0 ± 17.6 µmol O2 L–1 and 43.0 µmol O2 L–1, respectively. These values were substantially higher than the average AOU for the upper mesopelagic zone at N2S4 (17.8 ± 0.9 µmol O2 L–1; Figures 4e and 5e).

BA decreased with depth at all NAAMES 3 stations, without any clear latitudinal trends between them. Over the euphotic zone, BA was higher at the NAAMES 3 stations relative to N2S4, averaging 1.5 × 109 ± 5.0 × 108 cells L–1 compared to 1.3 × 109 ± 3.3 × 108. The NAAMES 3 stations, however, had lower BA (5.8 × 108 ± 3.6 × 108 cells L–1) in the upper mesopelagic zone relative to N2S4 (1.1 × 109 ± 3.6 × 108 cells L–1; Figures 4f and 5f).

At N3S3, N3S4, and N3S6, BP decreased to background values of 2.4 ± 1.9 nmol C L–1 d–1 by 75-m depth and did not exhibit latitudinal trends between the stations. BP was comparable on average in the euphotic zone at the NAAMES 3 stations (41 ± 9 nmol C L–1 d–1) relative to N2S4 (40 ± 10 nmol C L–1 d–1). In the upper mesopelagic zone, however, BP values were on average 10-fold higher on the first day of occupation of N2S4 (20 ± 4.7 nmol C L–1 d–1) than the average of those at the NAAMES 3 stations (2.1 ± 1.6 nmol C L–1 d–1; Figures 4g and 5g).

DOC concentrations decreased with depth at N3S3, N3S3.5, N3S4, and N3S6, with elevated surface concentrations at subtropical stations N3S3, N3S3.5, and N3S4 relative to subpolar station N3S6. In the euphotic zone, DOC concentrations averaged 63.4 ± 4.4 µmol C L–1 for the NAAMES 3 stations, which was greater than the average concentration of 53.4 ± 0.6 µmol C L–1 at N2S4. The average DOC concentration in the upper mesopelagic zone was slightly lower at the NAAMES 3 stations (51.4 ± 2.0 µmol C L–1) compared to N2S4 (53.1 ± 0.5 µmol C L–1; Figures 4h and 5h).

3.3. Bacterioplankton community composition

Bacterioplankton profiles were analyzed over the surface 200 m to assess community composition shifts during rapid restratification of the deeply mixed water column at N2S4 and to compare these dynamics with properties of a persistently stratified systems (i.e., N3S3, N3S3.5, N3S4, and N3S6). Bacterioplankton ASV evenness was not significantly different between the depth horizons at N2S4 (Shannon evenness, t test, p = 0.82), nor at the NAAMES 3 stations (Shannon evenness, t test, p = 0.16). Bacterioplankton ASV richness was also not significantly different between the depth horizons at N2S4 (Chao1, t test, p = 0.31) but was significantly greater in the upper mesopelagic zone (100–200 m) relative to the euphotic zone (Chao1, t test, p = 0.00013) at the NAAMES 3 stations. ASV richness at N2S4 was significantly higher than observed at the NAAMES 3 stations for both the euphotic (t test, p = 0.0006) and the upper mesopelagic zone (t test, p = 0.04). The average Chao1 index for the water column of 5–200 m at N2S4 was 679 ± 135 (n = 21) compared to 585 ± 102 (n = 11) for the upper mesopelagic of the NAAMES 3 stations. ASV evenness was significantly lower at N2S4 than at the NAAMES 3 stations for both the euphotic (t test, p < 0.0001) and the upper mesopelagic zones (t test, p = 0.04; Figure 6). NMDS ordination based on Bray–Curtis dissimilarities (stress = 0.05) showed that the N2S4 samples clustered tightly together and far from the NAAMES 3 samples (Figure 7a). A PERMANOVA test revealed that community structure differed significantly by cruise (r2 = 0.37, p < 0.01) and depth horizon (r2 = 0.04, p = 0.03), without a significant interaction effect (r2 = 0.02, p < 0.17). Within the samples collected at N2S4, the community over the entire water column clustered more closely over time (NMDS stress = 0.09). A PERMANOVA test showed that the N2S4 samples were significantly different based on the main effect of time (PERMANOVA r2 = 0.17, p < 0.01) and the interaction between time and depth horizon (PERMANOVA r2 = 0.18, p < 0.01), but not on depth horizon alone (PERMANOVA r2 = 0.04, p = 0.4; Figure 7b). There were decreasing trends in Shannon evenness and increasing trends in Chao1 richness for both the euphotic and upper mesopelagic communities over the occupation of N2S4, but these changes were not statistically different (Figure S5).

Figure 6.

Alpha diversity at NAAMES 2 Station 4 and the NAAMES 3 stations. Two diversity indices are depicted: (a) Shannon evenness index for bacterioplankton communities in the euphotic zone (left) and upper mesopelagic zone (right) and (b) Chao1 richness estimates for bacterioplankton communities in the euphotic zone (left) and upper mesopelagic zone (right). Filled circles represent the indices of each sample, with color representing the depth of collection. Boxes represent the 1.5 interquartile range, with the internal solid line representing the median; P values are reported for t tests between the means of the alpha diversity indices. At NAAMES 2 Station 4 (N2S4), the euphotic zone samples include those from 5 to 75 m, while the mesopelagic zone samples include those from 100 to 200 m. At the NAAMES 3 stations (Stations 3, 3.5, 4, and 6), the euphotic zone samples include those from 5 to 25 m, while the mesopelagic zone samples include those from 100 to 200 m. NAAMES = North Atlantic Aerosols and Marine Ecosystems Study.

Figure 6.

Alpha diversity at NAAMES 2 Station 4 and the NAAMES 3 stations. Two diversity indices are depicted: (a) Shannon evenness index for bacterioplankton communities in the euphotic zone (left) and upper mesopelagic zone (right) and (b) Chao1 richness estimates for bacterioplankton communities in the euphotic zone (left) and upper mesopelagic zone (right). Filled circles represent the indices of each sample, with color representing the depth of collection. Boxes represent the 1.5 interquartile range, with the internal solid line representing the median; P values are reported for t tests between the means of the alpha diversity indices. At NAAMES 2 Station 4 (N2S4), the euphotic zone samples include those from 5 to 75 m, while the mesopelagic zone samples include those from 100 to 200 m. At the NAAMES 3 stations (Stations 3, 3.5, 4, and 6), the euphotic zone samples include those from 5 to 25 m, while the mesopelagic zone samples include those from 100 to 200 m. NAAMES = North Atlantic Aerosols and Marine Ecosystems Study.

Close modal
Figure 7.

Beta diversity at NAAMES 2 Station 4 and the NAAMES 3 Stations. Nonmetric multidimensional scaling ordinations with Bray–Curtis dissimilarities for samples taken (a) at all stations (stress = 0.05), with NAAMES 3 data color-coded by station and (b) only at NAAMES 2 Station 4 (N2S4; stress = 0.09), with data color-coded by sampling time. At N2S4, the euphotic zone samples include those from 5 to 75 m, while the mesopelagic zone samples include those from 100 to 200 m. At the NAAMES 3 stations, the euphotic zone samples include those from 5 to 25 m, while the mesopelagic zone samples include those from 100 to 200 m. The 95% confidence interval ellipses are drawn around group centroids based on cruise and depth horizon (euphotic and upper mesopelagic zone) in (a), and on time in (b). In (a), samples were significantly different based on cruise (PERMANOVA r2 = 0.37, p < 0.01) and depth horizon (PERMANOVA r2 = 0.04, p = 0.03). In (b), samples were also significantly different based on the main effect of time (PERMANOVA r2 = 0.17, p < 0.01) and the interaction between time and depth horizon (PERMANOVA r2 = 0.18, p < 0.01), but not on depth horizon alone (PERMANOVA r2 = 0.04, p = 0.4). NAAMES = North Atlantic Aerosols and Marine Ecosystems Study; PERMANOVA = permutational multivariate analysis of variance.

Figure 7.

Beta diversity at NAAMES 2 Station 4 and the NAAMES 3 Stations. Nonmetric multidimensional scaling ordinations with Bray–Curtis dissimilarities for samples taken (a) at all stations (stress = 0.05), with NAAMES 3 data color-coded by station and (b) only at NAAMES 2 Station 4 (N2S4; stress = 0.09), with data color-coded by sampling time. At N2S4, the euphotic zone samples include those from 5 to 75 m, while the mesopelagic zone samples include those from 100 to 200 m. At the NAAMES 3 stations, the euphotic zone samples include those from 5 to 25 m, while the mesopelagic zone samples include those from 100 to 200 m. The 95% confidence interval ellipses are drawn around group centroids based on cruise and depth horizon (euphotic and upper mesopelagic zone) in (a), and on time in (b). In (a), samples were significantly different based on cruise (PERMANOVA r2 = 0.37, p < 0.01) and depth horizon (PERMANOVA r2 = 0.04, p = 0.03). In (b), samples were also significantly different based on the main effect of time (PERMANOVA r2 = 0.17, p < 0.01) and the interaction between time and depth horizon (PERMANOVA r2 = 0.18, p < 0.01), but not on depth horizon alone (PERMANOVA r2 = 0.04, p = 0.4). NAAMES = North Atlantic Aerosols and Marine Ecosystems Study; PERMANOVA = permutational multivariate analysis of variance.

Close modal

During the restratification at N2S4, subclades (approximately family level) of SAR11 dominated in relative abundance. During the occupation of N2S4, SAR11 relative abundance increased by approximately 7% in the euphotic zone to represent approximately 45% of the bacterial community, whereas in the upper mesopelagic, it increased by approximately 18% to represent approximately 47% of the community (Figure 8). The relative abundance of SAR 11 subclade 1a was greatest of all SAR11 subclades and increased by approximately 9% and 16% to represent approximately 35% and 36% of the total community in the euphotic and upper mesopelagic zones, respectively, throughout the station occupation (Figure 8). Of the remaining SAR11 subclades (I, IB, II, IV), subclades I, IB, and II showed increases (> 12%) in relative abundance in the upper 200 m, individually comprising less than or equal to 5% of the bacterioplankton community in each of the euphotic and mesopelagic zones (Figure 8). Members of the Oceanospirillales (of the Gammaproteobacteria) were the next most relatively abundant members of the total bacterioplankton community, particularly ZD0405, SAR86, and Oceanospirillaceae; altogether, they represented 17% and 18% of the community in the euphotic and mesopelagic zones, respectively. Apart from Oceanospirillaceae, the relative abundance of these families did not change significantly in the euphotic zone. However, in the upper mesopelagic zone, the relative abundance of Oceanospirillales decreased by approximately 19% with families ZD405 decreasing by approximately 19%, SAR86 decreasing by approximately 15%, and Oceanospirillaceae by approximately 32% (Figure 8). Flavobacteriales (8% ± 1%, including families Flavobacteriaceae, NS9_marine_group, Cryomorphaeceae, NS7_marine_group), Rhodobacterales (4% ± 1%, family Rhodobacteraceae), and Rhodospirillaceae (3% ± 1%, family Rhodospirillaceae) were the next most relatively abundant orders across both depth horizons, contributing to greater than or equal to approximately 2% of the total bacterioplankton community. These bacterial lineages showed < 12% changes in relative abundance over time (Figure 8). The change in relative abundance for all remaining orders and families in both depth horizons over the station occupation of N2S4 was up to 50%, but these taxa contributed to < 2% of the total bacterioplankton community (Figure 9).

Figure 8.

Taxonomic abundances at NAAMES 2 Station 4. Relative abundances and 16 S rRNA estimated cell abundances of taxonomic groups to the family level for the euphotic (0–75 m) and upper mesopelagic (100–200 m) zones at NAAMES 2 Station 4. Fill color within each box represents 16 S rRNA estimated cell abundances, while values within-box represent relative abundance. Blank boxes indicate the absence of a taxonomic group. Only families that represent at least 1% of the total community were included. For each zone, the relative abundance of a taxonomic group was calculated by summing its amplicon sequence variant counts across all depths in the depth horizon and then dividing the summed counts by the total counts in the depth horizon. The 16 S rRNA gene-estimated cell abundances for each family were computed as the product between its relative abundance and the total cell abundance in the respective depth horizon, normalized to its gene copy number. The 16 S rRNA gene-estimated cell abundances were not computed for unassigned families. NAAMES = North Atlantic Aerosols and Marine Ecosystems Study.

Figure 8.

Taxonomic abundances at NAAMES 2 Station 4. Relative abundances and 16 S rRNA estimated cell abundances of taxonomic groups to the family level for the euphotic (0–75 m) and upper mesopelagic (100–200 m) zones at NAAMES 2 Station 4. Fill color within each box represents 16 S rRNA estimated cell abundances, while values within-box represent relative abundance. Blank boxes indicate the absence of a taxonomic group. Only families that represent at least 1% of the total community were included. For each zone, the relative abundance of a taxonomic group was calculated by summing its amplicon sequence variant counts across all depths in the depth horizon and then dividing the summed counts by the total counts in the depth horizon. The 16 S rRNA gene-estimated cell abundances for each family were computed as the product between its relative abundance and the total cell abundance in the respective depth horizon, normalized to its gene copy number. The 16 S rRNA gene-estimated cell abundances were not computed for unassigned families. NAAMES = North Atlantic Aerosols and Marine Ecosystems Study.

Close modal
Figure 9.

Taxonomic abundances of select bacterioplankton in the upper mesopelagic zone at NAAMES 2 Station 4. Relative (open black circles) and 16 S rRNA gene-estimated (blue circles) cell abundances in the upper mesopelagic zone of bacterioplankton taxa whose differences over time were a significant contribution to the Bray–Curtis dissimilarities in the upper mesopelagic of N2S4 (similarity percentages p < 0.05), contributed to at least 1% of the total community and showed at least a doubling in 16 S rRNA gene-estimated cell abundances. Panel labels represent the taxonomic designation of order and family. NAAMES = North Atlantic Aerosols and Marine Ecosystems Study.

Figure 9.

Taxonomic abundances of select bacterioplankton in the upper mesopelagic zone at NAAMES 2 Station 4. Relative (open black circles) and 16 S rRNA gene-estimated (blue circles) cell abundances in the upper mesopelagic zone of bacterioplankton taxa whose differences over time were a significant contribution to the Bray–Curtis dissimilarities in the upper mesopelagic of N2S4 (similarity percentages p < 0.05), contributed to at least 1% of the total community and showed at least a doubling in 16 S rRNA gene-estimated cell abundances. Panel labels represent the taxonomic designation of order and family. NAAMES = North Atlantic Aerosols and Marine Ecosystems Study.

Close modal

The total bacterioplankton community abundance increased in the upper mesopelagic (Figures 3h and 4e); thus, if this increase in cell abundance is considered with the changes in relative abundance of each taxonomic family, 16 S rRNA-estimated cell abundances reveal increasing trends for many bacterioplankton families, even those that did not increase in relative abundance. Indeed, as total community abundance increased throughout the surface 200 m of the water column, 87% and 88% of the families represented in the euphotic zone and upper mesopelagic zone, respectively, increased in 16 S rRNA-estimated cell abundance (Figure 8b).

Similarity percentages breakdown (SIMPER) analysis identified 127 families of 75 orders in the upper mesopelagic zone at N2S4; their variability between May 24 and May 26 was primarily responsible for Bray–Curtis dissimilarities (Table S1). Of these groups, six contributed to at least 1% of the total community and showed at least a doubling in 16 S rRNA gene-estimated cell abundance between May 24 and May 26, as shown in Figure 9. These taxa represent the orders Rhodobacterales (110% increase), Puniceispirillales (100% increase), Flavobacteriales (140% increase), and the SAR11 clade (140%–170% increases). These groups showed relative abundance changes ranging from –5% to 29% (Figure 9).

The energy from storm events can erode thermal stratification and lead to convective mixing of accumulated organic matter from the euphotic zone to the upper mesopelagic zone. As storms subside, water column restratification traps the exported particulate and dissolved organic matter within the mesopelagic zone, supporting the carbon and energetic demands of heterotrophic microbes. Capturing one of these mixing events as well as the subsequent restratification is rare for ship-based sampling programs. Here, we have reported the biogeochemical and microbial changes measured during and shortly following a mixing event in the subtropical western North Atlantic. We have demonstrated that the organic matter delivered into and trapped within the upper mesopelagic zone stimulates rapid bacterial production and community structure shifts. However, the change in bacterioplankton community composition was subtle and more gradual than the changes in bacterioplankton stocks and secondary production rates on a timescale of days. We hypothesize that the continued availability of labile DOC, whether delivered from the euphotic zone or produced in situ by particle solubilization, temporarily supports the growth of euphotic zone-associated bacterioplankton at mesopelagic depths.

4.1. Microbial response at station N2S4

Over the 3-day occupation of N2S4, NPP coincided with N + N utilization and the reduction of AOU in the euphotic zone (0–75 m; Figures 3a, e, and g and 4c and d). Increases in relative abundance of phytoplankton were attributed primarily to prasinophytes, diatoms (Bolaños et al., 2020) and picoeukaryotes (Graff and Behrenfeld, 2018).

BA and BP in the euphotic zone also responded positively to mixed layer shoaling following deep convective mixing (Figures 3d and h and 4e and f). Despite NPP outpacing BP in the euphotic zone during the occupation of N2S4, the accumulation of bulk DOC was not resolvable over the timescale of station occupation (Figure 3d and g). Previous work on initial bloom conditions following nutrient entrainment by deep mixing demonstrated that a greater fraction of NPP was partitioned as particulate organic carbon compared to DOC (Carlson et al., 1998; Baetge et al., 2021). Despite our inability to resolve changes in the bulk DOC pool, increases in both the concentration and relative contribution of TDAA to the bulk DOC pool (Figures 3i and 4j) suggest that the quality of the accumulated DOC pool changed to a composition that was less altered diagenetically (i.e., fresh and less recalcitrant) within the euphotic zone throughout station occupation. Amino acids comprise a large proportion of labile DOC and are preferentially and rapidly consumed by heterotrophic bacterioplankton (Keil and Kirchman, 1993; Cowie and Hedges, 1994; Davis and Benner, 2005; Guerrero-Feijóo et al., 2017). The elevated contribution of TDAA to the bulk DOC pool is thus indicative of newly produced DOC (Davis and Benner, 2005; Kaiser and Benner, 2012; Stephens et al., 2020), which can result from a variety of pathways including direct release from phytoplankton (Granum et al., 2002) or archaea (Bayer et al., 2019), microzooplankton egestion (Nagata and Kirchman, 1991), mesozooplankton sloppy feeding, egestion, and excretion (Lampert, 1978; Maas et al., 2020), particle solubilization (Smith et al., 1992), and viral lysis (Middelboe and Jørgensen, 2006). The change in DOC composition supports the hypothesis that the flux of labile DOC at the nmol L–1 level partially supported the enhanced BP in the euphotic zone.

Chl a, phytoplankton cells, and DOC were exported to mesopelagic depths by deep mixing (Figure 3b, c, and f). The net effect of this redistribution is a reduction in the amount of organic matter in the euphotic zone as lower concentrations from depth are mixed to the surface. In turn, organic matter becomes enriched in the mesopelagic zone as a portion of the organic matter accumulated in the euphotic zone is mixed to depth. Subsequent thermal restratification of the upper 200 m then traps the exported organic matter within the mesopelagic zone, where it fuels the mesopelagic food web and metabolic demand (Carlson et al., 1994; Dall’Olmo et al., 2016; Lacour et al., 2019). The observed decreases in Chl a, phytoplankton cell abundance, and DOC in combination with increased AOU in the upper mesopelagic are consistent with enhanced heterotrophy over the 3-day occupation of N2S4 (Figure 3b, c, e, and f).

During the occupation of station N2S4, BA and BP increased throughout the upper mesopelagic as Chl a and phytoplankton cell abundance declined (Figures 3b–d and h and 4a, b, e, and f). While the exact source of available organic matter supporting the elevated BP in the upper mesopelagic is unknown, the increased BP led to an increase in BA and consumption of available organic matter. The extant bacterioplankton community may have been responding to labile DOC delivered from the euphotic zone during mixing. The loss of Chl a and phytoplankton cells suggests that phytoplankton were not yet acclimated to the deeper depths of the upper mesopelagic and may have been grazed upon or solubilized. Grazing or solubilization of organic particles delivered into the upper mesopelagic by deep convective mixing may have been additional potential sources for the freshly produced labile DOC needed to support increased BP (Cho and Azam, 1988; Smith and Azam, 1992; Smith et al., 1995). Significant increases of ASVs in the Flavobacteriaceae family within the upper mesopelagic of N2S4 suggest the potential for particle solubilization by heterotrophic microbes. Taxa from the phylum Bacteroidetes (including Flavobacteria) have been linked with particle-associated lifestyles throughout the global bathypelagic oceans (Salazar et al., 2015; Liu et al., 2018), potentially due to their capability to degrade complex polymeric organic matter (Cottrell and Kirchman, 2000). The presence of members of these bacterial clades in the upper mesopelagic at N2S4 supports the hypothesis of microbially driven organic particle solubilization, thus providing a potential source of labile DOC to the free-living microbial community (Cho and Azam, 1988; Smith et al., 1992; Smith et al., 1995).

Although most changes in bulk DOC and TDAA over the course of station N2S4 occupation were not statistically significant, there was a decreasing trend in bulk DOC, TDAA concentration, and percent contribution to bulk DOC in the upper mesopelagic zone (Figures 3i and 4h). This suggests that the DOC in the upper mesopelagic zone may have started to become qualitatively altered to more recalcitrant material as heterotrophic bacterioplankton preferentially consumed more labile components of DOC. This observation is consistent with the argument that heterotrophic bacterioplankton can alter DOC concentration and quality once isolated in the mesopelagic (Goldberg et al., 2009; Kaiser and Benner, 2009; Guerrero-Feijóo et al., 2017; Liu et al., 2022).

While the data presented here do not provide unequivocal evidence for the degradation of surface accumulated DOC by mesopelagic heterotrophic microbes, they do demonstrate that microbes respond rapidly to the introduction of less diagenetically altered organic matter from the euphotic zone (export of DOC or solubilization of exported suspended organic particles) as the water column physically stratifies following deep convection.

4.2. Gradual changes in bacterioplankton community composition

In the water columns of the NAAMES 3 stations, bacterioplankton communities were stratified over depth, with euphotic communities differentiating from upper mesopelagic communities (Figures 6 and 7a). These observations, reported previously for the NAAMES region (Bolaños et al., 2021), are consistent with studies in other oceanic systems (Giovannoni et al., 1996; Field et al., 1997; DeLong et al., 2006; Sunagawa et al., 2015). Taxonomic richness assessed by the Chao1 diversity index increased with increasing depth (Figure 3), a pattern also observed throughout the global ocean (Sunagawa et al., 2015) that reflects the range of niches resulting from heterogeneity in organic and inorganic matter quality and availability. Processes that result in patchiness of DOC concentration and quality include the spatial variability of excretory products of resident and vertical migrating zooplankton (Maas et al., 2020) and DOC heterogeneity associated with the solubilization of sinking particles (Azam and Long, 2001; Azam and Malfatti, 2007). In contrast to the vertical structure observed at the NAAMES 3 stations, the microbial community at N2S4 was homogenized throughout the surface 200 m during and shortly following deep convective mixing. The median alpha diversity indices of the N2S4 euphotic zone samples were higher relative to those of the NAAMES 3 stations, but comparable to the mesopelagic zone samples of N2S4 (Figure 6). This observation suggests that deep convection mixed the euphotic and mesopelagic microbial communities, thus enhancing community richness in the euphotic zone, while not simultaneously reducing richness in the upper mesopelagic. Similar patterns of enhanced euphotic zone richness due to the entrainment of mesopelagic microbial lineages following deep convective mixing have also been reported on seasonal timescales in studies off the coasts of Northwest Spain and California, USA (Cram et al., 2015; Hernando-Morales et al., 2018).

Previous work has reported shifts in mesopelagic microbial community structure on seasonal timescales following physical or chemical perturbations. For example, following deep convective mixing at the BATS site, shifts in upper mesopelagic bacterioplankton community composition have been observed with increases in the relative abundance of SAR202, Acidimicrobiales clade OM1, Salinisphaeraceae, OCS116, and SAR11 subclade II (Morris et al., 2005; Carlson et al., 2009; Treusch et al., 2009; Giovannoni and Vergin, 2012; Liu et al., 2022). Some of these mesopelagic bacterioplankton are hypothesized to retain the metabolic capability to degrade functionally recalcitrant DOC that accumulates in surface waters on a timescale of weeks (Morris et al., 2005; DeLong et al., 2006; Landry et al., 2017; Saw et al., 2020). SAR11, a group of free-living aerobic heterotrophic alphaproteobacteria that are ubiquitous in the ocean, is comprised of a variety of ecotypes adapted to different ocean niches (Brown et al., 2012; Vergin et al., 2013). Over a 7-year time series in the NW Mediterranean Sea, Salter et al. (2015) observed that the diversity of SAR11 ecotypes in the upper 5 m increased following physical mixing, with the elevated contributions of subclades Ib and II interrupting the predominance of subclade Ia during periods of stratification, low phytoplankton biomass, and severe phosphate limitation. Similarly at BATS, the members of the SAR11 subclade II become prevalent in the upper mesopelagic zone annually as the water column stratifies in late spring, after semi-labile DOC from the surface has been delivered to the upper mesopelagic by winter deep convection (Carlson et al., 2009). Throughout the subtropics in the NAAMES study region, Bolaños et al. (2021) showed that members of the SAR11 subclade II dominated the upper mesopelagic zone following deep convective overturn in the late winter and early spring. Taken together, these studies indicate that the upper mesopelagic bacterioplankton community responds to perturbations over the course of weeks to months.

As N2S4 thermally stratified over the station occupation, we expected that the microbial populations and organic matter trapped at depth would lead to a succession in bacterioplankton community structure, whereby the upper mesopelagic population would differentiate from that of the euphotic zone. Specifically, we hypothesized that the upper mesopelagic community would become more populated by lineages that generally predominate in the mesopelagic zones in the subtropical regions of the western North Atlantic, including SAR202 and SAR11 subclades Ic and II (Bolaños et al., 2021). While we observed statistically significant differentiation of bacterioplankton community structure throughout the water column over the 3-day occupation at N2S4 (PERMANOVA r2 = 0.17, p < 0.01), the bacterioplankton communities did not diverge between the euphotic and upper mesopelagic on time scales of hours to days (PERMANOVA r2 = 0.04, p = 0.4; Figure 7). These data suggest that on these short timescales, organic substrate availability (i.e., DOC flux or recently solubilized POM) resulting from episodic deep mixing events (e.g., via storm-induced winter convection) can supply enough labile organic matter to continue supporting the “euphotic zone” communities that were also mixed into and trapped within the mesopelagic zone (Hernando-Morales et al., 2018). Community composition changes at N2S4 may reflect a transient phase in microbial community structuring that, if observed over a longer period, may transition from a mixed system to a persistently stratified water column with vertical partitioning of euphotic- and mesopelagic-like communities similar to the vertical structure of the NAAMES 3 water columns (Figure 7).

Despite observing rapid changes in BA and BP, the duration of our occupation of N2S4 was simply not long enough to resolve a clear separation in community composition between the euphotic and mesopelagic zones. However, changes in community structure were beginning to take place throughout the water column (PERMANOVA r2 = 0.17, p < 0.01). In the upper mesopelagic zone, notable changes in relative abundance were attributed to the SAR11 clade, particularly subclades Ia, Ib, and II, which increased by 15%, 29%, and 19%, respectively (Figures 8 and 9). While there was little change, if any, in the relative abundance of other bacterioplankton orders and families, classifying an individual taxon as being a “non-responder” can be misleading. In a community that increases in cell density and biomass, an individual taxonomic group that does not change in relative abundance would increase in 16 S rRNA-estimated cell abundance and can be considered a responding taxon to new or changing environmental conditions. For example, the relative abundance of taxonomic groups such as Alteromonadaceae, Alcanivoracaceae, and Rhodobacteraceae did not change significantly in the upper mesopelagic zone throughout the occupation of N2S4, but all at least doubled in 16 S rRNA-estimated cell abundance (Figure 9, Table S2). Indeed, increases in 16 S rRNA-estimated cell abundance throughout the upper mesopelagic zone were observed for 88% of the community members, suggesting that nearly the entire upper mesopelagic community contributed to the increase in BP (Figure 8). We note, however, that our observations relied on 16 S rRNA gene amplification using V1–V2 primers that only target bacterioplankton and plastids. As such, our analysis excluded archaeal responses (Wear et al., 2018), which could have been important given that deep-sea archaea have the capacity to degrade complex carbon compounds (Ouverney and Fuhrman, 2000; Li et al., 2015; Bergauer et al., 2018) and have been observed to assimilate phytoplankton-derived exopolymeric substances (Boutrif et al., 2011). We further caution that our 16 S rRNA-estimated cell abundances cannot be taken as absolute values as they reflect only the relative abundances of bacterioplankton (and not archaeoplankton) applied to total prokaryotic counts. Even so, bacterioplankton families that did not change in relative abundance would still be revealed as responders because of the increase in total prokaryotic abundance over the station occupation.

Increases in the relative and/or 16 S rRNA-estimated cell abundances of bacterioplankton taxa in the upper mesopelagic zone that are typically predominant in the euphotic zone, such as SAR11 subclade Ia and SAR116 (Morris et al., 2005; Bolaños et al., 2021), suggest that some of the organic matter exported to depth by deep convection was labile in character and could support the growth of these taxa that are typically associated with euphotic zone lineages. Such rapid responses of euphotic zone lineages at mesopelagic depths would be difficult to capture in coarsely resolved time-series sampling (e.g., monthly sampling at BATS). Intriguing is that SAR11 subclade Ia increased in relative and 16 S rRNA gene-estimated cell abundances in the upper mesopelagic zone over the short occupation of N2S4, as it is associated with life in the surface ocean under nutrient-limited conditions (Salter et al., 2015; Giovannoni, 2017). Furthermore, this subclade typically blooms in the surface mixed layer during summer when UV irradiation is most intense (Zafiriou et al., 2008). SAR11 subclade Ia has been shown to decline in the mesopelagic following annual deep convection at BATS, being supplanted by more mesopelagic-associated taxa like members of SAR11 subclades Ib and II (Carlson et al., 2009; Vergin et al., 2013), which were also observed to increase in relative and 16 S rRNA gene-estimated cell abundance in the upper mesopelagic zone of N2S4. Members of the SAR11 clade possess a high abundance of transporters targeting low-molecular-weight DOC, such as amino acids and osmolytes, and have C1 pathways for oxidizing volatile compounds (Giovannoni, 2017); thus SAR11 subclade Ia may have benefited from scavenging hydrolysis products and leaked compounds released by other bacterioplankton, such as attached Flavobacteriaceae and Pseudoalteromonadaceae, that can solubilize exported POM (Baty et al., 2000; Williams et al., 2013; Fontanez et al., 2015; Salazar et al., 2015; Liu et al., 2018). As such bacterioplankton solubilized particles, they may have released amino acids and other DOC that fueled production of SAR11 subclade Ia (Giovannoni, 2017; Arnosti et al., 2018). Indeed, the presence and increase of Flavobacteriaceae and Pseudoalteromonadaceae in the upper mesopelagic zone at N2S4 may suggest that they were mixed into mesopelagic depths with the export of suspended and sinking particles and subsequently could facilitate the release of labile organic matter through particle solubilization (Figure 9, Table S2). Given more time as well as sustained water column stratification, we hypothesize that a clear vertical separation of community structure would become evident as the sources of labile DOC and particles are utilized and become limiting. We further hypothesize that as conditions become unfavorable to euphotic zone- or particle-associated bacterioplankton taxa, such as SAR11 subclade Ia, SAR116, Flavobacteriaceae, and Pseudoalteromonadaceae, they will be outcompeted in the upper mesopelagic by taxa that are better adapted to utilizing surface-derived, but more recalcitrant DOC. Although the data presented here do not capture an extensive time series of community structure shifts following deep convection, they do provide unique evidence for the rapid response of heterotrophic bacterioplankton in the mesopelagic zone to the delivery of surface-derived organic matter. These data illustrate how deep convective mixing can impact the mesopelagic food web by stimulating BP on a timescale of days.

The generated data sets for this study can be found in the SeaWiFS Bio-optical Archive and Storage System (SeaBASS; https://seabass.gsfc.nasa.gov/) as well as from the Biological & Chemical Oceanography Data Management Office (BCO-DMO, DOI: HTTPS://DOI.ORG/10.26008/1912/bco-dmo.824623.1). Amplicon sequence data sets can be found in the National Center for Biotechnology Information Sequence Read Archive (NCBI SRA) database under the BioProject identifier PRJNA627189. Seabird Navis BGCi float data are available via the MISC Lab FTP (ftp://misclab.umeoce.maine.edu/experiments/NAAMES/floats/) as well as the Argo Global Data Assembly Center (ftp://ftp.ifremer.fr/ifremer/argo). Phylogenetic databases are available at https://www.github.com/lbolanos32/NAAMES_2020. All processed data, analyses, and code presented here are available on GitHub (https://github.com/nbaetge/naames_multiday).

The supplemental files for this article can be found as follows:

Figures S1–S6. docx

Table S1. csv

Table S2. csv

We wish to thank the entire North Atlantic Aerosols and Marine Ecosystems Study team and the captains, officers, crews, and marine technicians of the R/V Atlantis. We thank Elisa Halewood for dissolved organic carbon sample processing and logistical support. Alyson Santoro provided thoughtful feedback that improved this manuscript. This manuscript also benefited from helpful discussions with Craig Nelson, Anna James, Jacqui Comstock, Brandon Stephens, and Chance English. Finally, we are grateful to the editors and reviewers for their time and comments that helped to strengthen this manuscript.

This project was supported by the National Science Foundation and National Aeronautics and Space Administration (Award Numbers NSF OCE-157943 and NASA 80NSSC18K0437 to CC). North Atlantic Aerosols and Marine Ecosystems Study was supported by the NASA Earth Venture Sub-Orbital program (EVS-2, NNX15AAF30G to MJB). Analysis was partially supported by the Simons Foundation International BIOS-SCOPE program.

The authors have no competing interests to declare.

This manuscript has not been published elsewhere.

Collected the samples: NB, CAC, LB, JRG.

Processed the samples: NB, CAC, LB, JRG, SL, KO.

Analyzed the data: NB, CAC.

Assisted with data reduction and contributed to the revision and editing of the final manuscript: All authors.

All authors are aware of and accept responsibility for this manuscript and have approved the submitted manuscript.

Aluwihare
,
LI
,
Repeta
,
DJ
,
Chen
,
RF.
1997
.
A major biopolymeric component to dissolved organic carbon in surface sea water
.
Nature
387
(
6629
):
166
169
. DOI: https://doi.org/10.1038/387166a0.
Anderson
,
MJ.
2001
.
A new method for non-parametric multivariate analysis of variance
.
Austral Ecology
26
(
1
):
32
46
. DOI: https://doi.org/10.1046/j.1442-9993.2001.01070.x.
Arakawa
,
N
,
Aluwihare
,
LI
,
Simpson
,
AJ
,
Soong
,
R
,
Stephens
,
BM
,
Lane-Coplen
,
D.
2017
.
Carotenoids are the likely precursor of a significant fraction of marine dissolved organic matter
.
Science Advances
3
(
9
):
1
12
. DOI: https://doi.org/10.1126/sciadv.1602976.
Argo
.
2022
.
Argo float data and metadata from the global data assembly centre (Argo GDAC)
.
SEANOE
. DOI: https://doi.org/10.17882/42182.
Available at
https://www.seanoe.org/html/argo-floats-example-of-doi-attribution-policy.htm.
Arnosti
,
C
,
Reintjes
,
G
,
Amann
,
R.
2018
.
A mechanistic microbial underpinning for the size-reactivity continuum of dissolved organic carbon degradation
.
Marine Chemistry
206
:
93
99
. DOI: https://doi.org/10.1016/j.marchem.2018.09.008.
Arrieta
,
JM
,
Mayol
,
E
,
Hansman
,
RL
,
Herndl
,
GJ
,
Dittmar
,
T
,
Duarte
,
CM.
2015
.
Ocean chemistry: Dilution limits dissolved organic carbon utilization in the deep ocean
.
Science
348
(
6232
):
331
333
. DOI: https://doi.org/10.1126/science.1258955.
Azam
,
F
,
Fenchel
,
T
,
Field
,
JG
,
Gray
,
JS
,
Meyer-Reil
,
LA
,
Thingstad
,
F
.
1983
.
The ecological role of water-column microbes in the sea
.
Marine Ecology Progress Series
10
(
3
):
257
263
.
Azam
,
F
,
Long
,
RA.
2001
.
Oceanography: Sea snow microcosms
.
Nature
414
:
495
497
.
Azam
,
F
,
Malfatti
,
F.
2007
.
Microbial structuring of marine ecosystems
.
Nature Reviews Microbiology
5
(
10
):
782
791
. DOI: https://doi.org/10.1038/nrmicro1747.
Azam
,
F
,
Smith
,
DC
,
Steward
,
GF
,
Hagström
,
Å.
1993
.
Bacteria-organic matter coupling and its significance for oceanic carbon cycling
.
Microbial Ecology
28
:
167
179
. DOI: https://doi.org/10.1002/9781118015841.
Baetge
,
N.
2022
.
Naames_multiday
.
Available at
https://github.com/nbaetge/naames_multiday.
Baetge
,
N
,
Behrenfeld
,
MJ
,
Fox
,
J
,
Halsey
,
KH
,
Mojica
,
KDA
,
Novoa
,
A
,
Stephens
,
BM
,
Carlson
,
CA.
2021
.
The seasonal flux and fate of dissolved organic carbon through bacterioplankton in the western North Atlantic
.
Frontiers in Microbiology
12
(
June
):
1
18
. DOI: https://doi.org/10.3389/fmicb.2021.669883.
Baty
,
AM
,
Eastburn
,
CC
,
Diwu
,
Z
,
Techkarnjanaruk
,
S
,
Goodman
,
AE
,
Geesey
,
GG
.
2000
.
Differentiation of chitinase-active and non-chitinase-active subpopulations of a marine bacterium during chitin degradation
.
Applied Environmental Microbiology
66
(
8
):
3566
3573
. DOI: https://doi.org/10.1128/AEM.66.8.3566-3573.2000.
Bayer
,
B
,
Hansman
,
RL
,
Bittner
,
MJ
,
Noriega-Ortega
,
BE
,
Niggemann
,
J
,
Dittmar
,
T
,
Herndl
,
GJ.
2019
.
Ammonia-oxidizing archaea release a suite of organic compounds potentially fueling prokaryotic heterotrophy in the ocean
.
Environmental Microbiology
21
(
11
):
4062
4075
. DOI: https://doi.org/10.1111/1462-2920.14755.
Behrenfeld
,
MJ.
2010
.
Abandoning Sverdrup’s critical depth hypothesis on phytoplankton blooms
.
Ecology
91
(
4
):
977
989
. DOI: https://doi.org/10.1890/09-1207.1.
Behrenfeld
,
MJ
,
O’Malley
,
RT
,
Boss
,
ES
,
Westberry
,
TK
,
Graff
,
JR
,
Halsey
,
KH
,
Milligan
,
AJ
,
Siegel
,
DA
,
Brown
,
MB.
2016
.
Revaluating Ocean warming impacts on global phytoplankton
.
Nature Climate Change
6
(
3
):
323
330
. DOI: https://doi.org/10.1038/nclimate2838.
Benner
,
R
,
Amon
,
RMW.
2015
.
The size-reactivity continuum of major bioelements in the ocean
.
Annual Review of Marine Science
7
(
1
):
185
205
. DOI: https://doi.org/10.1146/annurev-marine-010213-135126.
Benner
,
R
,
Biddanda
,
B.
1998
.
Photochemical transformations of surface and deep marine dissolved organic matter: Effects on bacterial growth
.
Limnology and Oceanography
43
(
6
):
1373
1378
. DOI: https://doi.org/10.4319/lo.1998.43.6.1373.
Bergauer
,
K
,
Fernandez-Guerra
,
A
,
Garcia
,
JAL
,
Sprenger
,
RR
,
Stepanauskas
,
R
,
Pachiadaki
,
MG
,
Jensen
,
ON
,
Herndl
,
GJ.
2018
.
Organic matter processing by microbial communities throughout the Atlantic water column as revealed by metaproteomics
.
Proceedings of the National Academy of Sciences of the United States of America
115
(
3
):
E400
E408
. DOI: https://doi.org/10.1073/pnas.1708779115.
Bolaños
,
LM.
2020
.
NAAMES_2020
.
Available at
https://www.github.com/lbolanos32/NAAMES_2020.
Bolaños
,
LM
,
Choi
,
CJ
,
Worden
,
AZ
,
Baetge
,
N
,
Carlson
,
CA
,
Giovannoni
,
S.
2021
.
Seasonality of the microbial community composition in the North Atlantic
.
Frontiers in Marine Science
8
(
February
):
1
16
. DOI: https://doi.org/10.3389/fmars.2021.624164.
Bolaños
,
LM
,
Karp-Boss
,
L
,
Choi
,
CJ
,
Worden
,
AZ
,
Graff
,
JR
,
Haëntjens
,
N
,
Chase
,
AP
,
Della Penna
,
A
,
Gaube
,
P
,
Morison
,
F
,
Menden-Deuer
,
S
,
Westberry
,
TK
,
O’Malley
,
RT
,
Boss
,
E
,
Behrenfeld
,
MJ
,
Giovannoni
,
SJ.
2020
.
Small phytoplankton dominate western North Atlantic biomass
.
The ISME Journal
14
:
1663
1664
. DOI: https://doi.org/10.1038/s41396-020-0636-0.
Boutrif
,
M
,
Garel
,
M
,
Cottrell
,
MT
,
Tamburini
,
C.
2011
.
Assimilation of marine extracellular polymeric substances by deep-sea prokaryotes in the NW Mediterranean Sea
.
Environmental Microbiology Reports
3
(
6
):
705
709
. DOI: https://doi.org/10.1111/j.1758-2229.2011.00285.x.
Brown
,
MV
,
Lauro
,
FM
,
Demaere
,
MZ
,
Muir
,
L
,
Wilkins
,
D
,
Thomas
,
T
,
Riddle
,
MJ
,
Fuhrman
,
JA
,
Andrews-Pfannkoch
,
C
,
Hoffman
,
JM
,
McQuaid
,
JB
,
Allen
,
A
,
Rintoul
,
SR
,
Cavicchioli
,
R.
2012
.
Global biogeography of SAR11 marine bacteria
.
Molecular Systems Biology
8
(
595
):
1
13
. DOI: https://doi.org/10.1038/msb.2012.28.
Callahan
,
BJ
,
McMurdie
,
PJ
,
Rosen
,
MJ
,
Han
,
AW
,
Johnson
,
AJA
,
Holmes
,
SP.
2016
.
DADA2: High-resolution sample inference from Illumina amplicon data
.
Nature Methods
13
(
7
):
581
583
. DOI: https://doi.org/10.1038/nmeth.3869.
Carlson
,
CA.
2020
.
Bacterial cell counts and dissolved organic carbon (DOC) measurements from R/V Atlantis AT32, AT34, AT38, and AT39-06 in the western North Atlantic Ocean (35°N to 57°N; 45°W) in November 2015, May 2016, September 2017, March–April 2018, 2020-09-22
. DOI: https://doi.org/10.26008/1912/bco-dmo.824623.1; https://hdl.handle.net/1912/26208.
Carlson
,
CA
,
Ducklow
,
HW
,
Hansell
,
DA
,
Smith
,
WO.
1998
.
Organic carbon partitioning during spring phytoplankton blooms in the Ross Sea polynya and the Sargasso Sea
.
Limnology and Oceanography
43
(
3
):
375
386
. DOI: https://doi.org/10.4319/lo.1998.43.3.0375.
Carlson
,
CA
,
Ducklow
,
HW
,
Michaels
,
AF.
1994
.
Annual flux of dissolved organic carbon from the euphotic zone in the northwestern Sargasso Sea
.
Nature
371
(
September
):
405
408
.
Carlson
,
CA
,
Giovannoni
,
SJ
,
Hansell
,
DA
,
Goldberg
,
SJ
,
Parsons
,
R
,
Vergin
,
K.
2004
.
Interactions among dissolved organic carbon, microbial processes, and community structure in the mesopelagic zone of the northwestern Sargasso Sea
.
Limnology and Oceanography
49
(
4 I
):
1073
1083
. DOI: https://doi.org/10.4319/lo.2004.49.4.1073.
Carlson
,
CA
,
Hansell
,
DA.
2015
.
DOM sources, sinks, reactivity, and budgets
, in
Carlson
,
CA
,
Hansell
,
DA
eds.,
Biogeochemistry of marine dissolved organic matter
. 2nd Edition.
San Diego, CA
:
Academic Press
:
65
126
. DOI: https://doi.org/10.1016/B978-0-12-405940-5.00003-0.
Carlson
,
CA
,
Hansell
,
DA
,
Nelson
,
NB
,
Siegel
,
DA
,
Smethie
,
WM
,
Khatiwala
,
S
,
Meyers
,
MM
,
Halewood
,
E.
2010
.
Dissolved organic carbon export and subsequent remineralization in the mesopelagic and bathypelagic realms of the North Atlantic basin
.
Deep-Sea Research II
57
(
16
):
1433
1445
. DOI: https://doi.org/10.1016/j.dsr2.2010.02.013.
Carlson
,
CA
,
Hansell
,
DA
,
Tamburini
,
C.
2011
. DOC persistence and its fate after export within the ocean interior, in
Jiao
,
N
,
Azam
,
F
,
Sanders
,
S
eds.,
Microbial carbon pump in the ocean
.
Washington, DC
:
Science/AAAS
:
57
59
.
Carlson
,
CA
,
Morris
,
R
,
Parsons
,
R
,
Treusch
,
AH
,
Giovannoni
,
SJ
,
Vergin
,
K.
2009
.
Seasonal dynamics of SAR11 populations in the euphotic and mesopelagic zones of the northwestern Sargasso Sea
.
The ISME Journal
3
(
3
):
283
295
. DOI: https://doi.org/10.1038/ismej.2008.117.
Cho
,
BC
,
Azam
,
F.
1988
.
Major role of bacteria in biogeochemical fluxes in the ocean’s interior
.
Nature
332
(
6163
):
441
443
. DOI: https://doi.org/10.1038/332441a0.
Clarke
,
KR.
1993
.
Non-parametric multivariate analyses of changes in community structure
.
Australian Journal of Ecology
18
(
1
):
117
143
. DOI: https://doi.org/10.1111/j.1442-9993.1993.tb00438.x.
Copin-Montégut
,
G
,
Avril
,
B.
1993
.
Vertical distribution and temporal variation of dissolved organic carbon in the north-western Mediterranean Sea
.
Deep-Sea Research I
40
(
10
):
1963
1972
. DOI: https://doi.org/10.1016/0967-0637(93)90041-Z.
Cottrell
,
MT
,
Kirchman
,
DL.
2000
.
Natural assemblages of marine proteobacteria and members of the Cytophaga-flavobacter cluster consuming low- and high-molecular-weight dissolved organic matter
.
Applied Environmental Microbiology
66
(
4
):
1692
1697
. DOI: https://doi.org/10.1128/AEM.66.4.1692-1697.2000.
Cowie
,
GL
,
Hedges
,
JI.
1994
.
Biochemical indicators of diagenetic alteration in natural organic matter mixtures
.
Nature
369
(
May
):
489
492
. DOI: https://doi.org/10.1038/369304a0.
Cram
,
JA
,
Chow
,
CET
,
Sachdeva
,
R
,
Needham
,
DM
,
Parada
,
AE
,
Steele
,
JA
,
Fuhrman
,
JA.
2015
.
Seasonal and interannual variability of the marine bacterioplankton community throughout the water column over ten years
.
The ISME Journal
9
(
3
):
563
580
. DOI: https://doi.org/10.1038/ismej.2014.153.
Dall’Olmo
,
G
,
Dingle
,
J
,
Polimene
,
L
,
Brewin
,
RJW
,
Claustre
,
H.
2016
.
Substantial energy input to the mesopelagic ecosystem from the seasonal mixed-layer pump
.
Nature Geoscience
9
(
11
):
820
823
. DOI: https://doi.org/10.1038/ngeo2818.
Davis
,
J
,
Benner
,
R.
2005
.
Seasonal trends in the abundance, composition and bioavailability of particulate and dissolved organic matter in the Chukchi/Beaufort Seas and western Canada Basin
.
Deep-Sea Research II
52
(
24–26
):
3396
3410
. DOI: https://doi.org/10.1016/j.dsr2.2005.09.006.
Della Penna
,
A
,
Gaube
,
P.
2019
.
Overview of (sub)mesoscale ocean dynamics for the NAAMES field program
.
Frontiers in Marine Science
6
(
July
):
384
. DOI: https://doi.org/10.3389/fmars.2019.00384.
DeLong
,
EF
,
Preston
,
CM
,
Mincer
,
T
,
Rich
,
V
,
Hallam
,
SJ
,
Frigaard
,
N
,
Martinez
,
A
,
Sullivan
,
MB
,
Edwards
,
R
,
Brito
,
BR
,
Chisholm
,
SW
,
Karl
,
DM.
2006
.
Community genomics among stratified microbial assemblages in the ocean’s interior
.
Science
311
(
5760
):
496
503
. DOI: https://doi.org/10.1126/science.1120250.
Dittmar
,
T.
2015
.
Reasons behind the long-term stability of dissolved organic matter
, in
Carlson
,
CA
,
Hansell
,
DA
eds.,
Biogeochemistry of marine dissolved organic matter
. 2nd Edition.
San Diego, CA
:
Academic Press
:
369
388
. DOI: https://doi.org/10.1016/B978-0-12-405940-5.00007-8.
Ducklow
,
HW.
1986
.
Bacterial biomass in warm-core Gulf Stream ring 82-B: Mesoscale distributions, temporal changes and production
.
Deep-Sea Research
33
(
11–12
):
1789
1812
. DOI: https://doi.org/10.1016/0198-0149(86)90080-4.
Ducklow
,
HW.
1999
.
The bacterial component of the oceanic euphotic zone
.
FEMS Microbiology Ecology
30
(
1
):
1
10
.
Ducklow
,
HW
,
Carlson
,
CA
,
Bates
,
NR
,
Knap
,
AH
,
Michaels
,
AF.
1995
.
Dissolved organic carbon as a component of the biological pump in the North Atlantic Ocean
.
Philosophical Transactions of the Royal Society B
348
(
1324
):
161
167
. DOI: https://doi.org/10.1098/rstb.1995.0058.
Ducklow
,
HW
,
Smith
,
DC
,
Campbell
,
L
,
Landry
,
MR
,
Quinby
,
HL
,
Steward
,
GF
,
Azam
,
F.
2001
.
Heterotrophic bacterioplankton in the Arabian Sea: Basinwide response to year-round high primary productivity
.
Deep-Sea Research Part II
48
(
6–7
):
1303
1323
. DOI: https://doi.org/10.1016/S0967-0645(00)00140-5.
Dunne
,
JP
,
Sarmiento
,
JL
,
Gnanadesikan
,
A.
2007
.
A synthesis of global particle export from the surface ocean and cycling through the ocean interior and on the seafloor
.
Global Biogeochemical Cycles
21
(
4
):
1
16
. DOI: https://doi.org/10.1029/2006GB002907.
Field
,
KG
,
Gordon
,
D
,
Wright
,
T
,
Rappé
,
M
,
Urbach
,
E
,
Vergin
,
K
,
Giovannoni
,
SJ.
1997
.
Diversity and depth-specific distribution of SAR11 cluster rRNA genes from marine planktonic bacteria
.
Applied Environmental Microbiology
63
(
1
):
63
70
. DOI: https://doi.org/10.1128/aem.63.1.63-70.1997.
Fontanez
,
KM
,
Eppley
,
JM
,
Samo
,
TJ
,
Karl
,
DM
,
DeLong
,
EF.
2015
.
Microbial community structure and function on sinking particles in the North Pacific Subtropical Gyre
.
Frontiers in Microbiology
6
(
May
):
1
14
. DOI: https://doi.org/10.3389/fmicb.2015.00469.
Fox
,
J
,
Behrenfeld
,
MJ
,
Haëntjens
,
N
,
Chase
,
A
,
Kramer
,
SJ
,
Boss
,
E
,
Karp-Boss
,
L
,
Fisher
,
NL
,
Penta
,
WB
,
Westberry
,
TK
,
Halsey
,
K.
2020
.
Phytoplankton growth and productivity in the western North Atlantic: Observations of regional variability from the NAAMES field campaigns
.
Frontiers in Marine Science
7
:
24
. DOI: https://doi.org/10.3389/fmars.2020.00024.
Gardner
,
WD
,
Chung
,
SP
,
Richardson
,
MJ
,
Walsh
,
ID.
1995
.
The oceanic mixed-layer pump
.
Deep-Sea Research Part II
42
(
2–3
):
757
775
. DOI: https://doi.org/10.1016/0967-0645(95)00037-Q.
Giovannoni
,
SJ.
2017
.
SAR11 bacteria: The most abundant plankton in the oceans
.
Annual Review of Marine Science
9
(
1
):
231
255
. DOI: https://doi.org/10.1146/annurev-marine-010814-015934.
Giovannoni
,
SJ
,
Rappé
,
MS
,
Vergin
,
KL
,
Adair
,
NL
.
1996
.
16 S rRNA genes reveal stratified open ocean bacterioplankton populations related to the green non-sulfur bacteria
.
Proceedings of the National Academy of Sciences of the United States of America
93
(
15
):
7979
7984
. DOI: https://doi.org/10.1073/pnas.93.15.7979.
Giovannoni
,
SJ
,
Vergin
,
KL.
2012
.
Seasonality in ocean microbial communities
.
Science
335
(
6069
):
671
676
. DOI: https://doi.org/10.1126/science.1198078.
Goldberg
,
SJ
,
Carlson
,
CA
,
Hansell
,
DA
,
Nelson
,
NB
,
Siegel
,
DA.
2009
.
Temporal dynamics of dissolved combined neutral sugars and the quality of dissolved organic matter in the Northwestern Sargasso Sea
.
Deep-Sea Research Part I
56
(
5
):
672
685
. DOI: https://doi.org/10.1016/j.dsr.2008.12.013.
Goto
,
S
,
Tada
,
Y
,
Suzuki
,
K
,
Yamashita
,
Y.
2020
.
Evaluation of the production of dissolved organic matter by three marine bacterial strains
.
Frontiers in Microbiology
11
(
October
):
1
15
. DOI: https://doi.org/10.3389/fmicb.2020.584419.
Graff
,
JR
,
Behrenfeld
,
MJ.
2018
.
Photoacclimation responses in subarctic Atlantic phytoplankton following a natural mixing-restratification event
.
Frontiers in Marine Science
5
(
June
):
209
. DOI: https://doi.org/10.3389/fmars.2018.00209.
Granum
,
E
,
Kirkvold
,
S
,
Myklestad
,
SM.
2002
.
Cellular and extracellular production of carbohydrates and amino acids by the marine diatom Skeletonema costatum: Diel variations and effects of N depletion
.
Marine Ecology Progress Series
242
:
83
94
. DOI: https://doi.org/10.3354/meps242083.
Gruber
,
DF
,
Simjouw
,
JP
,
Seitzinger
,
SP
,
Taghon
,
GL.
2006
.
Dynamics and characterization of refractory dissolved organic matter produced by a pure bacterial culture in an experimental predator-prey system
.
Applied Environmental Microbiology
72
(
6
):
4184
4191
. DOI: https://doi.org/10.1128/AEM.02882-05.
Guerrero-Feijóo
,
E
,
Nieto-Cid
,
M
,
Sintes
,
E
,
Dobal-Amador
,
V
,
Hernando-Morales
,
V
,
Álvarez
,
M
,
Balagué
,
V
,
Varela
,
MM.
2017
.
Optical properties of dissolved organic matter relate to different depth-specific patterns of archaeal and bacterial community structure in the North Atlantic Ocean
.
FEMS Microbiology Ecology
93
(
1
):
1
14
. DOI: https://doi.org/10.1093/femsec/fiw224.
Halewood
,
E
,
Opalk
,
K
,
Custals
,
L
,
Carey
,
M
,
Hansell
,
DA
,
Carlson
,
CA.
2022
.
GO-SHIP Repeat Hydrography: Determination of dissolved organic carbon (DOC) and total dissolved nitrogen (TDN) in seawater using high temperature combustion analysis [GOOS ENDORSED PRACTICE]: 1–52
.
Available at
https://repository.oceanbestpractices.org/handle/11329/1921.
Hansell
,
DA.
2013
.
Recalcitrant dissolved organic carbon fractions
.
Annual Review of Marine Science
5
:
421
445
. DOI: https://doi.org/10.1146/annurev-marine-120710-100757.
Hansell
,
DA
,
Carlson
,
CA.
2001
.
Biogeochemistry of total organic carbon and nitrogen in the Sargasso Sea: Control by convective overturn
.
Deep-Sea Research Part II
48
(
8–9
):
1649
1667
. DOI: https://doi.org/10.1016/S0967-0645(00)00153-3.
Hansell
,
DA
,
Carlson
,
CA
,
Repeta
,
DJ
,
Schlitzer
,
R.
2009
.
Dissolved organic matter in the ocean a controversy stimulates new insights
.
Oceanography
22
(
4
):
202
211
. DOI: https://doi.org/10.5670/oceanog.2009.109.
Hansell
,
DA
,
Carlson
,
CA
,
Schlitzer
,
R.
2012
.
Net removal of major marine dissolved organic carbon fractions in the subsurface ocean
.
Global Biogeochemical Cycles
26
(
1
):
1
9
. DOI: https://doi.org/10.1029/2011GB004069.
Hernando-Morales
,
V
,
Varela
,
MM
,
Needham
,
DM
,
Cram
J
,
Fuhrman
,
JA
,
Teira
,
E.
2018
.
Vertical and seasonal patterns control bacterioplankton communities at two horizontally coherent coastal upwelling sites off Galicia (NW Spain)
.
Microbial Ecology
76
(
4
):
866
884
. DOI: https://doi.org/10.1007/s00248-018-1179-z.
Jiao
,
N
,
Herndl
,
GJ
,
Hansell
,
DA
,
Benner
,
R
,
Kattner
,
G
,
Wilhelm
,
SW
,
Kirchman
,
DL
,
Weinbauer
,
MG
,
Luo
,
T
,
Chen
,
F
,
Azam
,
F.
2010
.
Microbial production of recalcitrant dissolved organic matter: Long-term carbon storage in the global ocean
.
Nature Reviews Microbiology
8
(
8
):
593
599
. DOI: https://doi.org/10.1038/nrmicro2386.
Kaiser
,
K
,
Benner
,
R.
2009
.
Biochemical composition and size distribution of organic matter at the Pacific and Atlantic time-series stations
.
Marine Chemistry
113
(
1–2
):
63
77
. DOI: https://doi.org/10.1016/j.marchem.2008.12.004.
Kaiser
,
K
,
Benner
,
R.
2012
.
Organic matter transformations in the upper mesopelagic zone of the North Pacific: Chemical composition and linkages to microbial community structure
.
Journal of Geophysical Research: Oceans
117
(
1
):
1
12
. DOI: https://doi.org/10.1029/2011JC007141.
Kattner
,
G
,
Meinhard
,
S
,
Koch
,
BP.
2011
. Molecular characterization of dissolved organic matter and constraints for prokaryotic utilization, in
Jiao
,
N
,
Azam
,
F
,
Sanders
,
S
eds.,
Microbial carbon pump in the Ocean
.
Washington, DC
:
Science/AAAS
:
60
61
.
Keil
,
RG
,
Kirchman
,
DL.
1993
.
Dissolved combined amino acids: Chemical form and utilization by marine bacteria
.
Limnology and Oceanography
38
(
6
):
1256
1270
. DOI: https://doi.org/10.4319/lo.1993.38.6.1256.
Kieber
,
RJ
,
Hydro
,
LH
,
Seaton
,
PJ.
1997
.
Photooxidation of triglycerides and fatty acids in seawater: Implication toward the formation of marine humic substances
.
Limnology and Oceanography
42
(
6
):
1454
1462
. DOI: https://doi.org/10.4319/lo.1997.42.6.1454.
Lacour
,
L
,
Briggs
,
N
,
Claustre
,
H
,
Ardyna
,
M
,
Dall’Olmo
,
G.
2019
.
The intraseasonal dynamics of the mixed layer pump in the subpolar North Atlantic Ocean: A biogeochemical-Argo float approach
.
Global Biogeochemical Cycles
33
(
3
):
266
281
. DOI: https://doi.org/10.1029/2018GB005997.
Lampert
,
W.
1978
.
Release of dissolved organic carbon by grazing zooplankton
.
Limnology and Oceanography
23
(
4
):
831
834
. DOI: https://doi.org/10.4319/lo.1978.23.4.0831.
Landry
,
Z
,
Swa
,
BK
,
Herndl
,
GJ
,
Stepanauskas
,
R
,
Giovannoni
,
SJ.
2017
.
SAR202 genomes from the dark ocean predict pathways for the oxidation of recalcitrant dissolved organic matter
.
MBio
8
(
2
):
e00413
17
. DOI: https://doi.org/10.1128/mBio.00413-17.
Li
,
M
,
Baker
,
BJ
,
Anantharaman
,
K
,
Jain
,
S
,
Breier
,
JA
,
Dick
,
GJ.
2015
.
Genomic and transcriptomic evidence for scavenging of diverse organic compounds by widespread deep-sea archaea
.
Nature Communications
6
(
1
):
1
6
. DOI: https://doi.org/10.1038/ncomms9933.
Liu
,
R
,
Wang
,
L
,
Liu
,
Q
,
Wang
,
Z
,
Li
,
Z
,
Fang
,
J
,
Zhang
,
L
,
Luo
,
M.
2018
.
Depth-resolved distribution of particle-attached and free-living bacterial communities in the water column of the New Britain Trench
.
Frontiers in Microbiology
9
(
April
):
1
12
. DOI: https://doi.org/10.3389/fmicb.2018.00625.
Liu
,
S
,
Baetge
,
N
,
Comstock
,
J
,
Opalk
,
K
,
Parsons
,
R
,
Halewood
,
E
,
English
,
CJ
,
Giovannoni
,
S
,
Bolaños
,
LM
,
Nelson
,
CE
,
Vergin
,
K
,
Carlson
,
C.
2020
a.
Stable isotope probing identifies bacterioplankton lineages capable of utilizing dissolved organic matter across a range of bioavailability
.
Frontiers in Microbiology
11
:
580397
. DOI: https://doi.org/10.3389/fmicb.2020.580397.
Liu
,
S
,
Longnecker
,
K
,
Kujawinski
,
EB
,
Vergin
,
K
,
Bolaños
,
LM
,
Giovannoni
,
SJ
,
Parsons
,
RJ
,
Opalk
,
KL
,
Halewood
,
E
,
Hansell
,
DA.
2022
.
Linkages among dissolved organic matter export, dissolved metabolites and associated microbial community structure response in the northwestern Sargasso Sea on a seasonal scale
.
Frontiers in Microbiology
13
:
833252
. DOI: https://doi.org/10.3389/fmicb.2022.833252.
Liu
,
S
,
Parsons
,
R
,
Opalk
,
K
,
Baetge
N
,
Giovannoni
,
S
,
Bolaños
,
LM
,
Kujawinski
,
EB
,
Longnecker
,
K
,
Lu
,
YH
,
Halewood
,
E
,
Carlson
,
C.
2020
b.
Different carboxyl-rich alicyclic molecules proxy compounds select distinct bacterioplankton for oxidation of dissolved organic matter in the mesopelagic Sargasso Sea
.
Limnology and Oceanography
65
(
7
):
1532
1553
. DOI: https://doi.org/10.1002/lno.11405.
Maas
,
AE
,
Liu
,
S
,
Bolaños
,
LM
,
Widner
,
B
,
Parsons
,
R
,
Kujawinski
,
EB
,
Blanco-Bercial
,
L
,
Carlson
,
CA.
2020
.
Migratory zooplankton excreta and its influence on prokaryotic communities
.
Frontiers in Marine Science
7
(
December
):
573268
. DOI: https://doi.org/10.3389/fmars.2020.573268.
Martin
,
M.
2011
.
Cutadapt removes adapter sequences from high-throughput sequencing reads
.
EMBnet Journal
17
(
1
):
10
12
.
McMurdie
,
PJ
,
Holmes
,
S.
2013
.
Phyloseq: An R package for reproducible interactive analysis and graphics of microbiome census data
.
PLoS One
8
(
4
):
e61217
.
Middelboe
,
M
,
Jørgensen
,
NOG.
2006
.
Viral lysis of bacteria: An important source of dissolved amino acids and cell wall compounds
.
Journal of the Marine Biological Association of the United Kingdom
86
(
3
):
605
612
. DOI: https://doi.org/10.1017/S0025315406013518.
Morison
,
F
,
Harvey
,
E
,
Franzè
,
G
,
Menden-Deuer
,
S.
2019
.
Storm-induced predator-prey decoupling promotes springtime accumulation of North Atlantic phytoplankton
.
Frontiers in Marine Science
6
(
SEP
):
608
. DOI: https://doi.org/10.3389/fmars.2019.00608.
Morris
,
RM
,
Vergin
,
KL
,
Cho
,
JC
,
Rappé
,
MS
,
Carlson
,
CA
,
Giovannoni
,
SJ.
2005
.
Temporal and spatial response of bacterioplankton lineages to annual convective overturn at the Bermuda Atlantic Time-Series Study site
.
Limnology and Oceanography
50
(
5
):
1687
1696
. DOI: https://doi.org/10.4319/lo.2005.50.5.1687.
Mueller
,
JL
,
Clark
,
DK
,
Kuwahara
,
VS
,
Lazin
,
G
,
Brown
,
SW
,
Fargion
,
GS
,
Yarbrough
,
MA
,
Feinholz
,
M
,
Flora
,
S
,
Broenkow
,
W
,
Kim
,
Y
,
Johnson
,
B
,
Yuen
,
M
,
Strutton
,
P
,
Dickey
,
T
,
Abbott
,
M
,
Letelier
,
R
,
Lewis
,
M
,
McLean
,
S
,
Chavez
,
F
,
Barnard
,
A
,
Morrison
,
J
,
Subramaniam
,
A
,
Manov
,
D
,
Zheng
,
X
,
Harding
,
L
,
Barnes
,
R
,
Lykke
,
K.
2003
.
Ocean optics protocols for satellite ocean color sensor validation, revision 4, Volume VI: Special topics in ocean protocols and appendices
.
NASA Technical Memorandum
VOL
VI
(
211621
):
1
124
.
Nagata
,
T
,
Kirchman
,
DL.
1991
.
Release of dissolved free and combined amino acids by bacterivorous marine flagellates
.
Limnology and Oceanography
36
(
3
):
433
443
. DOI: https://doi.org/10.4319/lo.1991.36.3.0433.
North Atlantic Aerosols and Marine Ecosystems Study
.
2018
.
SeaWiFS Bio-optical archive and storage system (SeaBASS)
.
NASA
. DOI: https://doi.org/10.5067/SeaBASS/NAAMES/DATA001.
North Atlantic Aerosols and Marine Ecosystems Study Floats
.
2018
.
Bio-argo floats used in the study area of NAAMES
.
University of Maine In-situ Sound & Color Lab FTP
.
Available at
http://misclab.umeoce.maine.edu/floats/.
Ogawa
,
H
,
Amagai
,
Y
,
Koike
,
I
,
Kaiser
,
K
,
Benner
,
R.
2001
.
Production of refractory dissolved organic matter by bacteria
.
Science
292
(
5518
):
917
920
. DOI: https://doi.org/10.1126/science.1057627.
Oksanen
,
J
,
Blanchet
,
FG
,
Kindt
,
R
,
Legendre
,
P
,
Minchin
,
PR
,
O’Hara
,
RB
,
Simpson
,
GL
,
Solymos
,
P
,
Stevens
,
MHH
,
Wagner
,
H.
2013
.
Package “vegan
.”
Community Ecology Package, version
2
(
9
):
1
295
.
Ouverney
,
CC
,
Fuhrman
,
JA.
2000
.
Marine planktonic archaea take up amino acids
.
Applied Environmental Microbiology
66
(
11
):
4829
4833
. DOI: https://doi.org/10.1128/AEM.66.11.4829-4833.2000.
Porter
,
KG
,
Feig
,
YS.
1980
.
The use of DAPI for identifying aquatic microfloral
.
Limnology and Oceanography
25
(
5
):
943
948
.
Quast
,
C
,
Pruesse
,
E
,
Yilmaz
,
P
,
Gerken
,
J
,
Schweer
,
T
,
Yarza
,
P
,
Peplies
,
J
,
Glöckner
,
FO.
2013
.
The SILVA ribosomal RNA gene database project: Improved data processing and web-based tools
.
Nucleic Acids Research
41
:
D590
D596
. DOI: https://doi.org/10.1093/nar/gks1219.
Salazar
,
G
,
Cornejo-Castillo
,
FM
,
Borrull
,
E
,
Díez-Vives
,
C
,
Lara
,
E
,
Vaqué
,
D
,
Arrieta
,
JM
,
Duarte
,
CM
,
Gasol
,
JM
,
Acinas
,
SG.
2015
.
Particle-association lifestyle is a phylogenetically conserved trait in bathypelagic prokaryotes
.
Molecular Ecology
24
(
22
):
5692
5706
. DOI: https://doi.org/10.1111/mec.13419.
Salter
,
I
,
Galand
,
PE
,
Fagervold
,
SK
,
Lebaron
,
P
,
Obernosterer
,
I
,
Oliver
,
MJ
,
Suzuki
,
MT
,
Tricoire
,
C.
2015
.
Seasonal dynamics of active SAR11 ecotypes in the oligotrophic Northwest Mediterranean Sea
.
The ISME Journal
9
(
2
):
347
360
. DOI: https://doi.org/10.1038/ismej.2014.129.
Saw
,
JHW
,
Nunoura
,
T
,
Hirai
,
M
,
Takaki
,
Y
,
Parsons
,
R
,
Michelsen
,
M
,
Longnecker
,
K
,
Kujawinski
,
EB
,
Stepanauskas
,
R
,
Landry
,
Z
,
Carlson
,
C
,
Giovannoni
,
S.
2020
.
Pangenomics analysis reveals diversification of enzyme families and niche specialization in globally abundant SAR202 bacteria
.
MBio
11
(
1
):
e02975-19
. DOI: https://doi.org/10.1128/mBio.02975-19.
Shen
,
Y
,
Benner
,
R.
2020
.
Molecular properties are a primary control on the microbial utilization of dissolved organic matter in the ocean
.
Limnology and Oceanography
65
(
5
):
1061
1071
. DOI: https://doi.org/10.1002/lno.11369.
Silsbe
,
GM
,
Behrenfeld
,
MJ
,
Halsey
,
KH
,
Milligan
,
AJ
,
Westberry
,
TK.
2016
.
The CAFE model: A net production model for global ocean phytoplankton
.
Global Biogeochem Cycles
30
(
12
):
1756
1777
. DOI: https://doi.org/10.1002/2016GB005521.
Simon
,
M
,
Azam
,
F.
1989
.
Protein content and protein synthesis rates of planktonic marine bacteria
.
Marine Ecology Progress Series
51
:
201
213
. DOI: https://doi.org/10.3354/meps051201.
Smith
,
DC
,
Azam
,
F.
1992
.
A simple, economical method for measuring bacterial protein synthesis rates in seawater using
.
Marine Microbial Food Webs
6
(
2
):
107
114
.
Smith
,
DC
,
Simon
,
M
,
Alldredge
,
AL
,
Azam
,
F.
1992
.
Intense hydrolytic enzyme activity on marine aggregates and implications for rapid particle dissolution
.
Nature
359
(
6391
):
139
142
. DOI: https://doi.org/10.1038/359139a0.
Smith
,
DC
,
Steward
,
GF
,
Long
,
RA
,
Azam
,
F.
1995
.
Bacterial mediation of carbon fluxes during a diatom bloom in a mesocosm
.
Deep-Sea Research Part II
50
(
14–16
):
2519
2536
. DOI: https://doi.org/10.1016/0967-0645(95)00005-B.
Stephens
,
BM
,
Opalk
,
KL
,
Petras
,
D
,
Liu
,
S
,
Comstock
,
J
,
Aluwihare
,
LI
,
Hansell
,
DA
,
Carlson
,
CA.
2020
.
Organic matter composition at Ocean Station Papa affects its bioavailability, bacterioplankton growth efficiency and the responding taxa
.
Frontiers in Marine Science
7
:
590273
.
Stoddard
,
SF
,
Smith
,
BJ
,
Hein
,
R
,
Roller
,
BRK
,
Schmidt
,
TM.
2015
.
Rrndb: Improved tools for interpreting rRNA gene abundance in bacteria and archaea and a new foundation for future development
.
Nucleic Acids Research
43
(
D1
):
D593
D598
. DOI: https://doi.org/10.1093/nar/gku1201.
Sunagawa
,
S
,
Coelho
,
LP
,
Chaffron
,
S
,
Kultima
,
JR
,
Labadie
,
K
,
Salazar
,
G
,
Djahanschiri
,
B
,
Zeller
,
G
,
Mende
,
DR
,
Alberti
,
A
,
Cornejo-Castillo
,
FM
,
Costea
,
PI
,
Cruaud
,
C
,
D’ovidio
,
F
,
Engelen
,
S
,
Ferrera
,
I
,
Gasol
,
JM
,
Guidi
,
L
,
Hildebrand
,
F
,
Kokoszka
,
F
,
Lepoivre
,
C
,
Lima-Mendez
,
G
,
Poulain
,
J
,
Poulos
,
BT
,
Royo-Llonch
,
M
,
Sarmento
,
H
,
Vieira-Silva
,
S
,
Dimier
,
C
,
Picheral
,
M
,
Searson
,
S
,
Kandels-Lewis
,
S
,
Tara Oceans
,
C
,
Bowler
,
C
,
De Vargas
,
C
,
Gorsky
,
G
,
Grimsley
,
N
,
Hingamp
,
P
,
Iudicone
,
D
,
Jaillon
,
O
,
Not
,
F
,
Ogata
,
H
,
Pesant
,
S
,
Speich
,
S
,
Stemmann
,
L
,
Sullivan
,
M
,
Weissenbach
,
J
,
Wincker
,
P
,
Karsenti
,
E
,
Raes
,
J
,
Acinas
,
SG
,
Bork
,
P
.
2015
.
Structure and function of the global ocean microbiome
.
Science
348
(
6237
):
1
10
. DOI: https://doi.org/10.1126/science.1261359.
Treusch
,
AH
,
Vergin
,
KL
,
Finlay
,
LA
,
Donatz
,
MG
,
Burton
,
RM
,
Carlson
,
CA
,
Giovannoni
,
SJ
.
2009
.
Seasonality and vertical structure of microbial communities in an ocean gyre
.
The ISME Journal
3
:
1148
1163
. DOI: https://doi.org/10.1038/ismej.2009.60.
Vergin
,
KL
,
Beszteri
,
B
,
Monier
,
A
,
Thrash
,
CJ
,
Temperton
,
B
,
Treusch
,
AH
,
Kilpert
,
F
,
Worden
,
AZ
,
Giovannoni
,
SJ
.
2013
.
High-resolution SAR11 ecotype dynamics at the Bermuda Atlantic Time-Series Study site by phylogenetic placement of pyrosequences
.
The ISME Journal
7
(
7
):
1322
1332
. DOI: https://doi.org/10.1038/ismej.2013.32.
Wear
,
EK
,
Wilbanks
,
EG
,
Nelson
,
CE
,
Carlson
,
CA
.
2018
.
Primer selection impacts specific population abundances but not community dynamics in a monthly time-series 16 S rRNA gene amplicon analysis of coastal marine bacterioplankton
.
Environmental Microbiology
20
(
8
):
2709
2726
. DOI: https://doi.org/10.1111/1462-2920.14091.
Weiss
,
RF
.
1970
.
The solubility of nitrogen, oxygen and argon in water and seawater
.
Deep-Sea Research and Oceanography Abstracts
17
(
4
):
721
735
. DOI: https://doi.org/10.1016/0011-7471(70)90037-9.
Westberry
,
T
,
Behrenfeld
,
MJ
,
Siegel
,
DA
,
Boss
,
E
.
2008
.
Carbon-based primary productivity modeling with vertically resolved photoacclimation
.
Global Biogeochemical Cycles
22
(
2
):
1
18
. DOI: https://doi.org/10.1029/2007GB003078.
Williams
,
PJ
.
1984
. A review of measurements of respiration rates of marine plankton populations, in
Hobbie
,
J
,
Williams
,
PJ
eds.,
Heterotrophic activity in the sea
.
Boston, MA
:
Springer
. (
NATO conference series; vol. 15
):
357
389
.
Williams
,
TJ
,
Wilkins
,
D
,
Long
,
E
,
Evans
,
F
,
Demaere
,
MZ
,
Raftery
,
MJ
,
Cavicchioli
,
R
.
2013
.
The role of planktonic Flavobacteria in processing algal organic matter in coastal East Antarctica revealed using metagenomics and metaproteomics
.
Environmental Microbiology
15
(
5
):
1302
1317
. DOI: https://doi.org/10.1111/1462-2920.12017.
Zafiriou
,
OC
,
Xie
,
H
,
Nelson
,
NB
,
Najjar
,
RG
,
Wang
,
W
.
2008
.
Diel carbon monoxide cycling in the upper Sargasso Sea near Bermuda at the onset of spring and in midsummer
.
Limnology and Oceanography
53
(
2
):
835
850
. DOI: https://doi.org/10.4319/lo.2008.53.2.0835.
Zakem
,
EJ
,
Cael
,
BB
,
Levine
,
NM
.
2020
.
A unified theory for organic matter accumulation
.
Proceedings of the National Academy of Sciences of the United States of America
118
(
6
):
e2016896118
.

How to cite this article: Baetge, N, Bolaños, LM, Della Penna, A, Gaube, P, Liu, S, Opalk, K, Graff, JR, Giovannoni, SJ, Behrenfeld, MJ, Carlson, CA. 2022. Bacterioplankton response to physical stratification following deep convection. Elementa: Science of the Anthropocene 10(1). DOI: https://doi.org/10.1525/elementa.2021.00078

Domain Editor-in-Chief: Jody W. Deming, University of Washington, Seattle, WA, USA

Associate Editor: Jeff S. Bowman, Integrative Oceanography Division, Scripps Institution of Oceanography, San Diego, CA, USA

Knowledge Domain: Ocean Science

This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International License (CC-BY 4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. See http://creativecommons.org/licenses/by/4.0/.

Supplementary Material