More stories

  • in

    Interactions between temperature and energy supply drive microbial communities in hydrothermal sediment

    The results are organized into subsections on in situ temperature profiles, geochemical gradients, and microbial community data. Geochemical data include concentration and isotopic data of dissolved electron acceptors (sulfate, dissolved inorganic carbon (DIC), δ13C-DIC), electron donors (methane, sulfide, SCOAs), and respiration end products (DIC, methane, sulfide), as well as solid-phase organic carbon pools (total organic carbon (TOC), δ13C-TOC, total nitrogen (TN), TOC:TN (C:N)). Microbial community data include bacterial and archaeal 16S rRNA gene copy numbers and bacterial and archaeal community trends. All geochemical and microbiological data are shown in Supplementary Data 1–4.Temperature profilesThe in situ temperatures and temperature gradients differ greatly among sites and hydrothermal areas (Table 1; Fig. 1a, b, 1st column). Certain locations within the SA (Cold Site) and NSA (MUC02, GC13, MUC12) are uniformly cold (~3–5 °C) and thus serve as low-temperature control sites. The fact that Cold Site has no measurable depth-dependent temperature increase suggests that this site, despite being located within the SA, only has minimal hydrothermal fluid seepage. At two sites from the NSA (GC09, GC10), temperatures increase strongly, reaching ~60 °C at 400 cm below the seafloor, with temperature gradients becoming linear below 50 cm. Everest Mound, Orange Mat, and Cathedral Hill in the SA have the steepest temperature gradients ( >165 °C m−1), reaching >80 °C within 25 cm, whereas Yellow Mat from the SA only reaches ~27 °C at 45 cm. Temperature gradients are near-linear at Everest Mound, Cathedral Hill, and Yellow Mat, and in the top ~15 cm of Orange Mat. Below ~15 cm, the temperatures at Orange Mat are nearly constant.Table 1 Overview of all sampling sites.Full size tableFig. 1: Microbial abundance and community structure in relation to temperature and geochemical gradients.Depth profiles of temperature (1st column), porewater dissolved sulfate, methane, and dissolved inorganic carbon (DIC) concentrations (2nd column), bacterial and archaeal 16S rRNA gene abundances (3rd column), bacterial (4th column) and archaeal community structure (5th column) across the 10 study sites. a Sites from the NSA. b Sites from the SA. Bacteria and Archaea community structure is shown at the phylum level, except in Proteobacteria, which are displayed at the class level (see asterisk). To improve visibility, we adjusted the depth axis range for bacterial and archaeal communities at Everest Mound, only showing the top 10 cm, where microbial 16S rRNA genes were above detection. Sulfate and methane data from the NSA, except those from MUC12, were previously published27.Full size imageConcentrations of methane, sulfate, sulfide, and DICPorewater concentration profiles of methane, sulfate and DIC are consistent with higher microbial activity and higher substrate supplies in hydrothermal seep sediments compared to cold control sites or hydrothermal non-seep sediments.Independent of temperature, sediments without fluid seepage, i.e. the hydrothermal NSA sites (GC09, GC10) and low-temperature control sites (MUC02, MUC12, GC13, Cold Site), have similar concentration profiles of sulfate, methane, and DIC (Fig. 1a, b, 2nd column). Methane remains at background concentrations (≤0.02 mM), suggesting minimal methane production. DIC concentrations increase with depth by ~1–2 mM relative to seawater values (~2 mM). Sulfate decreases but remains near seawater values (~28 mM) throughout MUC02, MUC12, and the hydrothermal GC10, but drops more clearly toward the bottom of the hydrothermal GC09 (to 26.4 mM) and the cold GC13 (to 23.8 mM). The only minor deviation is Cold Site from the SA. At this site, sulfate and DIC concentrations change more with depth (sulfate drops to 23.6 mM; DIC increases to 6.2 mM), suggesting higher microbial activity relative to all hydrothermal and control sites within the NSA. Consistent with this interpretation sulfide (HS−) concentrations increase strongly with depth at Cold Site (from 2500 to 6200 µM) but not at the NSA sites, where sulfide concentrations remain much lower (0–52 µM (Supplementary Fig. 1). Furthermore, δ13C-DIC decreases with sediment depth at Cold Site (from −3.3‰ to −10.3‰), suggesting strong input of DIC from organic carbon mineralization (Supplementary Fig. 2). By contrast, δ13C-DIC remains close to seawater values (~0‰) throughout sediments of all NSA sites (−1.7‰ to −0.2‰).Compared to all NSA sites and Cold Site, sulfate, methane, and DIC concentrations are more variable at the seep sites Yellow Mat, Cathedral Hill, Orange Mat, and Everest Mound (Fig. 1b, 2nd column). Methane concentrations at Yellow Mat, Cathedral Hill, and Orange Mat are much higher than at the non-seep sites, reaching 3.3, 5.2, and 2.8 mM, respectively (no data from Everest Mound). These high methane concentrations, which can be mainly attributed to the input of thermogenic methane from below24, almost certainly underestimate in situ concentrations due to outgassing during core retrieval. Sulfate concentrations decrease more strongly with depth than at the NSA sites or Control Site, consistent with previously observed high sulfate-reducing activity6,7 and advection of sulfate-depleted fluid from below29. Nonetheless, sulfate concentrations remain in the millimolar range throughout cores from Yellow and Orange Mat. By contrast, sulfate is below detection (≤0.1 mM) at ≥4.5 cm sediment depth at Everest Mound, and in an intermittent depth interval at Cathedral Hill (~7.5–19.5 cm), below which it increases back to ~6 mM. High, i.e. millimolar, concentrations of sulfide at Orange Mat and Cathedral Hill are consistent with high rates of in situ microbial sulfate reduction and advective input of sulfide from the thermochemical reduction of sulfate in hotter, abiotic layers below (Supplementary Fig. 1). DIC concentrations reach values of >10 mM at Orange Mat, Cathedral Hill, and Yellow Mat (no data from Everest Mound). DIC concentrations fluctuate around 20 mM DIC throughout the core from Cathedral Hill, suggesting high DIC input from deeper layers. C-isotopic values of this DIC are close to those of seawater (~−3‰), suggesting an inorganic source. By contrast, surface sedimentary DIC concentrations at Yellow Mat and Orange Mat are close to seawater values but increase with depth to ~20 and ~14 mM, respectively. Lower δ13C-DIC values in surface sediments, which decrease further to values of ~−20‰ to −24‰ at Yellow Mat and −14‰ to −18‰ at Orange Mat within the top 10–20 cm, suggest that most of this DIC comes from the microbial or thermogenic breakdown of organic matter and/or the microbial anaerobic oxidation of methane.Trends in dissolved SCOAs across locationsPorewater concentration profiles of SCOAs are consistent with higher input of reactive organic carbon substrates to hydrothermal seep sediments compared to cold control sites or hydrothermal non-seep sediments.SCOA concentrations at the cold control sites and hot NSA sites are low, showing no clear depth-related trends, consistent with absence of SCOA input from below and/or biological controlled SCOA concentrations. SCOAs are dominated by acetate (cold MUC02, MUC12, and GC13: 1–3 µM; hydrothermal GCs: 3–6 µM; Cold Site: 1–7 µM), which was detected along with formate, propionate, and lactate (Fig. 2).Fig. 2: Depth profiles of short-chain organic acid (SCOA) concentrations across locations.Note the differences in concentration ranges on the x-axis and depth ranges on the y-axis (Cathedral Hill: 0–50 cm; GC13, GC09, and GC10: 0–500 cm; all others: 0–40 cm).Full size imageBy contrast, SCOA concentrations at all hydrothermal seep sites except Orange Mat, increase with depth and temperature, consistent with a thermogenic source below the cored interval. At Yellow Mat, acetate concentrations are already elevated at the seafloor (32 µM) and increase to >100 µM at 20 cm depth. Cathedral Hill has a similar acetate concentration profile, but reaches even higher concentrations (250 µM). At the hottest site, Everest Mound, acetate concentrations increase from ~150 µM at the seafloor to steady concentrations of ~600 µM below 3 cm. Formate concentrations are also (locally) elevated at Yellow Mat (5-8 µM), Cathedral Hill (to 14 µM), and Everest Mound (94-265 µM), and propionate concentrations reach high values at Cathedral Hill (to 21.8 µM) and Everest Mound (to 125 µM). The only exception among the seep sites is Orange Mat, where acetate is only slightly elevated (10–20 µM), and formate and propionate remain at background concentrations. These concentrations suggest that either thermogenic SCOA input from below is low at this site, or SCOA concentrations are biologically controlled throughout the core. Unlike the other three SCOAs, lactate concentrations remain low at all seep sites, apart from one outlier at Cathedral Hill (34.5 cm: 17.3 µM), suggesting that lactate is not a major product of thermogenic organic matter breakdown.Trends in solid-phase organic matter poolsAll sites have similar δ13C-TOC isotopic compositions, with values ranging from −19‰ to −23‰, consistent with a predominant phytoplankton origin of sedimentary organic carbon (Supplementary Fig. 3). Yet, depth profiles of TOC and TN follow different patterns across the locations (Fig. 3). All cold control sites have similar TOC (~2–4 wt%) and TN contents (~0.3–0.6 wt%), with slight decreases in values from the seafloor downward. Compared to cold controls, GC09 and GC10 have lower TOC and TN contents (TOC: ~0.5–3 wt%; TN: ~0.0–0.3 wt%), in particular in deeper horizons with elevated temperatures. Seep sites within the SA have the widest ranges. Seep sites have higher TOC in surface sediment compared to control sites, suggesting net organic carbon assimilation and synthesis by microbial growth. TOC values are 16 wt% at the seafloor of Orange Mat and 6–7 wt% at the seafloor of the other three locations, and then decrease strongly within the top 10 cm, reaching values similar to those of cold sites or hot NSA sites below 10 cm. TN values in surface sediments of seep sites are generally higher than at control sites (~0.7–0.9 wt%), providing additional evidence of net organic matter synthesis by microbial biomass production, but then decrease steeply to values that are similar to those at hot NSA sites.Fig. 3: Carbon and nitrogen contents of bulk organic matter.Depth profiles of total organic carbon (TOC), total nitrogen (TN), and TOC:TN (C:N) across all sites.Full size imageAs a result of the stable TOC and TN trends, C:N does not change much with depth at the cold locations. Yet, while C:N ranges around 4.4–5.6 at Cold Site, values are considerably higher, around 8.1–10.1, at cold locations within the NSA. By comparison, the hot NSA sites and all seep sites except Orange Mat show increases in C:N with increasing temperature and depth. This increase in C:N is modest, from ~8 to 10 at Yellow Mat, and more pronounced at the hotter GC09 (to 15.9), GC10 (to 13.4), Cathedral Hill (to 14.6), and Everest Mound (to 15.7). Orange Mat has the highest C:N ratios (14.8–26.5), and unlike the other sites does not show an increase in C:N with depth.General trends in bacterial and archaeal 16S rRNA gene copy numbers16S rRNA gene copy numbers indicate distinct trends in bacterial and archaeal abundances that follow temperature increases with sediment depth (Fig. 1a and b, 3rd column).At the four cold locations, bacterial and archaeal gene copy numbers are relatively stable with depth (Bacteria: 108−109 g−1; Archaea: 107−108 g−1). By comparison, gene copy numbers of GC09 and GC10 are in a similar range near the seafloor but decrease strongly with depth. While Archaea are quantifiable throughout both cores to ≤103 gene copies g−1 sediment, bacterial gene copy numbers are not reliably distinguishable from extraction negative controls (~1 × 104 g−1) at temperatures >60 °C. Furthermore, unlike the cold sites, which consistently have higher bacterial gene copy numbers, there is a shift from bacterial to archaeal dominance in gene copy numbers (GC09: at ~50 cm; GC10: at ~150 cm) at both hot NSA sites.Compared to the hot GCs from the NSA, gene copies decrease over much shorter distances at sites with fluid seepage in the SA. This decrease in gene copy numbers appears related to the magnitude of the temperature increase with depth. At Yellow Mat, which only reaches moderately warm temperatures (27 °C), copy numbers of both domains decrease from ~108 g−1 at the seafloor to ~106 g−1 at the bottom of the core. While Orange Mat, Cathedral Hill, and Everest Mound have similar bacterial and archaeal gene copy numbers to Yellow Mat at the seafloor, these values drop off much more steeply with depth, matching the much steeper temperature increases. At Cathedral Hill and Everest Mound, Bacteria could not be reliably detected below 20 and 7.5 cm, respectively. As the only location, the detection limit of archaeal 16S gene sequences was reached at Everest Mound, at a depth of 9.5 cm.Relationships between microbial gene abundances and temperatureWe explored the relationship between 16S rRNA gene copy number and temperature further (Fig. 4a, b). While gene copy numbers of both domains generally decrease with increasing temperature, the shape of this temperature relationship differs between both domains. In bacteria the decrease in gene copy numbers in relation to temperature is nearly linear. By contrast, in Archaea gene copy numbers follow hump-shaped distributions, i.e. they remain stable or only decrease slightly up to a certain temperature threshold, beyond which their copy numbers decrease steeply. This apparent thermal threshold varies between sites, i.e. it is ~85 °C at Orange Mat, ~70 °C at Cathedral Hill, ~50 °C at the NSA sites, and ~20 °C at Everest Mound.Fig. 4: Gene copy trends in relation to temperature.a Bacterial and (b) archaeal 16S rRNA gene copy numbers vs. temperature. c Bacteria-to-Archaea 16S rRNA gene copy ratios vs. temperature (the exponential function and its coefficient of determination (R2), both calculated in Microsoft Excel, are shown in the graph). Symbol sizes indicate the sediment depth of each sample. Cold control sites from both locations are grouped together in the legend for easier viewing.Full size imageThe differences in relationships between bacterial and archaeal gene copy numbers and temperature are reflected in Bacteria-to-Archaea gene copy ratios (Fig. 4c). Bacterial always exceed archaeal gene copies at 45 °C. Between 10 and 45 °C, domain-level gene dominance varies with location. Despite the variability, Bacteria-to-Archaea gene copy ratios follow a highly significant, exponential relationship with temperature (R2 = 0.67, p  More

  • in

    Individual and population dietary specialization decline in fin whales during a period of ecosystem shift

    1.Hutchinson, G. E. Concluding remarks. Cold Spring Harb. Symp. Quant. Biol. 22, 415–427 (1957).Article 

    Google Scholar 
    2.Bolnick, D. I. et al. The ecology of individuals: Incidence and implications of individual specialization. Am. Nat. 161, 1–28 (2003).MathSciNet 
    PubMed 
    Article 

    Google Scholar 
    3.Araújo, M. S., Bolnick, D. I. & Layman, C. A. The ecological causes of individual specialisation. Ecology 14, 948–958 (2011).
    Google Scholar 
    4.Kassen, R. The experimental evolution of specialists, generalists, and the maintenance of diversity. J. Evol. Biol. 15, 173–190 (2002).Article 

    Google Scholar 
    5.Clavel, J., Julliard, R. & Devictor, V. Worldwide decline of specialist species: Toward a global functional homogenization?. Front. Ecol. Environ. 9, 222–228 (2011).Article 

    Google Scholar 
    6.Svanbäck, R. & Bolnick, D. I. Intraspecific competition drives increased resource use diversity within a natural population. Proc. R. Soc. B Biol. Sci. 274, 839–844 (2007).Article 

    Google Scholar 
    7.Tinker, M. et al. Structure and mechanism of diet specialization: Testing models of individual variation in resource use with sea otters. Ecology 15, 475–483 (2012).
    Google Scholar 
    8.Newsome, S. D. et al. The interaction of intraspecific competition and habitat on individual diet specialization: A near range-wide examination of sea otters. Oecologia 178, 45–59 (2015).ADS 
    PubMed 
    Article 

    Google Scholar 
    9.Layman, C. A., Newsome, S. D. & Crawford, T. G. Individual-level niche specialization within populations: Emerging areas of study. Oecologia 178, 1–4 (2015).ADS 
    PubMed 
    Article 

    Google Scholar 
    10.Moore, S. E., Haug, T., Víkingsson, G. A. & Stenson, G. B. Baleen whale ecology in arctic and subarctic seas in an era of rapid habitat alteration. Prog. Oceanogr. 176, 102–118 (2019).Article 

    Google Scholar 
    11.Møller, E. F. & Nielsen, T. G. Borealization of Arctic zooplankton – Smaller and less fat zooplankton species in Disko Bay, Western Greenland. Limnol. Oceanogr. 65, 1175–1188 (2020).ADS 
    Article 

    Google Scholar 
    12.COSEWIC. COSEWIC Assessment and Status Report on the Fin Whale Balaenoptera physalus, Atlantic Population and Pacific Population, in Canada. xv + 72 (COSEWIC, 2019).13.Kawamura, A. A review of food of balaenopterid whales. Sci. Rep. Whales Res. Inst. 32, 155–197 (1980).
    Google Scholar 
    14.Baumgartner, M. F. & Mate, B. R. Summertime foraging ecology of North Atlantic right whales. Mar. Ecol. Prog. Ser. 264, 123–135 (2003).ADS 
    Article 

    Google Scholar 
    15.Silva, M. A., Prieto, R., Jonsen, I., Baumgartner, M. F. & Santos, R. S. North Atlantic blue and fin whales suspend their spring migration to forage in middle latitudes: building up energy reserves for the journey?. PLoS ONE 8, e76507. https://doi.org/10.1371/journal.pone.0076507 (2013).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    16.Gavrilchuk, K. et al. Trophic niche partitioning among sympatric baleen whale species following the collapse of groundfish stocks in the Northwest Atlantic. Mar. Ecol. Prog. Ser. 497, 285–301 (2014).ADS 
    Article 

    Google Scholar 
    17.Aguilar, A. & García-Vernet, R. Fin whale. in Encyclopedia of Marine Mammals (eds. Würsig, B., Thewissen, J. G. M., Kovacs, K. M.) 368–371 (Elsevier, 2018).18.Silva, M. A. et al. A stable isotopes reveal winter feeding in different habitats in blue, fin and sei whales migrating through the Azores. R. Soc. Open Sci. 6, 181800. https://doi.org/10.1098/rsos.181800 (2019).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    19.Savenkoff, C. et al. Changes in the northern Gulf of St Lawrence ecosystem estimated by inverse modelling: Evidence of a fishery-induced regime shift?. Est. Coast. Shelf Sci. 73, 711–724 (2007).ADS 
    Article 

    Google Scholar 
    20.Galbraith, P. S. et al. Physical oceanographic conditions in the gulf of St. Lawrence during 2019. Can. Sci. Adv. Sec. Res. Doc. 2020/043, iv + 9 (2020).
    Google Scholar 
    21.Myers, R. A. & Worm, B. Extinction, survival or recovery of large predatory fishes. Philos. Trans. R. Soc. Lond. B Biol. Sci. 360, 13–20 (2005).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    22.Plourde, S. et al. Ecosystem perspective on changes and anomalies in the Gulf of St. Lawrence: A context in support of the management of the St. Lawrence beluga whale population. Can. Sci. Adv. Sec. Res. Doc. 2013/129, v + 29 (2014).
    Google Scholar 
    23.Hammill, M. O., Stenson, G. B., Doniol-Valcroze, T. & Mosnier, A. Conservation of northwest Atlantic harp seals: Past success, future uncertainty?. Biol. Conserv. 192, 181–191 (2015).Article 

    Google Scholar 
    24.Stenson, G. B., Haug, T. & Hammill, M. O. Harp seals: Monitors of change in differing ecosystems. Front. Mar. Sci. 7, 738. https://doi.org/10.3389/fmars.2020.569258 (2020).Article 

    Google Scholar 
    25.Comtois, S., Savenkoff, C., Bourassa, M. N., Brêthes, J. C. & Sears, R. Regional distribution and abundance of blue and humpback whales in the Gulf of St. Lawrence. Can. Tech. Rep. Fish. Aquat. Sci. 2877, viii+8 (2010).
    Google Scholar 
    26.Phillips, D. L. et al. Best practices for use of stable isotope mixing models in food-web studies. Can. J. Zool. 835, 823–835 (2014).Article 

    Google Scholar 
    27.Iverson, S. J., Field, C., Don Bowen, W. & Blanchard, W. Quantitative fatty acid signature analysis: A new method of estimating predator diets. Ecol. Monogr. 74, 211–235 (2004).Article 

    Google Scholar 
    28.Arregui, M., Borrell, A., Víkingsson, G., Ólafsdóttir, D. & Aguilar, A. Stable isotope analysis of fecal material provides insight into the diet of fin whales. Mar. Mamm. Sci. 34, 1059–1069 (2018).CAS 
    Article 

    Google Scholar 
    29.Bourdages, H. et al. Preliminary results from the ecosystemic survey in August 2019 in the Estuary and northern Gulf of St. Lawrence. Can. Sci. Adv. Sec. Res. Doc. 2020/009, iv+93 (2020).
    Google Scholar 
    30.Lesage, V., Lair, S., Turgeon, S. & Béland, P. Diet of St. Lawrence Estuary Beluga (Delphinapterus leucas) in a changing ecosystem. Can. Field Nat. 134, 21–35 (2020).Article 

    Google Scholar 
    31.Aubin, D. S., Smith, T. G. & Geraci, J. R. Seasonal epidermal molt in beluga whales, Delphinapterus leucas. Can. J. Zool. 68, 359–367 (1990).Article 

    Google Scholar 
    32.Busquets-Vass, G. et al. Estimating blue whale skin isotopic incorporation rates and baleen growth rates: Implications for assessing diet and movement patterns in mysticetes. PLoS ONE 12, e0177880. https://doi.org/10.1371/journal.pone.0177880 (2017).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    33.Vighi, M., Borrell, A. & Aguilar, A. Stable isotope analysis and fin whale subpopulation structure in the eastern North Atlantic. Mar. Mam. Sci. 32, 535–551 (2016).Article 

    Google Scholar 
    34.Ramp, C., Delarue, J., Bérubé, M., Hammond, P. S. & Sears, R. Fin whale survival and abundance in the Gulf of St. Lawrence, Canada. Endanger. Species Res. 23, 125–132 (2014).Article 

    Google Scholar 
    35.Lesage, V., Hammill, M. O. & Kovacs, K. M. Marine mammals and the community structure of the Estuary and Gulf of St. Lawrence, Canada: Evidence from stable isotope analysis. Mar. Ecol. Prog. Ser. 210, 203–221. https://doi.org/10.3354/meps210203 (2001).ADS 
    CAS 
    Article 

    Google Scholar 
    36.Ryan, C. et al. Prey preferences of sympatric fin (Balaenoptera physalus) and humpback (Megaptera novaeangliae) whales revealed by stable isotope mixing models. Mar. Mamm. Sci. 30, 242–258 (2014).CAS 
    Article 

    Google Scholar 
    37.Flinn, R. D., Trites, A. W., Gregr, E. J. & Perry, R. I. Diets of fin, sei, and sperm whales in British Columbia: An analysis of commercial whaling records, 1963–1967. Mar. Mamm. Sci. 18, 663–679 (2002).Article 

    Google Scholar 
    38.Bentaleb, I. et al. Foraging ecology of Mediterranean fin whales in a changing environment elucidated by satellite tracking and baleen plate stable isotopes. Mar. Ecol. Prog. Ser. 438, 285–302 (2011).ADS 
    Article 

    Google Scholar 
    39.Bolnick, D. I. & Fitzpatrick, B. M. Sympatric speciation: Models and empirical evidence. Annu. Rev. Ecol. Evol. Syst. 38, 459–487 (2007).Article 

    Google Scholar 
    40.Goldbogen, J. A. et al. Underwater acrobatics by the world’s largest predator: 360 rolling manoeuvres by lunge-feeding blue whales. Biol. Lett. 9, 20120986. https://doi.org/10.1098/rsbl.2012.0986 (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    41.Abrahms, B. et al. Memory and resource tracking drive blue whale migrations. Proc. Natl. Acad. Sci. 116, 5582–5587 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    42.Lesmerises, F., Johnson, C. J. & St-Laurent, M. H. Landscape knowledge is an important driver of the fission dynamics of an alpine ungulate. Anim. Behav. 140, 39–47 (2018).Article 

    Google Scholar 
    43.Serres, A. & Delfour, F. Social behaviors modulate bottlenose dolphins (Tursiops truncatus) breathing rate. Anim. Cogn. 6, 127–140 (2019).
    Google Scholar 
    44.Goldbogen, J. A. et al. Kinematics of foraging dives and lunge-feeding in fin whales. J. Exp. Biol. 209, 1231–1244 (2006).PubMed 
    Article 

    Google Scholar 
    45.Potvin, J., Goldbogen, J. A. & Shadwick, R. E. Metabolic expenditures of lunge feeding rorquals across scale: Implications for the evolution of filter feeding and the limits to maximum body size. PLoS ONE 7, e44854. https://doi.org/10.1371/journal.pone.0044854 (2012).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    46.Sexton, J. P., McIntyre, P. J., Angert, A. L. & Rice, K. J. Evolution and ecology of species range limits. Ann. Rev. Ecol. Evol. Syst. 40, 415–436 (2009).Article 

    Google Scholar 
    47.Schleimer, A. et al. Decline in abundance and apparent survival rates of fin whales (Balaenoptera physalus) in the northern Gulf of St. Lawrence. Ecol. Evol. 9, 4231–4244 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    48.Lesage, V. et al. Stable isotopes and trace elements as indicators of diet and habitat use in cetaceans: predicting errors related to preservation, lipid extraction, and lipid normalization. Mar. Ecol. Prog. Ser. 419, 249–265 (2010).ADS 
    Article 

    Google Scholar 
    49.Plourde, S., Winkler, G., Joly, P., St-Pierre, J. F. & Starr, M. Long-term seasonal and interannual variations of krill spawning in the lower St. Lawrence estuary, Canada, 1979–2009. J. Plankton Res. 33, 703–714 (2011).Article 

    Google Scholar 
    50.Plourde, S. et al. Daytime depth and thermal habitat of two sympatric krill species in response to surface salinity variability in the Gulf of St Lawrence, eastern Canada. ICES J. Mar. Sci. 71, 272–281 (2014).Article 

    Google Scholar 
    51.Cabrol, J. et al. Seasonal and large-scale spatial variability of the energy reserves and the feeding selectivity of Meganyctiphanes norvegica and Thysanoessa inermis in a Subarctic environment. Prog. Oceanogr. 179, 102203 (2019).Article 

    Google Scholar 
    52.Cabrol, J. et al. Functional feeding response of Nordic and Arctic krill on natural phytoplankton and zooplankton. J. Plankton Res. 42, 239–252 (2020).Article 

    Google Scholar 
    53.Cabrol, J. et al. Trophic niche partitioning of dominant North-Atlantic krill species, Meganyctiphanes norvegica, Thysanoessa inermis, and T. raschii. Limnol. Oceanogr. 64, 165–181 (2019).ADS 
    Article 

    Google Scholar 
    54.Guilpin, M. et al. Repeated vessel interactions and climate-or fishery-driven changes in prey density limit energy acquisition by foraging blue whales. Front. Mar. Sci. 7, 626. https://doi.org/10.3389/fmars.2020.00626 (2020).Article 

    Google Scholar 
    55.Guilpin, M. Étude des interactions bioénergétiques entre le rorqual bleu Balaenoptera musculus et le krill dans l’estuaire et le golfe du Saint-Laurent. Doctoral Thesis. (Université du Québec à Rimouski, 2020).56.MacArthur, R. H. & Pianka, E. R. On optimal use of a patchy environment. Am. Nat. 100, 603–609 (1966).Article 

    Google Scholar 
    57.Schleimer, A. et al. Spatio-temporal patterns in fin whale Balaenoptera physalus habitat use in the northern Gulf of St. Lawrence. Mar. Ecol. Prog. Ser. 623, 221–234 (2019).ADS 
    Article 

    Google Scholar 
    58.Bernier-Graveline, A. et al. Lipid metabolites as indicators of body condition in highly contaminant-exposed belugas from the endangered St. Lawrence Estuary population (Canada). Environ. Res. 192, 110272. https://doi.org/10.1016/j.envres.2020.110272 (2020).CAS 
    Article 
    PubMed 

    Google Scholar 
    59.Kershaw, J. L. et al. Declining reproductive success in the Gulf of St Lawrence’s humpback whales (Megaptera novaeangliae) reflects ecosystem shifts on their feeding grounds. Glob. Change Biol. 10, 1–15. https://doi.org/10.1111/gcb.15466 (2020).Article 

    Google Scholar 
    60.Mosnier, A. & Gosselin, J.-F. Seasonal distribution and concentration of four baleen whale species in the St. Lawrence Estuary based on 22 years of DFO observation data. Can. Sci. Adv. Sec. Res. Doc. 2020/053, iv+119 (2020).
    Google Scholar 
    61.Gowans, S., Simard, P., Giard, J., Vashro, C. & Sears, R. Photographic identification of fin whales (Balaenoptera physalus) off the Atlantic coast of Nova Scotia, Canada. Mar. Mamm. Sci. 21, 323–326 (2005).Article 

    Google Scholar 
    62.Bérubé, M. & Palsbøll, P. Identification of sex in cetaceans by multiplexing with three ZFX and ZFY specific primers. Mol. Ecol. 5, 283–287 (1996).PubMed 
    Article 

    Google Scholar 
    63.Williams, R. et al. Evidence for density-dependent changes in body condition and pregnancy rate of North Atlantic fin whales over four decades of varying environmental conditions. ICES J. Mar. Sci. 70, 1273–1280 (2013).Article 

    Google Scholar 
    64.Lockyer, C. Body fat condition in Northeast Atlantic fin whales, Balaenoptera physalus, and its relationship with reproduction and food resource. Can. J. Fish Aquat. Sci. 43, 142–147 (1986).Article 

    Google Scholar 
    65.Ryan, C. et al. Accounting for the effects of lipids in stable isotope (δ13C and δ15N values) analysis of skin and blubber of balaenopterid whales. Rapid Commun. Mass Spectrom. 26, 2745–2754 (2012).CAS 
    PubMed 
    Article 

    Google Scholar 
    66.Elliott, K. H., Roth, J. D. & Crook, K. Lipid extraction techniques for stable isotope analysis and ecological assays. Method Mol. Biol. 1609, 9–24 (2017).CAS 
    Article 

    Google Scholar 
    67.Newsome, S. D., Chivers, S. J. & Berman Kowalewski, M. The influence of lipid-extraction and long-term DMSO preservation on carbon (δ13C) and nitrogen (δ15N) isotope values in cetacean skin. Mar. Mamm. Sci. 34, 277–293 (2018).CAS 
    Article 

    Google Scholar 
    68.Budge, S. M., Iverson, S. J. & Koopman, H. N. Studying trophic ecology in marine ecosystems using fatty acids: A primer on analysis and interpretation. Mar. Mamm. Sci. 22, 759–801 (2006).Article 

    Google Scholar 
    69.Iverson, S. J., Frost, K. J. & Lang, S. Fat content and fatty acid composition of forage fish and invertebrates in Prince William Sound, Alaska: Factors contributing to among and within species variability. Mar. Ecol. Prog. Ser. 241, 161–181 (2002).ADS 
    CAS 
    Article 

    Google Scholar 
    70.Lesage, V. Trends in the trophic ecology of St. Lawrence beluga (Delphinapterus leucas) over the period 1988–2012, based on stable isotope analysis. Can. Sci. Adv. Sec. Res. Doc. 2013/126, iv+26 (2014).
    Google Scholar 
    71.Anderson, M. J. & Walsh, D. C. PERMANOVA, ANOSIM, and the Mantel test in the face of heterogeneous dispersions: What null hypothesis are you testing?. Ecol. Monogr. 83, 557–574 (2013).Article 

    Google Scholar 
    72.Clarke, K. R. & Gorley, R. N. PRIMER v7: User Manual/Tutorial 3rd edn. (Primer-E Ltd, 2015).
    Google Scholar 
    73.Bates, D. M., Mäechler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Soft. 67, 1–48 (2015).Article 

    Google Scholar 
    74.Wood, S. Mixed GAM computation vehicle with GCV/AIC/REML smoothness estimation and GAMMs by REML/PQL. in R Package Version 1–8 (2018).75.R Core Team. R: A Language and Environment for Statistical Computing. https://www.R-project.org (R Foundation for Statistical Computing, 2021).76.Stock, B. C. et al. Analyzing mixing systems using a new generation of Bayesian tracer mixing models. PeerJ 6, e5096. https://doi.org/10.7287/peerj.preprints.26884v1 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    77.Parnell, A., Inger, R., Bearhop, S. & Jackson, A. Source partitioning using stable isotopes: coping with too much variation. PLoS Biol. 5, e9672. https://doi.org/10.1371/journal.pone.0009672 (2010).ADS 
    CAS 
    Article 

    Google Scholar 
    78.Borrell, A., Abad-Oliva, N., Gómez-Campos, E., Giménez, J. & Aguilar, A. Discrimination of stable isotopes in fin whale tissues and application to diet assessment in cetaceans. Rapid Commun. Mass Spectrom. 26, 1596–1602 (2012).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    79.Gendron, D., Aguíñiga, S. & Carriquiry, J. D. d15N and d13C in skin biopsy samples: A note on their applicability for examining the relative trophic level in three rorqual species. J. Cetacean Res. Manag. 3, 41–44 (2001).
    Google Scholar 
    80.Giménez, J., Ramírez, F., Almunia, J., Forero, M. G. & de Stephanis, R. From the pool to the sea: Applicable isotope turnover rates and diet to skin discrimination factors for bottlenose dolphins (Tursiops truncatus). J. Exp. Mar. Biol. Ecol. 475, 54–61 (2016).Article 
    CAS 

    Google Scholar 
    81.Caut, S., Angulo, E. & Courchamp, F. Discrimination factors (Δ15N and Δ13C) in an omnivorous consumer: Effect of diet isotopic ratio. Funct. Ecol. 22, 255–263 (2008).Article 

    Google Scholar 
    82.Stock, B. C. & Semmens, B. X. Unifying error structures in commonly used biotracer mixing models. Ecology 97, 2562–2569 (2016).PubMed 
    Article 

    Google Scholar 
    83.Newsome, S. D., Yeakel, J. D., Wheatley, P. V. & Tinker, M. T. Tools for quantifying isotopic niche space and dietary variation at the individual and population level. J. Mammal. 93, 329–341 (2012).Article 

    Google Scholar 
    84.Nifong, J. C., Layman, C. A. & Silliman, B. R. Size, sex and individual-level behavior drive intrapopulation variation in cross-ecosystem foraging of a top-predator. J. Anim. Ecol. 84, 35–48 (2015).PubMed 
    Article 

    Google Scholar 
    85.Devries, M. S., Stock, B. C., Christy, J. H., Goldsmith, G. R. & Dawson, T. E. Specialized morphology corresponds to a generalist diet: Linking form and function in smashing mantis shrimp crustaceans. Oecologia 182, 429–442 (2016).ADS 
    PubMed 
    Article 

    Google Scholar  More

  • in

    Genetic engineering of marine cyanophages reveals integration but not lysogeny in T7-like cyanophages

    1.Flombaum P, Gallegos JL, Gordillo RA, Rincón J, Zabala LL, Jiao N, et al. Present and future global distributions of the marine cyanobacteria Prochlorococcus and Synechococcus. Proc Natl Acad Sci USA. 2013;110:9824–9.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    2.Goldin S, Hulata Y, Baran N, Lindell D. Quantification of T4-like and T7-like cyanophages using the polony method show they are significant members of the virioplankton in the North Pacific Subtropical Gyre. Front Microbiol. 2020;11:1210.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    3.Waterbury JB, Valois FW. Resistance to co-occurring phages enables marine Synechococcus communities to coexist with cyanophages abundant in seawater. Appl Environ Microbiol. 1993;59:3393–9.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    4.Marston MF, Sallee JL. Genetic diversity and temporal variation in the cyanophage community infecting marine Synechococcus species in Rhode Island’s coastal waters. Appl Environ Microbiol. 2003;69:4639–47.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    5.Clokie MRJ, Mann NH. Marine cyanophages and light. Environ Microbiol. 2006;8:2074–82.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    6.Avrani S, Wurtzel O, Sharon I, Sorek R, Lindell D. Genomic island variability facilitates Prochlorococcus -virus coexistence. Nature. 2011;474:604–8.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    7.Marston MF, Pierciey FJ, Shepard A, Gearin G, Qi J, Yandava C, et al. Rapid diversification of coevolving marine Synechococcus and a virus. Proc Natl Acad Sci USA. 2012;109:4544–9.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    8.Fuhrman JA. Marine virueses and their biogeochemical and ecological effects. Nature. 1999;399:541–8.9.Suttle CA. Marine viruses – major players in the global ecosystem. Nat Rev Microbiol. 2007;5:801–12.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    10.Breitbart M, Bonnain C, Malki K, Sawaya NA. Phage puppet masters of the marine microbial realm. Nat Microbiol. 2018;3:754–66.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    11.Suttle CA, Chan AM. Marine cyanophages infecting oceanic and coastal strains of Synechococcus: abundance, morphology, cross-infectivity and growth characteristics. Mar Ecol Prog Ser. 1993;92:99–109.Article 

    Google Scholar 
    12.Sullivan MB, Waterbury JB, Chisholm SW. Cyanophages infecting the oceanic cyanobacterium Prochlorococcus. Nature. 2003;424:1047–51.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    13.Sullivan MB, Coleman ML, Weigele P, Rohwer F, Chisholm SW. Three Prochlorococcus cyanophage genomes: Signature features and ecological interpretations. PLoS Biol. 2005;3:0790–806.CAS 
    Article 

    Google Scholar 
    14.Pope WH, Weigele PR, Chang J, Pedulla ML, Ford ME, Houtz JM, et al. Genome sequence, structural proteins, and capsid organization of the cyanophage Syn5: a “horned” bacteriophage of marine Synechococcus. J Mol Biol. 2007;368:966–81.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    15.Weigele PR, Pope WH, Pedulla ML, Houtz JM, Smith AL, Conway JF, et al. Genomic and structural analysis of Syn9, a cyanophage infecting marine Prochlorococcus and Synechococcus. Environ Microbiol. 2007;9:1675–95.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    16.Labrie SJ, Frois-Moniz K, Osburne MS, Kelly L, Roggensack SE, Sullivan MB, et al. Genomes of marine cyanopodoviruses reveal multiple origins of diversity. Environ Microbiol. 2013;15:1356–76.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    17.Huang S, Wang K, Jiao N, Chen F. Genome sequences of siphoviruses infecting marine Synechococcus unveil a diverse cyanophage group and extensive phage-host genetic exchanges. Environ Microbiol. 2012;14:540–58.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    18.Sullivan MB, Huang KH, Ignacio-Espinoza JC, Berlin AM, Kelly L, Weigele PR, et al. Genomic analysis of oceanic cyanobacterial myoviruses compared with T4-like myoviruses from diverse hosts and environments. Environ Microbiol. 2010;12:3035–56.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    19.Ghai R, Martin-Cuadrado AB, Molto AG, Heredia IG, Cabrera R, Martin J, et al. Metagenome of the Mediterranean deep chlorophyll maximum studied by direct and fosmid library 454 pyrosequencing. ISME J. 2010;4:1154–66.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    20.Ma Y, Allen LZ, Palenik B. Diversity and genome dynamics of marine cyanophages using metagenomic analyses. Environ Microbiol Rep. 2014;6:583–94.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Sabehi G, Shaulov L, Silver DH, Yanai I, Harel A, Lindell D. A novel lineage of myoviruses infecting cyanobacteria is widespread in the oceans. Proc Natl Acad Sci USA. 2012;109:2037–42.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    22.Huang S, Zhang S, Jiao N, Chen F. Comparative genomic and phylogenomic analyses reveal a conserved core genome shared by estuarine and oceanic cyanopodoviruses. PLoS One. 2015;10:1–17.
    Google Scholar 
    23.Ignacio-espinoza JC, Sullivan MB. Phylogenomics of T4 cyanophages: lateral gene transfer in the ‘core’ and origins of host genes. Environ Microbiol. 2012;14:2113–26.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    24.Crummett LT, Puxty RJ, Weihe C, Marston MF, Martiny JBHH. The genomic content and context of auxiliary metabolic genes in marine cyanomyoviruses. Virology. 2016;499:219–29.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    25.Millard AD, Zwirglmaier K, Downey MJ, Mann NH, Scanlan DJ. Comparative genomics of marine cyanomyoviruses reveals the widespread occurrence of Synechococcus host genes localized to a hyperplastic region: Implications for mechanisms of cyanophage evolution. Environ Microbiol. 2009;11:2370–87.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    26.Mann NH, Cook A, Millard A, Bailey S, Clokie M. Bacterial photosynthesis genes in a virus. Nature 2003;424:741–741.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    27.Lindell D, Sullivan MB, Johnson ZI, Tolonen AC, Rohwer F, Chisholm SW. Transfer of photosynthesis genes to and from Prochlorococcus viruses. Proc Natl Acad Sci USA. 2004;101:11013–8.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    28.Puxty RJ, Millard AD, Evans DJ, Scanlan DJ. Viruses inhibit CO2 fixation in the most abundant phototrophs on Earth. Curr Biol. 2016;26:1585–9.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    29.Kelly L, Ding H, Huang KH, Osburne MS, Chisholm SW. Genetic diversity in cultured and wild marine cyanomyoviruses reveals phosphorus stress as a strong selective agent. ISME J. 2013;7:1827–41.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    30.Yin Y, Fischer D. Identification and investigation of ORFans in the viral world. BMC Genomics. 2008;9:24.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    31.Clokie MR, Millard AD, Mann NH. T4 genes in the marine ecosystem: studies of the T4-like cyanophages and their role in marine ecology. Virol J. 2010;7:291.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    32.Rihtman B, Bowman‐Grahl S, Millard A, Corrigan RM, Clokie MRJ, Scanlan DJ. Cyanophage MazG is a pyrophosphohydrolase but unable to hydrolyse magic spot nucleotides. Environ Microbiol Rep. 2019;11:448–55.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    33.Dammeyer T, Bagby SC, Sullivan MB, Chisholm SW, Frankenberg-Dinkel N. Efficient phage-mediated pigment biosynthesis in oceanic cyanobacteria. Curr Biol. 2008;18:442–8.CAS 
    PubMed 
    Article 

    Google Scholar 
    34.Roitman S, Hornung E, Flores-Uribe J, Sharon I, Feussner I, Béjà O. Cyanophage-encoded lipid desaturases: Oceanic distribution, diversity and function. ISME J. 2018;12:343–55.CAS 
    PubMed 
    Article 

    Google Scholar 
    35.Thompson LR, Zeng Q, Kelly L, Huang KH, Singer AU, Stubbe J, et al. Phage auxiliary metabolic genes and the redirection of cyanobacterial host carbon metabolism. Proc Natl Acad Sci USA. 2011;108:E757–64.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    36.Puxty RJ, Perez-Sepulveda B, Rihtman B, Evans DJ, Millard AD, Scanlan DJ. Spontaneous deletion of an “ORFanage” region facilitates host adaptation in a “photosynthetic” cyanophage. PLoS One. 2015;10:e0132642.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    37.Howard-Varona C, Hargreaves KR, Abedon ST, Sullivan MB. Lysogeny in nature: mechanisms, impact and ecology of temperate phages. ISME J. 2017;11:1511–20.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    38.Ranade K, Poteete AR. Superinfection exclusion (sieB) genes of bacteriophages P22 and λ. J Bacteriol. 1993;175:4712–8.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    39.Fogg PCM, Allison HE, Saunders JR, McCarthy AJ. Bacteriophage Lambda: a paradigm revisited. J Virol. 2010;84:6876–9.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    40.van Houte S, Buckling A, Westra ER. Evolutionary ecology of prokaryotic immune mechanisms. Microbiol Mol Biol Rev. 2016;80:745–63.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    41.Tuttle MJ, Buchan A. Lysogeny in the oceans: lessons from cultivated model systems and a reanalysis of its prevalence. Environ Microbiol. 2020;22:4919–33.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.Knowles B, Silveira CB, Bailey BA, Barott K, Cantu VA, Cobian-Guëmes AG, et al. Lytic to temperate switching of viral communities. Nature. 2016;531:466–70.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    43.Touchon M, Bernheim A, Rocha EPC. Genetic and life-history traits associated with the distribution of prophages in bacteria. ISME J. 2016;10:2744–54.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    44.Roux S, Hallam SJ, Woyke T, Sullivan MB. Viral dark matter and virus–host interactions resolved from publicly available microbial genomes. Elife. 2015;4:e08490.PubMed Central 
    Article 
    PubMed 

    Google Scholar 
    45.Luo E, Eppley JM, Romano AE, Mende DR, DeLong EF. Double-stranded DNA. virioplankton dynamics and reproductive strategies in the oligotrophic open ocean water column. ISME J. 2020;14:1304–15.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    46.Touchon M, Moura de Sousa JA, Rocha EP. Embracing the enemy: the diversification of microbial gene repertoires by phage-mediated horizontal gene transfer. Curr Opin Microbiol. 2017;38:66–73.CAS 
    PubMed 
    Article 

    Google Scholar 
    47.Coleman ML, Sullivan MB, Martiny AC, Steglich C, Barry K, Delong EF, et al. Genomic islands and the ecology and evolution of Prochlorococcus. Science. 2006;311:1768–70.CAS 
    PubMed 
    Article 

    Google Scholar 
    48.Zhaxybayeva O, Gogarten JP, Charlebois RL, Doolittle WF, Papke RT. Phylogenetic analyses of cyanobacterial genomes: quantification of horizontal gene transfer events. Genome Res. 2006;16:1099–108.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    49.Lindell D, Jaffe JD, Coleman ML, Futschik ME, Axmann IM, Rector T, et al. Genome-wide expression dynamics of a marine virus and host reveal features of co-evolution. Nature. 2007;449:83–6.CAS 
    PubMed 
    Article 

    Google Scholar 
    50.Raytcheva DA, Haase-Pettingell C, Piret JM, King JA. Intracellular assembly of cyanophage Syn5 proceeds through a scaffold-containing procapsid. J Virol. 2011;85:2406–15.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    51.Flores-Uribe J, Philosof A, Sharon I, Fridman S, Larom S, Béjà O. A novel uncultured marine cyanophage lineage with lysogenic potential linked to a putative marine Synechococcus ‘relic’ prophage. Environ Microbiol Rep. 2019;11:598–604.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    52.Malmstrom RR, Rodrigue S, Huang KH, Kelly L, Kern SE, Thompson A, et al. Ecology of uncultured Prochlorococcus clades revealed through single-cell genomics and biogeographic analysis. ISME J. 2013;7:184–98.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    53.Kiro R, Shitrit D, Qimron U. Efficient engineering of a bacteriophage genome using the type I-E CRISPR-Cas system. RNA Biol. 2014;11:42–4.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    54.Sarkis GJ, Jacobs WR, Hatfull GF. L5 luciferase reporter mycobacteriophages: a sensitive tool for the detection and assay of live mycobacteria. Mol Microbiol. 1995;15:1055–67.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    55.Tanji Y, Furukawa C, Na SH, Hijikata T, Miyanaga K, Unno H. Escherichia coli detection by GFP-labeled lysozyme-inactivated T4 bacteriophage. J Biotechnol. 2004;114:11–20.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    56.Brahamsha B. A genetic manipulation system for oceanic cyanobacteria of the genus Synechococcus. Appl Environ Microbiol. 1996;62:1747–51.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    57.Mahichi F, Synnott AJ, Yamamichi K, Osada T, Tanji Y. Site-specific recombination of T2 phage using IP008 long tail fiber genes provides a targeted method for expanding host range while retaining lytic activity. FEMS Microbiol Lett. 2009;295:211–7.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    58.Le S, He X, Tan Y, Huang G, Zhang L, Lux R, et al. Mapping the tail fiber as the receptor binding protein responsible for differential host specificity of Pseudomonas aeruginosa bacteriophages PaP1 and JG004. PLoS One. 2013;8:e68562.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    59.Murphy KC. Phage recombinases and their applications. Adv Virus Res. 2012;83:367–414.60.Bujarski JJ. Recombination of viruses. In: Encyclopedia of Virology. Elsevier; 1999. p. 1446–53.61.Pires DP, Cleto S, Sillankorva S, Azeredo J, Lu TK. Genetically engineered phages: a review of advances over the last decade. Microbiol Mol Biol Rev. 2016;80:523–43.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    62.Qimron U, Marintcheva B, Tabor S, Richardson CC. Genomewide screens for Escherichia coli genes affecting growth of T7 bacteriophage. Proc Natl Acad Sci USA. 2006;103:19039–44.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    63.Selick HE, Kreuzer KN, Alberts BM. The bacteriophage T4 insertion/substitution vector system. A method for introducing site-specific mutations into the virus chromosome. J Biol Chem. 1988;263:11336–47.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    64.Martel B, Moineau S. CRISPR-Cas: an efficient tool for genome engineering of virulent bacteriophages. Nucleic Acids Res. 2014;42:9504–13.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    65.Dale JW, Greenaway PJ. Identification of recombinant phages by plaque hybridization. In: Walker JM, editor. Nucleic Acids. Totowa, NJ: Humana Press; 1984. p. 285–8.Chapter 

    Google Scholar 
    66.Dekel-Bird NP, Avrani S, Sabehi G, Pekarsky I, Marston MF, Kirzner S, et al. Diversity and evolutionary relationships of T7-like podoviruses infecting marine cyanobacteria. Environ Microbiol. 2013;15:1476–91.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    67.Doron S, Fedida A, Hernndez-Prieto MA, Sabehi G, Karunker I, Stazic D, et al. Transcriptome dynamics of a broad host-range cyanophage and its hosts. ISME J. 2016;10:1437–55.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    68.Krupovic M, Forterre P. Single-stranded DNA viruses employ a variety of mechanisms for integration into host genomes. Ann NY Acad Sci. 2015;1341:41–53.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    69.Morris RM, Cain KR, Hvorecny KL, Kollman JM. Lysogenic host–virus interactions in SAR11 marine bacteria. Nat Microbiol. 2020;5:1011–5.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    70.Martínez-García E, Jatsenko T, Kivisaar M, de Lorenzo V. Freeing Pseudomonas putida KT2440 of its proviral load strengthens endurance to environmental stresses. Environ Microbiol. 2015;17:76–90.PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    71.Mruwat N, Carlson MCG, Goldin S, Ribalet F, Kirzner S, Hulata Y, et al. A single-cell polony method reveals low levels of infected Prochlorococcus in oligotrophic waters despite high cyanophage abundances. ISME J. 2021;15:41–54.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    72.Oppenheim AB, Kobiler O, Stavans J, Court DL, Adhya S. Switches in bacteriophage lambda development. Annu Rev Genet. 2005;39:409–29.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    73.Ray U, Sakalka A. Lysogenization of Escherichia coli by bacteriophage Lambda: complementary activity of the host’s DNA polymerase I and ligase and bacteriophage replication proteins Q and P. J Virol. 1976;18:511–7.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    74.Greengrass E. Resistance of marine Synechococcus to podovirus infection: genetic basis and phenotypic characterization. M.Sc. thesis. Technion – Israel Inst Technol. 2013.75.Fedida A, Lindell D. Two Synechococcus genes, two different effects on cyanophage infection. Viruses. 2017;9:136.PubMed Central 
    Article 
    CAS 

    Google Scholar 
    76.Shao Q, Trinh JT, McIntosh CS, Christenson B, Balázsi G, Zeng L. Lysis-lysogeny coexistence: prophage integration during lytic development. Microbiol Open. 2017;6:e00395.Article 
    CAS 

    Google Scholar 
    77.Chen F, Lu J. Genomic sequence and evolution of marine cyanophage P60: a new insight on lytic and lysogenic phages. Appl Environ Microbiol. 2002;68:2589–94.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    78.Brum JR, Hurwitz BL, Schofield O, Ducklow HW, Sullivan MB. Seasonal time bombs: dominant temperate viruses affect Southern Ocean microbial dynamics. ISME J. 2016;10:437–49.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    79.Pratama AA, van Elsas JD. The ‘neglected’ soil virome—potential role and impact. Trends Microbiol. 2018;26:649–62.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    80.Srinivasiah S, Bhavsar J, Thapar K, Liles M, Schoenfeld T, Wommack KE. Phages across the biosphere: contrasts of viruses in soil and aquatic environments. Res Microbiol. 2008;159:349–57.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    81.Zhao Y, Temperton B, Thrash JC, Schwalbach MS, Vergin KL, Landry ZC, et al. Abundant SAR11 viruses in the ocean. Nature. 2013;494:357–60.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    82.Zhao Y, Qin F, Zhang R, Giovannoni SJ, Zhang Z, Sun J, et al. Pelagiphages in the Podoviridae family integrate into host genomes. Environ Microbiol. 2018;21:1989–2001.PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    83.Kashtan N, Roggensack SE, Berta-Thompson JW, Grinberg M, Stepanauskas R, Chisholm SW. Fundamental differences in diversity and genomic population structure between Atlantic and Pacific Prochlorococcus. ISME J. 2017;11:1997–2011.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    84.Berube PM, Biller SJ, Hackl T, Hogle SL, Satinsky BM, Becker JW, et al. Single cell genomes of Prochlorococcus, Synechococcus, and sympatric microbes from diverse marine environments. Sci Data. 2018;5:1–11.Article 
    CAS 

    Google Scholar 
    85.Wyman M, Gregory RPF, Carr NG. Novel role for phycoerythrin in a marine cyanobacterium, Synechococcus strain DC2. Science. 1985;230:818–20.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    86.Lindell D, Padan E, Post AF. Regulation of ntcA expression and nitrite uptake in the marine Synechococcus sp. strain WH 7803. J Bacteriol. 1998;180:1878–86.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    87.Moore LR, Coe A, Zinser ER, Saito MA, Sullivan MB, Lindell D, et al. Culturing the marine cyanobacterium Prochlorococcus. Limnol Oceanogr Methods. 2007;5:353–62.CAS 
    Article 

    Google Scholar 
    88.Lindell D. The genus Prochlorococcus, phylum cyanobacteria. In: The Prokaryotes: Other Major Lineages of Bacteria and The Archaea. Springer-Verlag Berlin Heidelberg; 2014. p. 829–45.89.Morris JJ, Kirkegaard R, Szul MJ, Johnson ZI, Zinser ER. Facilitation of robust growth of Prochlorococcus colonies and dilute liquid cultures by “helper” heterotrophic bacteria. Appl Environ Microbiol. 2008;74:4530–4.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    90.Wolk CP, Fan Q, Zhou R, Huang G, Lechno-Yossef S, Kuritz T, et al. Paired cloning vectors for complementation of mutations in the cyanobacterium Anabaena sp. strain PCC 7120. Arch Microbiol. 2007;188:551–63.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    91.Bryksin A, Matsumura I. Overlap extension PCR cloning: a simple and reliable way to create recombinant plasmids. Biotechniques. 2010;48:463–5.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    92.Simon R, Priefer U, Pühler A. A broad host range mobilization system for in vivo genetic engineering: Transposon mutagenesis in gram negative bacteria. Nat Biotechnol. 1983;1:784–91.CAS 
    Article 

    Google Scholar 
    93.Henn MR, Sullivan MB, Stange-Thomann N, Osburne MS, Berlin AM, Kelly L, et al. Analysis of high-throughput sequencing and annotation strategies for phage genomes. PLoS One. 2010;5:e9083.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    94.Zinser ER, Coe A, Johnson ZI, Martiny AC, Fuller NJ, Scanlan DJ, et al. Prochlorococcus ecotype abundances in the North Atlantic Ocean as revealed by an improved quantitative PCR method. Appl Environ Microbiol. 2006;72:723–32.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    95.Lindell D, Jaffe JD, Johnson ZI, Church GM, Chisholm SW. Photosynthesis genes in marine viruses yield proteins during host infection. Nature. 2005;438:86–9.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    96.Mitra RD, Church GM. In situ localized amplification and contact replication of many individual DNA molecules. Nucleic Acids Res. 1999;27:e34.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    97.Schwartz DA, Lindell D. Genetic hurdles limit the arms race between Prochlorococcus and the T7-like podoviruses infecting them. ISME J. 2017;11:1836–51.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    98.Bull JJ, Badgett MR, Wichman HA, Huelsenbeck JP, Hillis DM, Gulati A, et al. Exceptional convergent evolution in a virus. Genetics. 1997;147:1497–507.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    99.Quinlan AR. BEDTools: the Swiss-army Tool for genome feature analysis. Curr Protoc Bioinform. 2014;47:1–34. 11.12.Article 

    Google Scholar 
    100.Biller SJ, Berube PM, Dooley K, Williams M, Satinsky BM, Hackl T, et al. Marine microbial metagenomes sampled across space and time. Sci Data. 2018;5:180176.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    101.Haro-Moreno JM, López-Pérez M, de la Torre JR, Picazo A, Camacho A, Rodriguez-Valera F. Fine metagenomic profile of the Mediterranean stratified and mixed water columns revealed by assembly and recruitment. Microbiome. 2018;6:128.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    102.Philosof A, Yutin N, Flores-Uribe J, Sharon I, Koonin EV, Béjà O. Novel abundant oceanic viruses of uncultured marine group II Euryarchaeota. Curr Biol. 2017;27:1362–8.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    103.Wilson ST, Aylward FO, Ribalet F, Barone B, Casey JR, Connell PE, et al. Coordinated regulation of growth, activity and transcription in natural populations of the unicellular nitrogen-fixing cyanobacterium Crocosphaera. Nat Microbiol. 2017;2:17118.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    104.Sunagawa S, Coelho LP, Chaffron S, Kultima JR, Labadie K, Salazar G, et al. Structure and function of the global ocean microbiome. Science. 2015;348:1261359.PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    105.Bushnell B. BBMap: a fast, accurate, splice-aware aligner. 2014;URL: https://www.osti.gov/servlets/purl/1241166.106.Li H. Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. arXiv. 2013;1303:3997.
    Google Scholar 
    107.Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. The sequence alignment/map format and SAMtools. Bioinformatics. 2009;25:2078–9.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar  More

  • in

    Wolves, dogs and humans in regular contact can mutually impact each other’s skin microbiota

    Humans have the least and pet dogs have the most diverse skin microbiotaOut of all four groups, species richness and diversity were lowest in human skin microbiota, whereas the pet dog group had the highest species richness and diversity. All four groups differed significantly to each other (Kruskal–Wallis; Chao1, chi-squared = 20.828, df = 3, p  More

  • in

    Nonlinear shifts in infectious rust disease due to climate change

    1.Harvell, C. D. et al. Climate warming and disease risks for terrestrial and marine Biota. Science 296, 2158–2162 (2002).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    2.Gautam, H. R., Bhardwaj, M. L. & Kumar, R. Climate change and its impact on plant diseases. Curr. Sci. 105, 1685–1691 (2013).
    Google Scholar 
    3.Bebber, D. P. & Gurr, S. J. Crop-destroying fungal and oomycete pathogens challenge food security. Fungal Genet. Biol. 74, 62–64 (2015).PubMed 
    Article 

    Google Scholar 
    4.Lukanda, M. et al. First report of maize chlorotic mottle virus infecting maize in the Democratic Republic of the Congo. Plant Dis. 98, 1448–1448 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    5.Brasier, C. M. in The Elms: Breeding,Conservation, and Disease Management (ed. Dunn, C. P.) 61–72 (Springer US, 2000). https://doi.org/10.1007/978-1-4615-4507-1_4.6.Boyd, I. L., Freer-Smith, P. H., Gilligan, C. A. & Godfray, H. C. J. The consequence of tree pests and diseases for ecosystem services. Science 342, 1235773 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    7.Chaloner, T. M., Gurr, S. J. & Bebber, D. P. Geometry and evolution of the ecological niche in plant-associated microbes. Nat. Commun. 11, 2955 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    8.Donald, F., Green, S., Searle, K., Cunniffe, N. J. & Purse, B. V. Small scale variability in soil moisture drives infection of vulnerable juniper populations by invasive forest pathogen. Ecol. Manag. 473, 118324 (2020).Article 

    Google Scholar 
    9.Sturrock, R. N. et al. Climate change and forest diseases. Plant Pathol. 60, 133–149 (2011).Article 

    Google Scholar 
    10.Pathak, R., Singh, S. K., Tak, A. & Gehlot, P. Impact of climate change on host, pathogen and plant disease adaptation regime: a review. Biosci. Biotechnol. Res. Asia 15, 529–540 (2018).Article 

    Google Scholar 
    11.Lafferty, K. D. The ecology of climate change and infectious diseases. Ecology 90, 888–900 (2009).PubMed 
    Article 

    Google Scholar 
    12.Ghelardini, L., Pepori, A. L., Luchi, N., Capretti, P. & Santini, A. Drivers of emerging fungal diseases of forest trees. Ecol. Manag. 381, 235–246 (2016).Article 

    Google Scholar 
    13.Wyka, S. A. et al. Emergence of white pine needle damage in the northeastern United States is associated with changes in pathogen pressure in response to climate change. Glob. Change Biol. 23, 394–405 (2017).ADS 
    Article 

    Google Scholar 
    14.Garrett, K. A. et al. in Climate Change 2nd edn (ed. Letcher, T. M.) 325–338 (Elsevier, 2016). https://doi.org/10.1016/B978-0-444-63524-2.00021-X.15.Bebber, D. P. Range-expanding pests and pathogens in a warming world. Annu. Rev. Phytopathol. 53, 335–356 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    16.Bebber, D. P., Ramotowski, M. A. T. & Gurr, S. J. Crop pests and pathogens move polewards in a warming world. Nat. Clim. Change 3, 985–988 (2013).ADS 
    Article 

    Google Scholar 
    17.Altizer, S., Ostfeld, R. S., Johnson, P. T. J., Kutz, S. & Harvell, C. D. Climate change and infectious diseases: from evidence to a predictive framework. Science 341, 514–519 (2013).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    18.Mordecai, E. A. et al. Optimal temperature for malaria transmission is dramatically lower than previously predicted. Ecol. Lett. 16, 22–30 (2013).PubMed 
    Article 

    Google Scholar 
    19.Huey, R. B. & Berrigan, D. Temperature, demography, and ectotherm fitness. Am. Nat. 158, 204–210 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    20.Huey, R. B. & Stevenson, R. D. Integrating thermal physiology and ecology of ectotherms: a discussion of approaches. Integr. Comp. Biol. 19, 357–366 (1979).
    Google Scholar 
    21.Rohr, J. R. et al. Frontiers in climate change—disease research. Trends Ecol. Evol. 26, 270–277 (2011).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    22.Peterson, A. T. Shifting suitability for malaria vectors across Africa with warming climates. BMC Infect. Dis. 9, 59 (2009).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    23.Garamszegi, L. Z. Climate change increases the risk of malaria in birds. Glob. Change Biol. 17, 1751–1759 (2011).ADS 
    Article 

    Google Scholar 
    24.Cook, B. I., Mankin, J. S. & Anchukaitis, K. J. Climate change and drought: from past to future. Curr. Clim. Change Rep. 4, 164–179 (2018).Article 

    Google Scholar 
    25.Desprez-Loustau, M.-L., Marçais, B., Nageleisen, L.-M., Piou, D. & Vannini, A. Interactive effects of drought and pathogens in forest trees. Ann. Sci. 63, 597–612 (2006).Article 

    Google Scholar 
    26.Brodribb, T. J. & McAdam, S. A. M. Passive origins of stomatal control in vascular plants. Science 331, 582–585 (2011).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    27.Jactel, H. et al. Drought effects on damage by forest insects and pathogens: a meta-analysis. Glob. Change Biol. 18, 267–276 (2012).ADS 
    Article 

    Google Scholar 
    28.Baptista-Rosas, R. C. et al. Molecular detection of Coccidioides spp. from environmental samples in Baja California: linking Valley Fever to soil and climate conditions. Fungal Ecol. 5, 177–190 (2012).Article 

    Google Scholar 
    29.Cohen, J. M. et al. The thermal mismatch hypothesis explains host susceptibility to an emerging infectious disease. Ecol. Lett. 20, 184–193 (2017).PubMed 
    Article 

    Google Scholar 
    30.Mcelrone, A. J., Reid, C. D., Hoye, K. A., Hart, E. & Jackson, R. B. Elevated CO2 reduces disease incidence and severity of a red maple fungal pathogen via changes in host physiology and leaf chemistry. Glob. Change Biol. 11, 1828–1836 (2005).ADS 
    Article 

    Google Scholar 
    31.Berzitis, E. A., Minigan, J. N., Hallett, R. H. & Newman, J. A. Climate and host plant availability impact the future distribution of the bean leaf beetle (Cerotoma trifurcata). Glob. Change Biol. 20, 2778–2792 (2014).ADS 
    Article 

    Google Scholar 
    32.Bebber, D. P. & Gurr, S. J. Biotic interactions and climate in species distribution modelling. bioRxiv 520320 https://doi.org/10.1101/520320 (2019).33.Parker, I. M. et al. Phylogenetic structure and host abundance drive disease pressure in communities. Nature 520, 542–544 (2015).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    34.Morgan, E. R., Milner-Gulland, E. J., Torgerson, P. R. & Medley, G. F. Ruminating on complexity: macroparasites of wildlife and livestock. Trends Ecol. Evol. 19, 181–188 (2004).PubMed 
    Article 

    Google Scholar 
    35.Paull, S. H., LaFonte, B. E. & Johnson, P. T. J. Temperature-driven shifts in a host-parasite interaction drive nonlinear changes in disease risk. Glob. Change Biol. 18, 3558–3567 (2012).ADS 
    Article 

    Google Scholar 
    36.Pearson, R. G. & Dawson, T. P. Predicting the impacts of climate change on the distribution of species: are bioclimate envelope models useful? Glob. Ecol. Biogeogr. 12, 361–371 (2003).Article 

    Google Scholar 
    37.Bebber, D. P. Climate change effects on Black Sigatoka disease of banana. Philos. Trans. R. Soc. B: Biol. Sci. 374, 20180269 (2019).Article 

    Google Scholar 
    38.Soberón, J. & Peterson, A. T. Interpretation of models of fundamental ecological niches and species’ distributional areas. https://doi.org/10.17161/bi.v2i0.4 (2005).39.Garrett, K. A. et al. Complexity in climate-change impacts: an analytical framework for effects mediated by plant disease. Plant Pathol. 60, 15–30 (2011).Article 

    Google Scholar 
    40.Scherm, H. Climate change: can we predict the impacts on plant pathology and pest management? Can. J. Plant Pathol. 26, 267–273 (2004).Article 

    Google Scholar 
    41.Simler-Williamson, A. B., Rizzo, D. M. & Cobb, R. C. Interacting effects of global change on forest pest and pathogen dynamics. Annu. Rev. Ecol. Evol. Syst. https://doi.org/10.1146/annurev-ecolsys-110218-024934 (2019).42.Campbell, E. M. & Antos, J. A. Distribution and severity of white pine blister rust and mountain pine beetle on whitebark pine in British Columbia. Can. J. Res. 30, 1051–1059 (2000).Article 

    Google Scholar 
    43.Larsen, A. E., Meng, K. & Kendall, B. E. Causal analysis in control–impact ecological studies with observational data. Methods Ecol. Evol. 10, 924–934 (2019).Article 

    Google Scholar 
    44.McDonald, G. I., Richardson, B. A., Zambino, P. J., Klopfenstein, N. B. & Kim, M.-S. Pedicularis and Castilleja are natural hosts of Cronartium ribicola in North America: a first report. Pathol. 36, 73–82 (2006).Article 

    Google Scholar 
    45.Geils, B. W., Hummer, K. E. & Hunt, R. S. White pines, Ribes, and blister rust: a review and synthesis. Pathol. 40, 147–185 (2010).Article 

    Google Scholar 
    46.Kinloch, B. B. White pine blister rust in North America: past and prognosis. Phytopathology 93, 1044–1047 (2003).PubMed 
    Article 

    Google Scholar 
    47.Arsdel, E. P. V., Geils, B. W. & Zambino, P. J. Epidemiology for hazard rating of white pine blister rust. In: Guyon JC Comp Proc. 53rd Western International Forest Disease Work Conference 2005 September 26–30 Jackson WY USA (Department of Agriculture, Forest Service, Intermountain Region, Ogden UT, 2006).48.Dudney, J. Characterizing and Managing Drivers of Change in Mediterranean Forest and Grassland Communities (UC Berkeley, 2019).49.Kreyling, J. et al. To replicate, or not to replicate—that is the question: how to tackle nonlinear responses in ecological experiments. Ecol. Lett. 21, 1629–1638 (2018).PubMed 
    Article 

    Google Scholar 
    50.Larson, E. R. & Kipfmueller, K. F. Ecological disaster or the limits of observation? reconciling modern declines with the long-term dynamics of whitebark pine communities. Geogr. Compass 6, 189–214 (2012).Article 

    Google Scholar 
    51.Parmesan, C. & Yohe, G. A globally coherent fingerprint of climate change impacts across natural systems. Nature 421, 37–42 (2003).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    52.Kinloch, B. B. et al. Patterns of variation in blister rust resistance in sugar pine (Pinus lambertiana). In: Proc. IUFRO joint conference: Genetics of five-needle pines, rusts of forest trees, and Strobusphere; 2014 June 15–20; Fort Collins, CO. Proc. RMRS-P-76 (eds Schoettle, A. W., Sniezko, R. A. & Kliejunas, J. T.) 124–128 (Department of Agriculture, Forest Service, Rocky Mountain Research Station, 2018).53.King, J. N., David, A., Noshad, D. & Smith, J. A review of genetic approaches to the management of blister rust in white pines. Pathol. 40, 292–313 (2010).Article 

    Google Scholar 
    54.Maloney, P. E. Incidence and distribution of white pine blister rust in the high-elevation forests of California. Forest Pathol. 41, 308–316 (2011).Article 

    Google Scholar 
    55.Dunlap, J. M. Variability in and environmental correlates to white pine blister rust incidence in five California white pine species. Northwest Sci. 86, 248–263 (2012).Article 

    Google Scholar 
    56.Thoma, D. P., Shanahan, E. K. & Irvine, K. M. Climatic correlates of white pine blister rust infection in whitebark pine in the greater yellowstone ecosystem. Forests 10, 666 (2019).Article 

    Google Scholar 
    57.Talley, S. M., Coley, P. D. & Kursar, T. A. The effects of weather on fungal abundance and richness among 25 communities in the Intermountain West. BMC Ecol. 2, 7 (2002).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    58.Davis, J. K. et al. Improving the prediction of arbovirus outbreaks: A comparison of climate-driven models for West Nile virus in an endemic region of the United States. Acta Trop. 185, 242–250 (2018).PubMed 
    Article 

    Google Scholar 
    59.Manstretta, V. & Rossi, V. Effects of weather variables on ascospore discharge from Fusarium graminearum Perithecia. PLoS ONE 10, e0138860 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    60.Seager, R. et al. Climatology, variability, and trends in the U.S. vapor pressure deficit, an important fire-related meteorological quantity. J. Appl. Meteorol. Climatol. 54, 1121–1141 (2015).ADS 
    Article 

    Google Scholar 
    61.Taylor, K. E., Stouffer, R. J. & Meehl, G. A. An overview of CMIP5 and the experiment design. Bull. Am. Meteorol. Soc. 93, 485–498 (2012).ADS 
    Article 

    Google Scholar 
    62.Dudney, J. C. et al. Compounding effects of white pine blister rust, mountain pine beetle, and fire threaten four white pine species. Ecosphere 11, e03263 (2020).Article 

    Google Scholar 
    63.Schwandt, J. W., Lockman, I. B., Kliejunas, J. T. & Muir, J. A. Current health issues and management strategies for white pines in the western United States and Canada. Forest Pathol. 40, 226–250 (2010).Article 

    Google Scholar 
    64.Dudney, J. et al. Overstory removal and biological legacies influence long-term forest management outcomes on introduced species and native shrubs. Forest Ecol. Manag. 491, 119149 (2021).Article 

    Google Scholar 
    65.Abatzoglou, J. T. & Williams, A. P. Impact of anthropogenic climate change on wildfire across western US forests. Proc. Natl Acad. Sci. 113, 11770–11775 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    66.Goodsman, D. W., Lusebrink, I., Landhäusser, S. M., Erbilgin, N. & Lieffers, V. J. Variation in carbon availability, defense chemistry and susceptibility to fungal invasion along the stems of mature trees. N. Phytol. 197, 586–594 (2013).CAS 
    Article 

    Google Scholar 
    67.McDowell, N. et al. Mechanisms of plant survival and mortality during drought: why do some plants survive while others succumb to drought? N. Phytol. 178, 719–739 (2008).Article 

    Google Scholar 
    68.Bockino, N. K. & Tinker, D. B. Interactions of white pine blister rust and mountain pine beetle in whitebark pine ecosystems in the southern Greater Yellowstone Area. Nat. Areas J. 32, 31–40 (2012).Article 

    Google Scholar 
    69.Stephenson, N. L., Das, A. J., Ampersee, N. J., Bulaon, B. M. & Yee, J. L. Which trees die during drought? The key role of insect host-tree selection. J. Ecol. 107, 2383–2401 (2019).70.Griffin, D. & Anchukaitis, K. J. How unusual is the 2012–2014 California drought? Geophys. Res. Lett. 41, 2014GL062433 (2014).Article 

    Google Scholar 
    71.Paz‐Kagan, T. et al. What mediates tree mortality during drought in the southern Sierra Nevada? Ecol. Appl. 27, 2443–2457 (2017).PubMed 
    Article 

    Google Scholar 
    72.Zambino, P. J. Biology and pathology of Ribes and their implications for management of white pine blister rust. Pathol. 40, 264–291 (2010).Article 

    Google Scholar 
    73.Anderegg, W. R. L., Anderegg, L. D. L., Kerr, K. L. & Trugman, A. T. Widespread drought-induced tree mortality at dry range edges indicates that climate stress exceeds species’ compensating mechanisms. Glob. Change Biol. 25, 3793–3802 (2019).ADS 
    Article 

    Google Scholar 
    74.Bebber, D. P., Holmes, T. & Gurr, S. J. The global spread of crop pests and pathogens. Glob. Ecol. Biogeogr. 23, 1398–1407 (2014).Article 

    Google Scholar 
    75.Deyle, E. R., May, R. M., Munch, S. B. & Sugihara, G. Tracking and forecasting ecosystem interactions in real time. Proc. R. Soc. B: Biol. Sci. 283, 20152258 (2016).Article 

    Google Scholar 
    76.Deyle, E. R., Maher, M. C., Hernandez, R. D., Basu, S. & Sugihara, G. Global environmental drivers of influenza. Proc. Natl Acad. Sci. 113, 13081–13086 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    77.Cohen, J. M., Civitello, D. J., Venesky, M. D., McMahon, T. A. & Rohr, J. R. An interaction between climate change and infectious disease drove widespread amphibian declines. Glob. Change Biol. 25, 927–937 (2019).ADS 
    Article 

    Google Scholar 
    78.Kolb, T. E. et al. Observed and anticipated impacts of drought on forest insects and diseases in the United States. Ecol. Manag. 380, 321–334 (2016).Article 

    Google Scholar 
    79.Kutz, S. J. et al. The Arctic as a model for anticipating, preventing, and mitigating climate change impacts on host–parasite interactions. Vet. Parasitol. 163, 217–228 (2009).PubMed 
    Article 

    Google Scholar 
    80.Flower, C. E. & Gonzalez-Meler, M. A. Responses of temperate forest productivity to insect and pathogen disturbances. Annu. Rev. Plant Biol. 66, 547–569 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    81.Trant, A., Higgs, E. & Starzomski, B. M. A century of high elevation ecosystem change in the Canadian Rocky Mountains. Sci. Rep. 10, 9698 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    82.Wong, C. M. & Daniels, L. D. Novel forest decline triggered by multiple interactions among climate, an introduced pathogen and bark beetles. Glob. Change Biol. 23, 1926–1941 (2017).ADS 
    Article 

    Google Scholar 
    83.Endangered and Threatened Wildlife and Plants; Threatened Species Status for Pinus albicaulis (Whitebark Pine) With Section 4(d) Rule. Federal Register https://www.federalregister.gov/documents/2020/12/02/2020-25331/endangered-and-threatened-wildlife-and-plants-threatened-species-status-for-pinus-albicaulis (2020).84.Garrett, K. A., Dendy, S. P., Frank, E. E., Rouse, M. N. & Travers, S. E. Climate change effects on plant disease: genomes to ecosystems. Annu. Rev. Phytopathol. 44, 489–509 (2006).CAS 
    PubMed 
    Article 

    Google Scholar 
    85.PRISM Climate Group. PRISM Climate Group, Oregon State U. http://www.prism.oregonstate.edu/normals/.86.Abatzoglou, J. T. & Brown, T. J. A comparison of statistical downscaling methods suited for wildfire applications. Int. J. Climatol. 32, 772–780 (2012).Article 

    Google Scholar 
    87.Abatzoglou, J. T. Development of gridded surface meteorological data for ecological applications and modelling. Int. J. Climatol. 33, 121–131 (2013).Article 

    Google Scholar 
    88.Mitchell, K. E. et al. The multi-institution North American Land Data Assimilation System (NLDAS): Utilizing multiple GCIP products and partners in a continental distributed hydrological modeling system. J. Geophys. Res. Atmospheres https://doi.org/10.1029/2003JD003823@10.1002/(ISSN)2169-8996.GCIP3 (2018).89.Ritchie, J. & Dowlatabadi, H. Why do climate change scenarios return to coal? Energy 140, 1276–1291 (2017).Article 

    Google Scholar 
    90.R Core Team. R: A Language and Environment for Statistical Computing https://www.rproject.org/ (2017).91.Burns, K. S., Schoettle, A. W., Jacobi, W. R. & Mahalovich, M. F. White pine blister rust in the Rocky Mountain Region and options for management. Management. https://www.fs.fed.us/rm/pubs/rmrs_gtr206.pdf (2007).92.Fox, J. et al. car: Companion to applied regression (2019).93.Nakagawa, S. & Schielzeth, H. A general and simple method for obtaining R2 from generalized linear mixed-effects models. Methods Ecol. Evol. 4, 133–142 (2013).Article 

    Google Scholar 
    94.Zuur, A. F., Ieno, E. N., Walker, N., Saveliev, A. A. & Smith, G. M. Mixed Effects Models and Extensions in Ecology with R (Springer New York, 2009).95.Baker-Austin, C. et al. Emerging Vibrio risk at high latitudes in response to ocean warming. Nat. Clim. Change 3, 73–77 (2013).ADS 
    Article 

    Google Scholar 
    96.Lüdecke, D. ggeffects: Tidy data frames of marginal effects from regression models. J. Open Source Softw. 3, 772 (2018).ADS 
    Article 

    Google Scholar 
    97.Rohr, J. R., Raffel, T. R., Romansic, J. M., McCallum, H. & Hudson, P. J. Evaluating the links between climate, disease spread, and amphibian declines. Proc. Natl Acad. Sci. 105, 17436–17441 (2008).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    98.Wooldridge, J. M. Introductory Econometrics: A Modern Approach. 6th ed. (Cengage learning. Boston, MA, 2015).99.Berge, L. fixest: Fast Fixed-Effects Estimations. https://cran.rproject.org/web/packages/fixest/index.html (2020).100.Harrell, F. E. rms: Regression Modeling Strategies https://CRAN.R-project.org/package=rms (2020).101.Kelly, M., Guo, Q., Liu, D. & Shaari, D. Modeling the risk for a new invasive forest disease in the United States: An evaluation of five environmental niche models. Comput. Environ. Urban Syst. 31, 689–710 (2007).Article 

    Google Scholar 
    102.Meentemeyer, R. K. et al. Epidemiological modeling of invasion in heterogeneous landscapes: spread of sudden oak death in California (1990–2030). Ecosphere 2, 1–24 (2011).Article 

    Google Scholar 
    103.QGIS Development Team. QGIS Geographic Information System. http://qgis.osgeo.org/ (2020).104.Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).105.Hülsmann, L., Bugmann, H., Cailleret, M. & Brang, P. How to kill a tree: empirical mortality models for 18 species and their performance in a dynamic forest model. Ecol. Appl. 28, 522–540 (2018).PubMed 
    Article 

    Google Scholar 
    106.Cribbs, J., Nesmith, J., van Mantgem, P. & Dudney, J. Using stable isotope analysis and foliar growth measurements to understand physiological responses to drought in whitebark pine. Presented at the Ecological Society of America Symposium (2020).107.Farquhar, G. D. & Richards, R. A. Isotopic composition of plant carbon correlates with water-use efficiency of wheat genotypes. Funct. Plant Biol. 11, 539–552 (1984).CAS 
    Article 

    Google Scholar 
    108.Dudney, J. et al. Climate change and white pine blister rust. https://doi.org/10.17605/OSF.IO/PC9FM. (2021).109.Soberón, J. & Nakamura, M. Niches and distributional areas: concepts, methods, and assumptions. Proc. Natl Acad. Sci. 106, 19644–19650 (2009).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar  More

  • in

    Non-lethal effects of entomopathogenic nematode infection

    1.Gaugler, R. Entomopathogenic nematology (2002).2.Gaugler, R. Entomopathogenic Nematodes in Biological Control (CRC Press, 2018).Book 

    Google Scholar 
    3.Grewal, P. S., Ehlers, R.-U. & Shapiro-Ilan, D. I. Nematodes as Biocontrol Agents (CABI, 2005).Book 

    Google Scholar 
    4.Duncan, L. & McCoy, C. Vertical distribution in soil, persistence, and efficacy against citrus root weevil (coleoptera: Curculionidae) of two species of entomogenous nematodes (rhabditida: Steinernematidae; heterorhabditidae). Environ. Entomol. 25, 174–178 (1996).Article 

    Google Scholar 
    5.Duncan, L., McCoy, C. & Terranova, A. Estimating sample size and persistence of entomogenous nematodes in sandy soils and their efficacy against the larvae of Diaprepes abbreviatus in Florida. J. Nematol. 28, 56 (1996).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    6.Bullock, R., Pelosi, R. & Killer, E. Management of citrus root weevils (coleoptera: Curculionidae) on florida citrus with soil-applied entomopathogenic nematodes (nematoda: Rhabditida). Florida Entomologist 1–7 (1999).7.Koppenhöfer, A. M. & Fuzy, E. M. Steinernema scarabaei for the control of white grubs. Biol. Control 28, 47–59 (2003).Article 

    Google Scholar 
    8.Grewal, P., Power, K., Grewal, S., Suggars, A. & Haupricht, S. Enhanced consistency in biological control of white grubs (coleoptera: Scarabaeidae) with new strains of entomopathogenic nematodes. Biol. Control 30, 73–82 (2004).Article 

    Google Scholar 
    9.Georgis, R. et al. Successes and failures in the use of parasitic nematodes for pest control. Biol. Control 38, 103–123 (2006).Article 

    Google Scholar 
    10.Labaude, S. & Griffin, C. T. Transmission success of entomopathogenic nematodes used in pest control. Insects 9, 72 (2018).Article 

    Google Scholar 
    11.Li, X.-Y., Cowles, R., Cowles, E., Gaugler, R. & Cox-Foster, D. Relationship between the successful infection by entomopathogenic nematodes and the host immune response. Int. J. Parasitol. 37, 365–374 (2007).CAS 
    Article 

    Google Scholar 
    12.Castillo, J. C., Reynolds, S. E. & Eleftherianos, I. Insect immune responses to nematode parasites. Trends Parasitol. 27, 537–547 (2011).CAS 
    Article 

    Google Scholar 
    13.Ribeiro, C. et al. Insect immunity-effects of factors produced by a nematobacterial complex on immunocompetent cells. J. Insect Physiol. 45, 677–685 (1999).CAS 
    Article 

    Google Scholar 
    14.Garriga, A., Mastore, M., Morton, A., Garcia del Pino, F. & Brivio, M. F. Immune response of drosophila suzukii larvae to infection with the nematobacterial complex steinernema carpocapsae-xenorhabdus nematophila. Insects 11, 210 (2020).Article 

    Google Scholar 
    15.Ebrahimi, L., Niknam, G., Dunphy, G. & Toorchi, M. Side effects of immune response of colorado potato beetle, leptinotarsa decemlineata against the entomopathogenic nematode, steinernema carpocapsae infection. Invertebr. Surviv. J. 11, 132–142 (2014).
    Google Scholar 
    16.Ebrahimi, L., Niknam, G. & Lewis, E. Lethal and sublethal effects of iranian isolates of steinernema feltiae and heterorhabditis bacteriophora on the colorado potato beetle, leptinotarsa decemlineata. Biocontrol 56, 781–788 (2011).Article 

    Google Scholar 
    17.Chen, S., Li, J., Han, X. & Moens, M. Effect of temperature on the pathogenicity of entomopathogenic nematodes (Steinernema and Heterorhabditis spp.) to delia radicum. Biocontrol 48, 713–724 (2003).Article 

    Google Scholar 
    18.Mastore, M., Quadroni, S., Toscano, A., Mottadelli, N. & Brivio, M. F. Susceptibility to entomopathogens and modulation of basal immunity in two insect models at different temperatures. J. Therm. Biol 79, 15–23 (2019).CAS 
    Article 

    Google Scholar 
    19.Wojda, I. Temperature stress and insect immunity. J. Therm. Biol 68, 96–103 (2017).CAS 
    Article 

    Google Scholar 
    20.Lee, J. H., Dillman, A. R. & Hallem, E. A. Temperature-dependent changes in the host-seeking behaviors of parasitic nematodes. BMC Biol. 14, 1–17 (2016).Article 

    Google Scholar 
    21.Girling, R., Ennis, D., Dillon, A. & Griffin, C. The lethal and sub-lethal consequences of entomopathogenic nematode infestation and exposure for adult pine weevils, Hylobius abietis (coleoptera: Curculionidae). J. Invertebr. Pathol. 104, 195–202 (2010).CAS 
    Article 

    Google Scholar 
    22.Mastore, M., Arizza, V., Manachini, B. & Brivio, M. F. Modulation of immune responses of Rhynchophorus ferrugineus (insecta: Coleoptera) induced by the entomopathogenic nematode Steinernema carpocapsae (nematoda: Rhabditida). Insect Sci. 22, 748–760 (2015).CAS 
    Article 

    Google Scholar 
    23.Willett, D. S., Filgueiras, C. C., Nyrop, J. P. & Nault, B. A. Attract and kill: spinosad containing spheres to control onion maggot (Delia antiqua). Pest Manag. Sci. 76, 2720–2725 (2020).CAS 
    Article 

    Google Scholar 
    24.Willett, D. S., Filgueiras, C. C., Nyrop, J. P. & Nault, B. A. Field monitoring of onion maggot (Delia antiqua) fly through improved trapping. J. Appl. Entomol. 144, 382–387 (2020).Article 

    Google Scholar 
    25.Kaya, H. K. & Stock, S. P. Techniques in insect nematology. In Manual of Techniques in Insect Pathology, 281–324 (Elsevier, 1997).26.White, G. et al. A method for obtaining infective nematode larvae from cultures. Science (Washington) 66, 302–303 (1927).ADS 
    CAS 
    Article 

    Google Scholar 
    27.R Core Team. R: A. Language and Environment for Statistical Computing. R Foundation for Statistical Computing (Vienna, Austria, 2021).28.Wickham, H. et al. Welcome to the tidyverse. J. Open Sour. Softw. 4, 1686 (2019). https://doi.org/10.21105/joss.01686ADS 
    Article 

    Google Scholar 
    29.Fox, J. & Weisberg, S. An R Companion to Applied Regression third. (Sage, 2019).
    Google Scholar 
    30.Lenth, R. V. emmeans: Estimated Marginal Means, aka Least-Squares Means (2021). R package version 1.5.5-1.31.Franceschi, C. et al. Genes involved in immune response/inflammation, igf1/insulin pathway and response to oxidative stress play a major role in the genetics of human longevity: The lesson of centenarians. Mech. Ageing Dev. 126, 351–361 (2005).CAS 
    Article 

    Google Scholar 
    32.Kumar, S. et al. Lifespan extension in C. elegans caused by bacterial colonization of the intestine and subsequent activation of an innate immune response. Dev. Cell 49, 100–117 (2019).CAS 
    Article 

    Google Scholar 
    33.Bruno, P. et al. Entomopathogenic nematodes from Mexico that can overcome the resistance mechanisms of the western corn rootworm. Sci. Rep. 10, 1–12 (2020).Article 

    Google Scholar 
    34.Stock, S. P., Campos-Herrera, R., El-Borai, F. & Duncan, L. Steinernema khuongi n. sp. (panagrolaimomorpha, steinernematidae), a new entomopathogenic nematode species from Florida, USA. J. Helminthol. 93, 226–241 (2019).CAS 
    Article 

    Google Scholar 
    35.Nagelkerke, N. J. et al. A note on a general definition of the coefficient of determination. Biometrika 78, 691–692 (1991).MathSciNet 
    Article 

    Google Scholar  More

  • in

    High efficacy of microbial larvicides for malaria vectors control in the city of Yaounde Cameroon following a cluster randomized trial

    1.Guengant, J.-P. & May, J. F. African demography. Glob. J. Emerg. Mark. Econ. 5, 215–267 (2013).
    Google Scholar 
    2.Neiderud, C.-J. How urbanization affects the epidemiology of emerging infectious diseases. Infect. Ecol. Epidemiol. 5, 27060 (2015).PubMed 
    PubMed Central 

    Google Scholar 
    3.Eder, M. et al. Scoping review on vector-borne diseases in urban areas: Transmission dynamics, vectorial capacity and co-infection. Infect. Dis. Poverty 7, 1–24 (2018).Article 

    Google Scholar 
    4.Keiser, J. et al. Urbanization in sub-Saharan Africa and implication for malaria control. Am. J. Trop. Med. Hyg. 71, 118–127 (2004).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    5.Tusting, L. S. et al. Environmental temperature and growth faltering in African children: A cross-sectional study. Lancet Planet. Health 4, e116–e123. https://doi.org/10.1016/S2542-5196(20)30037-1 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    6.Hay, S. I. et al. Climate variability and malaria epidemics in the highlands of East Africa. Trends Parasitol. 21, 52–53 (2005).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    7.Bhatt, S. et al. The effect of malaria control on Plasmodium falciparum in Africa between 2000 and 2015. Nature 526, 207–211. https://doi.org/10.1038/nature15535 (2015).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    8.Kamau, A., Mogeni, P., Okiro, E. A., Snow, R. W. & Bejon, P. A systematic review of changing malaria disease burden in sub-Saharan Africa since 2000: Comparing model predictions and empirical observations. BMC Med. 18, 1–11 (2020).Article 

    Google Scholar 
    9.Nkumama, I. N., O’Meara, W. P. & Osier, F. H. Changes in malaria epidemiology in Africa and new challenges for elimination. Trends Parasitol. 33, 128–140 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    10.World Health Organization. World Malaria Report 2020: 20 Years of Global Progress and Challenges (WHO, 2020).Book 

    Google Scholar 
    11.World Health Organization. World Malaria Report 2012 (World Health Organization, 2012).Book 

    Google Scholar 
    12.Report, W. M. . Licence: CC BY-NC-SA 3.0 IGO (World Health Organization, 2019).
    Google Scholar 
    13.Ranson, H. & Lissenden, N. Insecticide resistance in African Anopheles mosquitoes: A worsening situation that needs urgent action to maintain malaria control. Trends Parasitol. 32, 187–196 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    14.Ranson, H. et al. Pyrethroid resistance in African anopheline mosquitoes: What are the implications for malaria control?. Trends Parasitol. 27, 91–98. https://doi.org/10.1016/j.pt.2010.08.004 (2011).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    15.Moyes, C. L. et al. Evaluating insecticide resistance across African districts to aid malaria control decisions. Proc. Natl. Acad. Sci. 117, 22042–22050 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    16.Moyes, C. L. et al. Contemporary status of insecticide resistance in the major Aedes vectors of arboviruses infecting humans. PLoS Negl. Tropi. Dis. 11, e0005625 (2017).Article 
    CAS 

    Google Scholar 
    17.Staedke, S. G. et al. LLIN evaluation in Uganda Project (LLINEUP)—Impact of long-lasting insecticidal nets with, and without, piperonyl butoxide on malaria indicators in Uganda: Study protocol for a cluster-randomised trial. Trials 20, 1–13 (2019).Article 

    Google Scholar 
    18.Hemingway, J. et al. Country-level operational implementation of the global plan for insecticide resistance management. Proc. Natl. Acad. Sci. 110, 9397–9402. https://doi.org/10.1073/pnas.1307656110 (2013).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    19.Chanda, E. et al. Scale-up of integrated malaria vector control: Lessons from Malawi. Bull. World Health Organ. 94, 475 (2016).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    20.Derua, Y. A., Kweka, E. J., Kisinza, W. N., Githeko, A. K. & Mosha, F. W. Bacterial larvicides used for malaria vector control in sub-Saharan Africa: Review of their effectiveness and operational feasibility. Parasit. Vectors 12, 1–18 (2019).Article 

    Google Scholar 
    21.WHO. Larval Source Management: A Supplementary Measure for Malaria Vector Control (World Health Organization, 2013).
    Google Scholar 
    22.Dambach, P. et al. Reduction of malaria vector mosquitoes in a large-scale intervention trial in rural Burkina Faso using Bti based larval source management. Malar. J. 18, 1–9 (2019).Article 

    Google Scholar 
    23.Fillinger, U. & Lindsay, S. W. Larval source management for malaria control in Africa: Myths and reality. Malar J. https://doi.org/10.1186/1475-2875-10-353 (2011).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    24.Shousha, A. T. Species-eradication: The eradication of Anopheles gambioe from Upper Egypt, 1942–1945. Bull. World Health Organ. 1, 309 (1948).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    25.Utzinger, J., Tozan, Y. & Singer, B. H. Efficacy and cost-effectiveness of environmental management for malaria control. Trop Med. Int. Health 6, 677–687. https://doi.org/10.1046/j.1365-3156.2001.00769.x (2001).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    26.Fillinger, U., Ndenga, B., Githeko, A. & Lindsay, S. W. Integrated malaria vector control with microbial larvicides and insecticide-treated nets in western Kenya: A controlled trial. Bull. World Health Organ. 87, 655–665 (2009).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    27.Choi, L., Majambere, S. & Wilson, A. L. Larviciding to prevent malaria transmission. Cochrane Database Syst. Rev. 8, CD012736 (2019).
    PubMed 
    PubMed Central 

    Google Scholar 
    28.Tusting, L. S. et al. Socioeconomic development as an intervention against malaria: A systematic review and meta-analysis. Lancet 382, 963–972. https://doi.org/10.1016/s0140-6736(13)60851-x (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    29.Choi, L. & Wilson, A. Larviciding to control malaria. Cochrane Database Syst. Rev. 2017, CD012736. https://doi.org/10.1002/14651858.CD012736 (2017).Article 
    PubMed Central 

    Google Scholar 
    30.Tusting, L. S. et al. Mosquito larval source management for controlling malaria. Cochrane Database Syst. Rev. 8, CD008923 (2013).
    Google Scholar 
    31.Geissbuhler, Y. et al. Microbial larvicide application by a large-scale, community-based program reduces malaria infection prevalence in urban Dar es Salaam. Tanzania. PLoS One. 4, e5107. https://doi.org/10.1371/journal.pone.0005107 (2009).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    32.Castro C. M. et al. Community-based environmental management for malaria control: evidence from a small-scale intervention in Dar es Salaam, Tanzania. Malar J. 8, 57 (2009)PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    33.Maheu-Giroux, M. & Castro, M. C. Impact of community-based larviciding on the prevalence of malaria infection in Dar es Salaam, Tanzania. PLoS ONE 8, e71638. https://doi.org/10.1371/journal.pone.0071638 (2013).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    34.World Health Organization. A Framework for Malaria Elimination (World Health Organization, 2017).
    Google Scholar 
    35.Wold Health Organization. 11th World Malaria Day “Ready to Beat Malaria” We are the Generation that can End Malaria (ed. World Health Organization). (World Health Oraganization, 2018).
    Google Scholar 
    36.Institut National de Statistiques and IFC. Enquête Démographique et de Santé du Cameroun (ed. INS. ICF.) 1–515 (INS, 2020).
    Google Scholar 
    37.Barbazan, P. et al. Control of Culex quinquefasciatus (Diptera: Culicidae) with Bacillus sphaericus in Maroua, Cameroon. J. Am. Mosq. Control Assoc. 13, 263–269 (1997).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    38.Hougard, J.-M. et al. Lutte contre Culex quinquefasciatus par Bacillus sphaericus: résultats d’une campagne pilote dans une agglomération urbaine d’Afrique équatoriale. Bulletin de l’organisation mondiale de la santé 71, 367–375 (1994).
    Google Scholar 
    39.Antonio-Nkondjio, C. et al. Anopheles gambiae distribution and insecticide resistance in the cities of Douala and Yaoundé (Cameroon): Influence of urban agriculture and pollution. Malar J. https://doi.org/10.1186/1475-2875-10-154 (2011).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    40.Tene Fossog, B. et al. Water quality and Anopheles gambiae larval tolerance to pyrethroids in the cities of Douala and Yaounde (Cameroon). J. Trop. Med. 2012, 1–10. https://doi.org/10.1155/2012/429817 (2012).Article 

    Google Scholar 
    41.Fondjo, E., Robert, V., Le Goff, G., Toto, J. & Carnevale, P. Urban malaria transmission in Yaounde (Cameroon). 2. Entomologic study in 2 semi urban districts. Bull. Soc. Pathol. Exot. 85, 57–63 (1992).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    42.Manga, L., Robert, V., Messi, J., Desfontaines, M. & Le Carnevale, P. paludisme urbain à Yaoundé, Cameroun. 1-Etude entomologique dans deux quartiers centraux. Mém. Soc. R Belge Entomol. 35, 155–162 (1992).
    Google Scholar 
    43.Doumbe-Belisse, P. et al. High malaria transmission sustained by Anopheles gambiae sl occurring both indoors and outdoors in the city of Yaoundé, Cameroon. Wellcome Open Res. 3, 164 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    44.Talipouo, A. et al. Malaria prevention in the city of Yaoundé: Knowledge and practices of urban dwellers. Malar. J. 18, 167. https://doi.org/10.1186/s12936-019-2799-6 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    45.Djamouko-Djonkam, L. et al. Implication of Anopheles funestus in malaria transmission in the city of Yaoundé, Cameroon. Parasite 27, 10 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    46.Djamouko-Djonkam, L. et al. Spatial distribution of Anopheles gambiae sensu lato larvae in the urban environment of Yaoundé, Cameroon. Infect. Dis. Poverty 8, 1–15 (2019).Article 

    Google Scholar 
    47.Nchoutpouen, E. et al. Culex species diversity, susceptibility to insecticides and role as potential vector of Lymphatic filariasis in the city of Yaoundé, Cameroon. PLoS Negl. Trop. Dis. 13, e0007229 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    48.Bamou, R. et al. Status of insecticide resistance and its mechanisms in Anopheles gambiae and Anopheles coluzzii populations from forest settings in south Cameroon. Genes 10, 741 (2019).CAS 
    PubMed Central 
    Article 

    Google Scholar 
    49.Ngadjeu, C. S. et al. Influence of house characteristics on mosquito distribution and malaria transmission in the city of Yaoundé, Cameroon. Malar. J. 19, 53 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    50.Nkahe, D. L. et al. Fitness cost of insecticide resistance on the life-traits of a Anopheles coluzzii population from the city of Yaoundé, Cameroon. Wellcome Open Res. 5, 171 (2020).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    51.Historique-Meteo. https://www.historique-meteo.net/afrique/cameroun/ (2020).52.Govella, N. et al. A new tent trap for sampling exophagic and endophagic members of the Anopheles gambiae complex. Malar. J. 8, 157. https://doi.org/10.1186/1475-2875-8-157 (2009).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    53.Wilson, A. L. et al. Evidence-based vector control? Improving the quality of vector control trials. Trends Parasitol. 31, 380–390 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    54.Fillinger, U., Sonye, G., Killeen, G. F., Knols, B. G. & Becker, N. The practical importance of permanent and semipermanent habitats for controlling aquatic stages of Anopheles gambiae sensu lato mosquitoes: Operational observations from a rural town in western Kenya. Trop. Med. Int. Health 9, 1274–1289 (2004).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    55.Majambere, S., Lindsay, S. W., Green, C., Kandeh, B. & Fillinger, U. Microbial larvicides for malaria control in The Gambia. Malar. J. 6, 1–14 (2007).Article 

    Google Scholar 
    56.Gillies, M. T. & Coetzee, M. A Supplement to the Anophelinae of Africa South of the Sahara (Afrotropical Region) (South African Institute for Medical Research, 1987).
    Google Scholar 
    57.Majambere, S. et al. Is mosquito larval source management appropriate for reducing malaria in areas of extensive flooding in The Gambia? A cross-over intervention trial. Am. J. Trop. Med. Hyg. 82, 176–184 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    58.Fossog Tene, B. et al. Resistance to DDT in an urban setting: Common mechanisms implicated in both M and S forms of Anopheles gambiae in the City of Yaoundé, Cameroon. PLoS ONE 8, e61408. https://doi.org/10.1371/journal.pone.0061408 (2013).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    59.Antonio-Nkondjio, C. et al. Investigation of mechanisms of bendiocarb resistance in Anopheles gambiae populations from the city of Yaoundé, Cameroon. Malar. J. 15, 424. https://doi.org/10.1186/s12936-016-1483-3 (2016).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    60.Antonio-Nkondjio, C. et al. Rapid evolution of pyrethroid resistance prevalence in Anopheles gambiae populations from the cities of Douala and Yaoundé (Cameroon). Malar. J. 14, 155. https://doi.org/10.1186/s12936-015-0675-6 (2015).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    61.Bamou, R. et al. Assessment of the Anophelinae blood seeking bionomic and pyrethroids resistance of local malaria vectors in the forest region of Southern Cameroon. JEZS 8, 1054–1062 (2020).
    Google Scholar 
    62.World Health Organization. The Role of Larviciding for Malaria Control in Sub-Saharan Africa: Interim Position Statement (World Health Organization, 2012).
    Google Scholar 
    63.Talipouo, A. et al. High insecticide resistance mediated by different mechanisms in Culex quinquefasciatus populations from the city of Yaoundé, Cameroon. Sci. Rep. 11, 1–11 (2021).Article 
    CAS 

    Google Scholar 
    64.Haines, A. et al. Promoting health and advancing development through improved housing in low-income settings. J. Urban Health 90, 810–831 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    65.Tusting, L. S. et al. Housing improvements and malaria risk in sub-Saharan Africa: A multi-country analysis of survey data. PLoS Med. 14, e1002234. https://doi.org/10.1371/journal.pmed.1002234 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    66.Tusting, L. S. et al. The evidence for improving housing to reduce malaria: A systematic review and meta-analysis. Malar. J. 14, 209. https://doi.org/10.1186/s12936-015-0724-1 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    67.Organization, W. H. Report of the Nineteenth WHOPES Working Group Meeting: WHO/HQ, Geneva, 8–11 February 2016: Review of Veeralin LN, VectoMax GR, Bactivec SC.(World Health Organization, 2016).68.Service M. Mosquito Ecology. Field Sampling Methods (Elsevier Applied Science, 1993).Book 

    Google Scholar 
    69.Edwards, F. W. Mosquitoes of the Ethiopian Region. HI.-Culicine Adults and Pupae. Mosquitoes of the Ethiopian Region. HI.-Culicine Adults and Pupae (1941).70.Edwards, F. W. Mosquitoes of the Ethiopian region. III, Culicine adults and pupae (Brit. Mus. Nat. Hist., 1941).71.Gillies, M. & Coetzee, M. A supplement to the Anophelinae of Africa south of the Sahara (Afrotropical region). Pub South Afr Inst Med Res 55, 143 (1987).
    Google Scholar 
    72.Wagtech, P. Water Quality Testing 1–139. https://www.palintest.com/product-categories/wagtech/ (Wagtech, 2012).73.Gillies, M. T. & DeMeillon, B. The Anopheline of Africa, south of the Sahara (Ethiopian zoogeographical region) Johannesburg: publication of south African Institute of Medical Research no. 54. (SAIMR, 1968).74.Gillies, T. & Coetzee, M. Supplement of the Anopheles of Africa south of Sahara (Afrotropical region) (Publication of the South African Institute of Medical Research, 1987).
    Google Scholar 
    75.Santolamazza, F. et al. Distribution of knock-down resistance mutations in Anopheles gambiae molecular forms in west and west-central Africa. Malar. J. 7, 74 (2008).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    76.Koekemoer, L., Kamau, L., Hunt, R. & Coetzee, M. A cocktail polymerase chain reaction assay to identify members of the Anopheles funestus (Diptera: culicidae) group. Am. J. Trop. Med. Hyg. 66, 804–811 (2002).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    77.Livak, K. J. Organization and mapping of a sequence on the Drosophila melanogaster X and Y chromosomes that is transcribed during spermatogenesis. Genetics 107, 611–634 (1984).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    78.Wirtz, R., Burkot, T., Graves, P. & Andre, R. Field evaluation of enzymelinked immunosorbent assays for P. falciparum and P. vivax sporozoites in mosquitoes (Diptera: Culicidae) from Papua, new Guinea. J. Med. Entomol. 24, 433–437. https://doi.org/10.1093/jmedent/24.4.433 (1987).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    79.Bass, C. et al. PCR-based detection of Plasmodium in Anopheles mosquitoes: A comparison of a new high-throughput assay with existing methods. Malar. J. 7, 177 (2008).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    80.WHO. Test Procedures for Insecticide Resistance Monitoring in Malaria Vector Mosquitoes (WHO, 2013). https://doi.org/10.1371/journal.pone.0013140.Book 

    Google Scholar 
    81.Bass, C. et al. Detection of knockdown resistance (kdr) mutations in Anopheles gambiae: A comparison of two new high-throughput assays with existing methods. Malar. J. 6, 1–14. https://doi.org/10.1186/1475-2875-6-111 (2007).CAS 
    Article 

    Google Scholar 
    82.Mulla, M. S., Norland, L. R., Fanara, D. M., Darwazeh, H. A. & McKean, D. W. Control of chironomid midges in recreational lakes. J. Econ. Entomol. 64, 300–307 (1971).CAS 
    Article 

    Google Scholar  More

  • in

    The rate and fate of N2 and C fixation by marine diatom-diazotroph symbioses

    Abundances of N2 fixing symbioses in the WTNATo date, the various marine symbiotic diatoms are notoriously understudied, and hence our understanding of their abundances and distribution patterns is limited [7]. In general, these symbiotic populations are capable of forming expansive blooms, but largely co-occur at low densities in tropical and subtropical waters with a few rare reports in temperate waters [26,27,28,29, 39,40,41,42]. The Rhizosolenia-Richelia symbioses have been more commonly reported in the North Pacific gyre [26, 27, 31], and the western tropical North Atlantic (WTNA) near the Amazon and Orinoco River plumes is an area where widespread blooms of the H. hauckii-Richelia symbioses are consistently recorded [28, 29, 42,43,44,45,46,47].In the summer of 2010, bloom densities (105−106 cells L−1) of the H. hauckii-Richelia symbioses were encountered at multiple stations with mesohaline (30–35 PSU) surface salinities (Supplementary Table 1). The R. clevei-Richelia symbioses were less abundant (2–30 cells L−1). Similar densities of H. hauckii-Richelia have been reported in the WTNA during spring (April–May) and summer seasons (June–July) (28–29; 46). In fall 2011, less dense symbiotic populations (0–50 cells L−1) were observed, and the dominant symbioses was the larger cell diameter (30–50 µm) H. membranaceus associated with Richelia. Previous observations of H. membranaeus-Richelia in this region are limited and reported as total cells (i.e., 12-218 cells) and highest numbers recorded in Aug–Sept in waters near the Bahama Islands [43]. On the other hand, Rhizosolenia-Richelia are even less reported in the WTNA, and most studies by quantitative PCR assays based on the nifH gene (for nitrogenase enzyme for N2 fixation) of the symbiont (44; 46–7). Unlike qPCR which cannot resolve if the populations are symbiotic or active for N2 fixation, the densities and activity reported here represent quantitative counts and measures of activity for symbiotic Rhizosolenia.The WTNA is largely influenced by both riverine and atmospheric dust deposition (e.g., Saharan dust) [48], including the silica necessary for the host diatom frustules, and trace metals (e.g., iron) necessary for photosynthesis by both partners and the nitrogenase enzyme (for N2 fixation) of the symbiont. We observed similar hydrographic conditions (i.e., low to immeasurable concentrations of dissolved N, sufficient concentrations of dissolved inorganic P and silicates, and variable surface salinities; 22; 28–29; 40–47) as reported earlier that favor high densities of H. hauckii-Richelia blooms. Unfortunately our data is too sparse to determine if these conditions are in fact priming and favoring the observed blooms of the H.hauckii-Richelia symbioses in summer 2010, and to a lesser extent in the Fall 2011.A biometric relationship between C and N activity and host biovolumeThe diatom-Richelia symbioses are considered highly host specific [10, 11], however, the driver of the specificity between partners remains unknown. We initially hypothesized that host selectivity could be related to the N2 fixation capacity of the symbiont. Moreover, it would be expected that the larger H. membranaceus and R. clevei hosts which are ~2–2.5 and 3.5–5 times, respectively, larger in cell dimensions than the H. hauckii cells would have higher N requirements (Supplementary Table 2). In fact, recently it was reported that the filament length of Richelia is positively correlated with the diameter of their respective hosts [22]. Thus, to determine if there is also a size dependent relationship between activity and cell biovolume, the enrichment of both 15N and 13C measured by SIMS was plotted as a function of symbiotic cell biovolume.Given the long incubation times (12 h) and previous work [32] that show fixation and transfer of reduced N to the host is rapid (i.e., within 30 min), we expected most if not all of the reduced N, or enrichment of 15N, to be transferred to the host diatom during the experiment (Fig. 1). Therefore, we measured and report the enrichment for the whole symbiotic cell, rather than the enrichment in the individual partners (Supplementary Table 2; Fig. 2). The enrichment of both 13C/12C and 15N/14N was significantly higher in the larger H. membranaceus-Richelia cells (atom % 13C: 1.5628–2.0500; atom % 15N: 0.8645–1.0200) than the enrichment measured in the smaller H. hauckii-Richelia cells (atom % 13C: 1.0700–1.3078; atom % 15N: 0.3642–0.7925) (Fig. 2) (13C, Mann–Whitney p = 0.009; 15N, Mann–Whitney p 50 symbiotic cells in a chain) were reported at station 2 with fully intact symbiotic Richelia filaments (2–3 vegetative cells and terminal heterocyst), and at station 25 chains were short (1–2 symbiotic cells) and associated with short Richelia filaments (only terminal heterocyst). Moreover, the symbiotic H. hauckii hosts possessed poor chloroplast auto-fluorescence at station 25 [46]. Given that the cells selected for NanoSIMS were largely single cells, rather than chains, we suspect that these cells were in a less than optimal cell state, which was also reflected in the low 13C/12C enrichment ratios and low estimated C-based growth rates (0.30–57 div d−1). These are particularly reduced compared to the growth rates recently reported for enrichment cultures of H. hauckii-Richelia (0.74–93 div d−1§) (Supplementary Table 2) [33].In 2011, higher cellular N2 fixation rates (15.4–27.2 fmols N cell−1 h−1) were measured for the large cell diameter H. membranaceus-Richelia, symbioses. Despite high rates of fixation, cell abundances were low (4–19 cells L−1), and resulted in a low overall contribution of the symbiotic diatoms to the whole water N2 ( >1%) and C-fixation ( >0.01%). The estimated C-based growth rates for H. membranaceus were high (1.9–3.5 div d−1), whereas estimated N-based growth rates (0.3–4 div d−1) were lower than previously published (33; 52–53). Hence the populations in 2011 were likely in a pre-bloom condition given the low cell densities.Estimating symbiotically derived reduced N to surface oceanTo date, determining the fate of the newly fixed N from these highly active but fragile symbiotic populations has been difficult. Thus, we attempted to estimate the excess N fixed and potentially available for release to the surround by using the numerous single cell-specific rates of N2 fixation determined by SIMS on the Hemiaulus spp.-Richelia symbioses (Supplementary Materials). Because the populations form chains during blooms and additionally sink, we calculated the size-dependent sinking rates for both single cells and chains ( >50 cells). Initially we hypothesized that sinking rates of the symbiotic associations would be more rapid than the N excretion rates, such that most newly fixed N would contribute less to the upper water column (sunlit).The sinking velocities were plotted (Fig. 5) as a function of cell radius at a range (min, max) of densities and included two different form resistances (∅ = 0.3 and 1.5). As expected, the combination of form resistance and density has a large impact on the sinking velocity. For example, a H. hauckii cell of similar radius (10 μm) and density (3300 kg m−3) but higher form resistance (0.3 vs. 1.5) sinks twice as fast at the lower form resistance (Fig. 5). This points to chain formation (e.g., increased form resistance) as a potential ecological adaptation to reduce sinking rates. Recently, colony formation was identified as an important phenotypic trait that could be traced back ancestrally amongst both free-living and symbiotic diatoms that presumably functions for maintaining buoyancy and enhancing light capture [22].Fig. 5: The influence of cell characteristics on estimated sinking velocity for symbiotic Hemiaulus spp.The range of diatom sinking speed predicted using the modified Stokes approximation for diatoms [74] and accounting for the symbioses (cylinders) having varying cell size characteristics (form resistance by altering chain length, density; Supplementary Table 4). Note that form resistance increases with chain length and that the longest chains would have sinking speeds less than 10 m d−1.Full size imageThe concentration of fixed N surrounding a H. hauckii and H. membranaceus cell were modeled (Supplementary Materials; Supplementary Table 4; Fig. 6). First, the cellular N requirement (QN, mol N cell−1) for a cell of known volume, V, as per the allometric formulation of Menden-Deuer and Lessard [71] is calculated by the following.$${{{{{{{mathrm{Q}}}}}}}}_{{{{{{{mathrm{N}}}}}}}} = (10^{ – 12}/12) times 0.76 ;times, {{{{{{{mathrm{V}}}}}}}}^{^{0.189}}$$
    (1)
    Fig. 6: The simplified case of diffusive nitrogen (N) exudate plumes for non-motile symbioses.The concentration of dissolved N (nmol L−1) is presented at of varying cell sizes (3 µm and 30 µm) for H. hauckii-Richelia (A and B, respectively) and H. membranaceus-Richelia (C and D, respectively) growing at specific growth rates of 0.4 d−1 (dashed red lines) or 0.68 d−1 (solid black lines). Exudation follows the same principle as diffusive uptake as per Kiorboe [72] in the absence of turbulence.Full size imageVolume calculations assume a cylindrical shape; whereas exudation assumes an equivalent spherical volume. Then, using published growth rates of 0.4 d−1 and 0.68 d−1 for the symbioses [52, 53], N uptake rate (VN) necessary to sustain the QN was determined. N loss was assumed to be a constant fraction (f) of the VN; this fraction was assumed to be 7.5% and 11% for H. hauckii and H. membranaceus, respectively, or the estimated excess N which was fixed given the assumed growth rate [31]. The excretion rate (EN) of the individual cells was then calculated as$${{{{{{{mathrm{E}}}}}}}}_{{{{{{{mathrm{N}}}}}}}} = {{{{{{{mathrm{fQ}}}}}}}}_{{{{{{{mathrm{N}}}}}}}}$$
    (2)
    The concentration of fixed N surrounding the cell (Cr) was iteratively calculated by the following:$${{{{{{{mathrm{C}}}}}}}}_{{{{{{{mathrm{r}}}}}}}} = {{{{{{{mathrm{E}}}}}}}}_{{{{{{{mathrm{N}}}}}}}}/(4pi * {{{{{mathrm{D}}}}}}* {{{{{mathrm{r}}}}}}_{{{{{mathrm{{x}}}}}}}) + {{{{{{{mathrm{C}}}}}}}}_{{{{{{{mathrm{i}}}}}}}}$$
    (3)
    The concentric radius (rx) as per Kiørboe [72] uses a diffusivity of N assumed to be 1.860 × 10−5 cm2 sec−1 and the background concentration of N (Ci) is assumed to be negligible. Figure 5 presents the results for the two symbioses: H. membranaceus and H. hauckii at the two growth rates and as chains or singlets. Mean sinking rates for cells with a high form resistance (e.g., chains) are More