More stories

  • in

    A metagenomic insight into the microbiomes of geothermal springs in the Subantarctic Kerguelen Islands

    MAG binning and general featuresFrom the four hot springs, we assembled four associated metagenomes and then binned a total of 42 MAGs. We recovered 12 MAGs from RB10 hot spring, 13 from RB13, 14 from RB32 and 3 from RB108. Out of these 42 MAGs, 7 were of high-quality, 25 of nearly-high quality, 9 of medium quality and 1 of low quality (Table 1) based on metagenomic standards26. The GC% was quite variable, ranging from 25.76 to 70.35% among all MAGs and between 32.15 and 69.21% only among the high- and near high-quality MAGs. With the exception of RB108 from which we only recovered bacterial MAGs, we retrieved both bacterial and archaeal MAGs in the other hot springs. Two thirds of the MAGs (26/42) were assigned to the domain Bacteria and the rest to the domain Archaea (16/42) (Table 2).Table 1 General characteristics of the 42 MAGs obtained from RB10, RB13, RB32 and RB108 samples.Full size tableTable 2 Classification of the MAGs based on the taxonomic classification of GTDB-Tk software (v2.1.0) and the Genome Taxonomy Database (07-RS207 release).Full size tableTaxonomic and phylogenomic analyses of MAGsThe taxonomic affiliation of the MAGs was investigated in detail through the workflow classify of GTDB-Tk (v 2.1.0; GTDB reference tree 07-RS207) (Table 2) and through de novo phylogenomic analyses (Fig. S1a–i). We also tried to classify MAGs on the basis of overall genome relatedness indices (OGRI), which is detailed in supplementary material (Text S1, Table S2, Fig. S2).De novo phylogenomic analyses globally confirmed the positioning of MAGs provided by GTDB-Tk, with high branching support. For Bacteria, GTDB-Tk analyses allowed us to place the MAGs in the following clades: six in the phylum Aquificota from the four different springs, comprising four MAGs belonging to the genus Hydrogenivirga (family Aquificaceae) (RB10-MAG07, RB13-MAG10, RB32-MAG07, RB108-MAG02), and two belonging to the family ‘Hydrogenobaculaceae’ (RB10-MAG12, RB32-MAG11) (Table 2, Fig. S1a). Their closest cultured relatives originated either from hot springs or from deep-sea hydrothermal vents27. Three MAGs from three geothermal springs belonged to the phylum Armatimonadota (RB10-MAG03, RB13-MAG04, RB32-MAG03) and had no close cultured relatives. Seven MAGs have been classified into the phylum Chloroflexota: three MAGs belonging to the genus Thermoflexus from three different springs (RB10-MAG04, RB13-MAG05, RB32-MAG02), one affiliating with the genus Thermomicrobium (RB32-MAG08), one falling into the family Ktedonobacteraceae (RB108-MAG03), one belonging to the class Dehalococcoidia (RB32-MAG04) and another one whose phylogenetic position is more difficult to assert because it is a MAG of medium quality (RB32-MAG14). Six MAGs from four various hot springs belonged to the phylum Deinococcota, and to the genera Thermus (RB10-MAG08, RB10-MAG11, RB13-MAG09, RB32-MAG10, RB108-MAG01) and Meiothermus (RB13-MAG13). One MAG belonged to the family ‘Sulfurifustaceae’ (RB13-MAG01), in the phylum Proteobacteria (Gamma-class). The MAG referenced as RB32-MAG13 was classified into the phylum ‘Patescibacteria’, in the class ‘Paceibacteria’, and was distantly related to MAGs originating from groundwater and from hot springs. Finally, two MAGs from two different springs belonged to the phylum WOR-3, in the Candidatus genus ‘Caldipriscus’ (RB32-MAG12, RB10-MAG09).For Archaea, almost all the MAGs reconstructed in this study, e.g. 15 of the 16 archaeal MAGs, belonged to the phylum Thermoproteota. Among them, four belonged to the genus Ignisphaera (RB10-MAG05, RB13-MAG08, RB13-MAG11, RB32-MAG05), three to the genus Infirmifilum (RB10-MAG06, RB13-MAG03, RB32-MAG09), two to the genus Zestosphaera (RB10-MAG02, RB13-MAG06), three to the family Acidilobaceae (RB10-MAG01, RB13-MAG02, RB32-MAG01) and two to the order Geoarchaeales (RB10-MAG10, RB32-MAG06). Additionally, one belonged to the family Thermocladiaceae (RB13-MAG07). Lastly, the MAG belonging to another phylum (RB13-MAG12) was affiliated with the ‘Aenigmatarchaeota’, class ‘Aenigmatarchaeia’, and was distantly related to MAGs from hot springs and from deep-sea hydrothermal vent sediments28,29.Out of these 42 MAGs, at least 19 MAGs corresponded to different taxa at the taxonomic rank of species or higher according to GTDB (Table 2). Eighteen of them belonged to lineages with several cultivated representatives including the species Thermus thermophilus. 13 new genomic species within the GTDB genera Hydrogenivirga, HRBIN17, Thermoflexus, SpSt-223, CADDYT01, Zestosphaera, Ignisphaera, Infirmifilum, Thermus, Thermus_A, Meiothermus_B, JAHLMO01 and Caldipriscus, and 6 putative new genomic genera belonging to the GTDB families Hydrogenobaculaceae, Acidilobaceae, WAQG01, Thermocladiaceae, Sulfurifustaceae and HR35 could be identified (Table 2). In addition, 9 MAGs belonged to lineages that are predominantly or exclusively known through environmental DNA sequences. Thus, these 42 MAGs comprised a broad phylogenetic range of Bacteria and Archaea at different levels of taxonomic organization, of which a large majority were not reported before.The approaches implemented here were not intended to describe the microbial diversity present in these sources in an exhaustive way or to compare them in a fine way, and cannot allow it because of a 2-year storage at 4 °C. This long storage has probably led to changes in the microbial communities and to the selective loss or enrichment of some taxa. As a result, no analysis of abundance or absence of taxa can be conducted from these metagenomes and the results are discussed taking this bias into account. However, they do provide an overview of the microbial diversity effectively present. If we compare the phylogenetic diversity of the MAGs found in the four hot springs, we can observe that 3 shared phyla (Deinococcota, Aquificota and Chloroflexota: phyla names according to GTDB), 2 shared families (Thermaceae and Aquificaceae), and one shared genus (Hydrogenivirga) were found among the four sources (Fig. 2). In addition, hot springs RB10, RB13 and RB32, that are geographically close ( More

  • in

    Economic and biophysical limits to seaweed farming for climate change mitigation

    Monte Carlo analysisSeaweed production costs and net costs of climate benefits were estimated on the basis of outputs of the biophysical and technoeconomic models described below. The associated uncertainties and sensitivities were quantified by repeatedly sampling from uniform distributions of plausible values for each cost and economic parameter (n = 5,000 for each nutrient scenario from the biophysical model, for a total of n = 10,000 simulations; see Supplementary Figs. 14 and 15)47,48,49,50,51,52. Parameter importance across Monte Carlo simulations (Fig. 3 and Supplementary Fig. 9) was determined using decision trees in LightGBM, a gradient-boosting machine learning framework.Biophysical modelG-MACMODS is a nutrient-constrained, biophysical macroalgal growth model with inputs of temperature, nitrogen, light, flow, wave conditions and amount of seeded biomass30,53, that we used to estimate annual seaweed yield per area (either in tons of carbon or tons of dry weight biomass per km2 per year)33,34. In the model, seaweed takes up nitrogen from seawater, and that nitrogen is held in a stored pool before being converted to structural biomass via growth54. Seaweed biomass is then lost via mortality, which includes breakage from variable ocean wave intensity. The conversion from stored nitrogen to biomass is based on the minimum internal nitrogen requirements of macroalgae, and the conversion from biomass to units of carbon is based on an average carbon content of macroalgal dry weight (~30%)55. The model accounts for farming intensity (sub-grid-scale crowding) and employs a conditional harvest scheme, where harvest is optimized on the basis of growth rate and standing biomass33.The G-MACMODS model is parameterized for four types of macroalgae: temperate brown, temperate red, tropical brown and tropical red. These types employed biophysical parameters from genera that represent over 99.5% of present-day farmed macroalgae (Eucheuma, Gracilaria, Kappahycus, Sargassum, Porphyra, Saccharina, Laminaria, Macrocystis)39. Environmental inputs were derived from satellite-based and climatological model output mapped to 1/12-degree global resolution, which resolves continental shelf regions. Nutrient distributions were derived from a 1/10-degree resolution biogeochemical simulation led by the National Center for Atmospheric Research (NCAR) and run in the Community Earth System Model (CESM) framework35.Two nutrient scenarios were simulated with G-MACMODS and evaluated using the technoeconomic model analyses described below: the ‘ambient nutrient’ scenario where seaweed growth was computed using surface nutrient concentrations without depletion or competition, and ‘limited nutrient’ simulations where seaweed growth was limited by an estimation of the nutrient supply to surface waters (computed as the flux of deep-water nitrate through a 100 m depth horizon). For each Monte Carlo simulation in the economic analysis, the technoeconomic model randomly selects either the 5th, 25th, 50th, 75th or 95th percentile G-MACMODS seaweed yield map from a normal distribution to use as the yield map for that simulation. Figures and numbers reported in the main text are based on the ambient-nutrient scenario; results based on the limited-nutrient scenario are shown in Supplementary Figures.Technoeconomic modelAn interactive web tool of the technoeconomic model is available at https://carbonplan.org/research/seaweed-farming.We estimated the net cost of seaweed-related climate benefits by first estimating all costs and emissions related to seaweed farming, up to and including the point of harvest at the farm location, then estimating costs and emissions related to the transportation and processing of harvested seaweed, and finally estimating the market value of seaweed products and either carbon sequestered or GHG emissions avoided.Production costs and emissionsSpatially explicit costs of seaweed production ($ tDW−1) and production-related emissions (tCO2 tDW−1) were calculated on the basis of ranges of capital costs ($ km−2), operating costs (including labour, $ km−2), harvest costs ($ km−2) and transport emissions per distance travelled (tCO2 km−1) in the literature (Table 1, Supplementary Tables 1 and 2); annual seaweed biomass (tDW km−2, for the preferred seaweed type in each grid cell), line spacing and number of harvests (species-dependent) from the biophysical model; as well as datasets of distances to the nearest port (km), ocean depth (m) and significant wave height (m).Capital costs were calculated as:$$c_{cap} = c_{capbase} + left( {c_{capbase} times left( {k_d + k_w} right)} right) + c_{sl}$$
    (1)
    where ccap is the total annualized capital costs per km2, ccapbase is the annualized capital cost per km2 (for example, cost of buoys, anchors, boats, structural rope) before applying depth and wave impacts, kd and kw are the impacts of depth and waviness on capital cost, respectively, each expressed as a multiplier between 0 and 1 modelled using our Monte Carlo method and applied only to grid cells with depth >500 m and/or significant wave height >3 m, respectively, and csl is the total annual cost of seeded line calculated as:$$c_{sl} = c_{slbase} times p_{sline}$$
    (2)
    where cslbase is the cost per metre of seeded line, and psline is the total length of line per km2, based on the optimal seaweed type grown in each grid cell.Operating and maintenance costs were calculated as:$$c_{op} = c_{ins} + c_{lic} + c_{lab} + c_{opbase}$$
    (3)
    where cop is the total annualized operating and maintenance costs per km2, cins is the annual insurance cost per km2, clic is the annual cost of a seaweed aquaculture license per km2, clab is the annual cost of labour excluding harvest labour, and copbase is all other operating and maintenance costs.Harvest costs were calculated as:$$c_{harv} = c_{harvbase} times n_{harv}$$
    (4)
    where charv is the total annual costs associated with harvesting seaweed per km2, charvbase is the cost per harvest per km2 (including harvest labour but excluding harvest transport), and nharv is the total number of harvests per year.Costs associated with transporting equipment to the farming location were calculated as:$$c_{eqtrans} = c_{transbase} times m_{eq} times d_{port}$$
    (5)
    where ceqtrans is total annualized cost of transporting equipment, ctransbase is the cost to transport 1 ton of material 1 km on a barge, meq is the annualized equipment mass in tons and dport is the ocean distance to the nearest port in km.The total production cost of growing and harvesting seaweed was therefore calculated as:$$c_{prod} = frac{{left( {c_{cap}} right) + left( {c_{op}} right) + left( {c_{harv}} right) + (c_{eqtrans})}}{{s_{dw}}}$$
    (6)
    where cprod is total annual cost of seaweed production (growth + harvesting), ccap is as calculated in equation (1), cop is as calculated in equation (3), charv is as calculated in equation (4), ceqtrans is as calculated in equation (5) and sdw is the DW of seaweed harvested annually per km2.Emissions associated with transporting equipment to the farming location were calculated as:$$e_{eqtrans} = e_{transbase} times m_{eq} times d_{port}$$
    (7)
    where eeqtrans is the total annualized CO2 emissions in tons from transporting equipment, etransbase is the CO2 emissions from transporting 1 ton of material 1 km on a barge, meq is the annualized equipment mass in tons and dport is the ocean distance to the nearest port in km.Emissions from maintenance trips to/from the seaweed farm were calculated as:$$e_{mnt} = left( {left( {2 times d_{port}} right) times e_{mntbase} times left( {frac{{n_{mnt}}}{{a_{mnt}}}} right)} right) + (e_{mntbase} times d_{mnt})$$
    (8)
    where emnt is total annual CO2 emissions from farm maintenance, dport is the ocean distance to the nearest port in km, nmnt is the number of maintenance trips per km2 per year, amnt is the area tended to per trip, dmnt is the distance travelled around each km2 for maintenance and emntbase is the CO2 emissions from travelling 1 km on a typical fishing maintenance vessel (for example, a 14 m Marinnor vessel with 2 × 310 hp engines) at an average speed of 9 knots (16.67 km h−1), resulting in maintenance vessel fuel consumption of 0.88 l km−1 (refs. 28,56).Total emissions from growing and harvesting seaweed were therefore calculated as:$$e_{prod} = frac{{(e_{eqtrans}) + (e_{mnt})}}{{s_{dw}}}$$
    (9)
    where eprod is total annual emissions from seaweed production (growth + harvesting), eeqtrans is as calculated in equation (7), emnt is as calculated in equation (8) and sdw is the DW of seaweed harvested annually per km2.Market value and climate benefits of seaweedFurther transportation and processing costs, economic value and net emissions of either sinking seaweed in the deep ocean for carbon sequestration or converting seaweed into usable products (biofuel, animal feed, pulses, vegetables, fruits, oil crops and cereals) were calculated on the basis of ranges of transport costs ($ tDW−1 km−1), transport emissions (tCO2-eq t−1 km−1), conversion cost ($ tDW−1), conversion emissions (tCO2-eq tDW−1), market value of product ($ tDW−1) and the emissions avoided by product (tCO2-eq tDW−1) in the literature (Table 1). Market value was treated as globally homogeneous and does not vary by region. Emissions avoided by products were determined by comparing estimated emissions related to seaweed production to emissions from non-seaweed products that could potentially be replaced by seaweed (including non-CO2 greenhouse gas emissions from land use)24. Other parameters used are distance to nearest port (km), water depth (m), spatially explicit sequestration fraction (%)57 and distance to optimal sinking location (km; cost-optimized for maximum emissions benefit considering transport emissions combined with spatially explicit sequestration fraction; see ‘Distance to sinking point calculation’ below). Each Monte Carlo simulation calculated the cost of both CDR via sinking seaweed and GHG emissions mitigation via seaweed products.For seaweed CDR, after the seaweed is harvested, it can either be sunk in the same location that it was grown, or be transported to a more economically favourable sinking location where more of the seaweed carbon would remain sequestered for 100 yr (see ‘Distance to optimal sinking point’ below). Immediately post-harvest, the seaweed still contains a large amount of water, requiring a conversion from dry mass to wet mass for subsequent calculations33:$$s_{ww} = frac{{s_{dw}}}{{0.1}}$$
    (10)
    where sww is the annual wet weight of seaweed harvested per km2 and sdw is the annual DW of seaweed harvested per km2.The cost to transport harvested seaweed to the optimal sinking location was calculated as:$$c_{swtsink} = c_{transbase} times d_{sink} times s_{ww}$$
    (11)
    where cswtsink is the total annual cost to transport harvested seaweed to the optimal sinking location, ctransbase is the cost to transport 1 ton of material 1 km on a barge, dsink is the distance in km to the economically optimized sinking location and sww is the annually harvested seaweed wet weight in t km−2 as in equation (10).The costs associated with transporting replacement equipment (for example, lines, buoys,anchors) to the farming location and hauling back used equipment at the end of its assumed lifetime (1 yr for seeded line, 5–20 yr for capital equipment by equipment type) in the sinking CDR pathway were calculated as:$$c_{eqtsink} = left( {c_{transbase} times left( {2 times d_{sink}} right) times m_{eq}} right) + (c_{transbase} times d_{port} times m_{eq})$$
    (12)
    where ceqtsink is the total annualized cost to transport both used and replacement equipment, ctransbase is the cost to transport 1 ton of material 1 km on a barge, meq is the annualized equipment mass in tons, dsink is the distance in km to the economically optimized sinking location and dport is the ocean distance to the nearest port in km. We assumed that the harvesting barge travels from the farming location directly to the optimal sinking location with harvested seaweed and replaced (used) equipment in tow (including used seeded line and annualized mass of used capital equipment), sinks the harvested seaweed, returns to the farm location and then returns to the nearest port (see Supplementary Fig. 16). These calculations assumed the shortest sea-route distance (see Distance to optimal sinking point).The total value of seaweed that is sunk for CDR was therefore calculated as:$$v_{sink} = frac{{left( {v_{cprice} – left( {c_{swtsink} + c_{eqtsink}} right)} right)}}{{s_{dw}}}$$
    (13)
    where vsink is the total value (cost, if negative) of seaweed farmed for CDR in $ tDW−1, vcprice is a theoretical carbon price, cswtsink is as calculated in equation (11), ceqtsink is as calculated in equation (12) and sdw is the annually harvested seaweed DW in t km−2. We did not assume any carbon price in our Monte Carlo simulations (vcprice is equal to zero), making vsink negative and thus representing a net cost.To calculate net carbon impacts, our model included uncertainty in the efficiency of using the growth and subsequent deep-sea deposition of seaweed as a CDR method. The uncertainty is expected to include the effects of reduced phytoplankton growth from nutrient competition, the relationship between air–sea gas exchange and overturning circulation (hereafter collectively referred to as the ‘atmospheric removal fraction’) and the fraction of deposited seaweed carbon that remains sequestered for at least 100 yr. The total amount of atmospheric CO2 removed by sinking seaweed was calculated as:$$e_{seqsink} = k_{atm} times k_{fseq} times frac{{tC}}{{tDW}} times frac{{tCO_2}}{{tC}}$$
    (14)
    where eseqsink is net atmospheric CO2 sequestered annually in t km−2, katm is the atmospheric removal fraction and kfseq is the spatially explicit fraction of sunk seaweed carbon that remains sequestered for at least 100 yr (see ref. 57).The emissions from transporting harvested seaweed to the optimal sinking location were calculated as:$$e_{swtsink} = e_{transbase} times d_{sink} times s_{ww}$$
    (15)
    where eswtsink is the total annual CO2 emissions from transporting harvested seaweed to the optimal sinking location in tCO2 km−2, etransbase is the CO2 emissions (tons) from transporting 1 ton of material 1 km on a barge (tCO2 per t-km), dsink is the distance in km to the economically optimized sinking location and sww is the annually harvested seaweed wet weight in t km−2 as in equation (10). Since the unit for etransbase is tCO2 per t-km, the emissions from transporting seaweed to the optimal sinking location are equal to (e_{mathrm{transbase}} times d_{mathrm{sink}} times s_{mathrm{ww}}), and the emissions from transporting seaweed from the optimal sinking location back to the farm are equal to 0 (since the seaweed has already been deposited, the seaweed mass to transport is now 0). Note that this does not yet include transport emissions from transport of equipment post-seaweed-deposition (see equation 16 below and Supplementary Fig. 16).The emissions associated with transporting replacement equipment (for example, lines, buoys, anchors) to the farming location and hauling back used equipment at the end of its assumed lifetime (1 yr for seeded line, 5–20 yr for capital equipment by equipment type)28,41 in the sinking CDR pathway were calculated as:$$e_{eqtsink} = left( {e_{transbase} times left( {2 times d_{sink}} right) times m_{eq}} right) + (e_{transbase} times d_{port} times m_{eq})$$
    (16)
    where eeqtsink is the total annualized CO2 emissions in tons from transporting both used and replacement equipment, etransbase is the CO2 emissions from transporting 1 ton of material 1 km on a barge, meq is the annualized equipment mass in tons, dsink is the distance in km to the economically optimized sinking location and dport is the ocean distance to the nearest port in km. We assumed that the harvesting barge travels from the farming location directly to the optimal sinking location with harvested seaweed and replaced (used) equipment in tow (including used seeded line and annualized mass of used capital equipment), sinks the harvested seaweed, returns to the farm location and then returns to the nearest port. These calculations assumed the shortest sea-route distance (see Distance to optimal sinking point).Net CO2 emissions removed from the atmosphere by sinking seaweed were thus calculated as:$$e_{remsink} = frac{{left( {e_{seqsink} – left( {e_{swtsink} + e_{eqtsink}} right)} right)}}{{s_{dw}}}$$
    (17)
    where eremsink is the net atmospheric CO2 removed per ton of seaweed DW, eseqsink is as calculated in equation (14), eswtsink is as calculated in equation (15), eeqtsink is as calculated in equation (16) and sdw is the annually harvested seaweed DW in t km−2.Net cost of climate benefitsSinkingTo calculate the total net cost and emissions from the production, harvesting and transport of seaweed for CDR, we combined the cost and emissions from the sinking-pathway cost and value modules. The total net cost of seaweed CDR per DW ton of seaweed was calculated as:$$c_{sinknet} = c_{prod} – v_{sink}$$
    (18)
    where csinknet is the total net cost of seaweed for CDR per DW ton harvested, cprod is the net production cost per DW ton as calculated in equation (6) and vsink is the net value (or cost, if negative) per ton seaweed DW as calculated in equation (13).The total net CO2 emissions removed per DW ton of seaweed were calculated as:$$e_{sinknet} = e_{remsink} – e_{prod}$$
    (19)
    where esinknet is the total net atmospheric CO2 removed per DW ton of seaweed harvested annually (tCO2 tDW−1 yr−1), eremsink is the net atmospheric CO2 removed via seaweed sinking annually as calculated in equation (17) and eprod is the net CO2 emitted from production and harvesting of seaweed annually as calculated in equation (9). For each Monte Carlo simulation, locations where esinknet is negative (that is, net emissions rather than net removal) were not included in subsequent calculations since they would not be contributing to CDR in that location under the given scenario. Note that these net emissions cases only occur in areas far from port in specific high-emissions scenarios. Even in such cases, most areas still contribute to CO2 removal (negative emissions), hence costs from locations with net removal were included.Total net cost was then divided by total net emissions to get a final value for cost per ton of atmospheric CO2 removed:$$c_{pertonsink} = frac{{c_{sinknet}}}{{e_{sinknet}}}$$
    (20)
    where cpertonsink is the total net cost per ton of atmospheric CO2 removed via seaweed sinking ($ per tCO2 removed), csinknet is total net cost per ton seaweed DW harvested as calculated in equation (18) ($ tDW−1) and esinknet is the total net atmospheric CO2 removed per ton seaweed DW harvested as calculated in equation (19) (tCO2 tDW−1).GHG emissions mitigationInstead of sinking seaweed for CDR, seaweed can be used to make products (including but not limited to food, animal feed and biofuels). Replacing convention products with seaweed-based products can result in ‘avoided emissions’ if the emissions from growing, harvesting, transporting and converting seaweed into products are less than the total greenhouse gas emissions (including non-CO2 GHGs) embodied in conventional products that seaweed-based products replace.When seaweed is used to make products, we assumed it is transported back to the nearest port immediately after being harvested. The annualized cost to transport the harvested seaweed and replacement equipment (for example, lines, buoys, anchors) was calculated as:$$c_{transprod} = frac{{left( {c_{transbase} times d_{port} times left( {s_{ww} + m_{eq}} right)} right)}}{{s_{dw}}}$$
    (21)
    where ctransprod is the annualized cost per ton seaweed DW to transport seaweed and equipment back to port from the farm location, ctransbase is the cost to transport 1 ton of material 1 km on a barge, meq is the annualized equipment mass in tons, dport is the ocean distance to the nearest port in km, sww is the annual wet weight of seaweed harvested per km2 as calculated in equation (10) and sdw is the annual DW of seaweed harvested per km2.The total value of seaweed that is used for seaweed-based products was calculated as:$$v_{product} = v_{mkt} – left( {c_{transprod} + c_{conv}} right)$$
    (22)
    where vproduct is the total value (cost, if negative) of seaweed used for products ($ tDW−1), vmkt is how much each ton of seaweed would sell for, given the current market price of conventional products that seaweed-based products replace ($ tDW−1), ctransprod is as calculated in equation (21) and cconv is the cost to convert each ton of seaweed to a usable product ($ tDW−1).The annualized CO2 emissions from transporting harvested seaweed and equipment back to port were calculated as:$$e_{transprod} = frac{{left( {e_{transbase} times d_{port} times left( {s_{ww} + m_{eq}} right)} right)}}{{s_{dw}}}$$
    (23)
    where etransprod is the annualized CO2 emissions per ton seaweed DW to transport seaweed and equipment back to port from the farm location, etransbase is the CO2 emissions from transporting 1 ton of material 1 km on a barge, meq is the annualized equipment mass in tons, dport is the ocean distance to the nearest port in km, sww is the annual wet weight of seaweed harvested per km2 as calculated in equation (10) and sdw is the annual DW of seaweed harvested per km2.Total emissions avoided by each ton of harvested seaweed DW were calculated as:$$e_{avprod} = e_{subprod} – left( {e_{transprod} + e_{conv}} right)$$
    (24)
    where eavprod is total CO2-eq emissions avoided per ton of seaweed DW per year (including non-CO2 GHGs using a GWP time period of 100 yr), esubprod is the annual CO2-eq emissions avoided per ton seaweed DW by replacing a conventional product with a seaweed-based product, etransprod is as calculated in equation (23) and econv is the annual CO2 emissions per ton seaweed DW from converting seaweed into usable products. esubprod was calculated by converting seaweed DW to caloric content58 for food/feed and comparing emissions intensity per kcal to agricultural products24, or by converting seaweed DW into equivalent biofuel content with a yield of 0.25 tons biofuel per ton DW59 and dividing the CO2 emissions per ton fossil fuel by the seaweed biofuel yield.To calculate the total net cost and emissions from the production, harvesting, transport and conversion of seaweed for products, we combined the cost and emissions from the product-pathway cost and value modules. The total net cost of seaweed for products per ton DW was calculated as:$$c_{prodnet} = c_{prod} – v_{product}$$
    (25)
    where cprodnet is the total net cost per ton DW of seaweed harvested for use in products, cprod is the net production cost per ton DW as calculated in equation (6) and vproduct is the net value (or cost, if negative) per ton DW as calculated in equation (22).The total net CO2-eq emissions avoided per ton DW of seaweed used in products were calculated as:$$e_{prodnet} = e_{avprod} – e_{prod}$$
    (26)
    where eprodnet is the total net CO2-eq emissions avoided per ton DW of seaweed harvested annually (tCO2 tDW−1 yr−1), eavprod is the net CO2-eq emissions avoided by seaweed products annually as calculated in equation (24) and eprod is the net CO2 emitted from production and harvesting of seaweed annually as calculated in equation (9). For each Monte Carlo simulation, locations where eprodnet is negative (that is, net emissions rather than net emissions avoided) were not included in subsequent calculations since they would not be avoiding any emissions in that scenario.Total net cost was then divided by total net emissions avoided to get a final value for cost per ton of CO2-eq emissions avoided:$$c_{pertonprod} = frac{{c_{prodnet}}}{{e_{prodnet}}}$$
    (27)
    where cpertonprod is the total net cost per ton of CO2-eq emissions avoided by seaweed products ($ per tCO2-eq avoided), cprodnet is total net cost per ton seaweed DW harvested for products as calculated in equation (25) ($ tDW−1) and eprodnet is total net CO2-eq emissions avoided per ton seaweed DW harvested for products as calculated in equation (26) (tCO2 tDW−1).Parameter ranges for Monte Carlo simulationsFor technoeconomic parameters with two or more literature values (see Supplementary Table 1), we assumed that the maximum literature value reflected the 95th percentile and the minimum literature value represented the 5th percentile of potential costs or emissions. For parameters with only one literature value, we added ±50% to the literature value to represent greater uncertainty within the modelled parameter range. Values at each end of parameter ranges were then rounded before Monte Carlo simulations as follows: capital costs, operating costs and harvest costs to the nearest $10,000 km−2, labour costs and insurance costs to the nearest $1,000 km−2, line costs to the nearest $0.05 m−1, transport costs to the nearest $0.05 t−1 km−1, transport emissions to the nearest 0.000005 tCO2 t−1 km−1, maintenance transport emissions to the nearest 0.0005 tCO2 km−1, product-avoided emissions to the nearest 0.1 tCO2-eq tDW−1, conversion cost down to the nearest $10 tDW−1 on the low end of the range and up to the nearest $10 tDW−1 on the high end of the range, and conversion emissions to the nearest 0.01 tCO2 tDW−1.We extended the minimum range values of capital costs to $10,000 km−2 and transport emissions to 0 to reflect potential future innovations, such as autonomous floating farm setups that would lower capital costs and net-zero emissions boats that would result in 0 transport emissions. To calculate the minimum value of $10,000 km−2 for a potential autonomous floating farm, we assumed that the bulk of capital costs for such a system would be from structural lines and flotation devices, and we therefore used the annualized structural line (system rope) and buoy costs from ref. 41 rounded down to the nearest $5,000 km−2. The full ranges used for our Monte Carlo simulations and associated literature values are shown in Supplementary Table 1.Distance to optimal sinking pointDistance to the optimal sinking point was calculated using a weighted distance transform (path-finding algorithm, modified from code in ref. 60) that finds the shortest ocean distance from each seaweed growth pixel to the location at which the net CO2 removed is maximized (including impacts of both increased sequestration fraction and transport emissions for different potential sinking locations) and the net cost is minimized. This is not necessarily the location in which the seaweed was grown, since the fraction of sunk carbon that remains sequestered for 100 yr is spatially heterogeneous (see ref. 57). For each ocean grid cell, we determined the cost-optimal sinking point by iteratively calculating equations (11–20) and assigning dsink as the distance calculated by weighted distance transform to each potential sequestration fraction (0.01–1.00) in increments of 0.01. Except for transport emissions, the economic parameter values used for these calculations were the averages of unrounded literature value ranges; we assumed that the maximum literature value reflected the 95th percentile and the minimum literature value represented the 5th percentile of potential costs or emissions, or for parameters with only one literature value, we added ±50% to the literature value to represent greater uncertainty within the modelled parameter range. For transport and maintenance transport emissions, we extended the minimum values of the literature ranges to zero to reflect potential net-zero emissions transport options and used the mean values of the resulting ranges. The dsink that resulted in minimum net cost per ton CO2 for each ocean grid cell was saved as the final dsink map, and the associated sequestration fraction value that the seaweed is transported to via dsink was assigned to the original cell where the seaweed was farmed and harvested (Supplementary Fig. 19). If the cost-optimal location to sink using this method was the same cell where the seaweed was harvested, then dsink was 0 km and the sequestration fraction was not modified from its original value (Supplementary Fig. 18).Comparison of gigaton-scale sequestration area to previous estimatesPrevious related work estimating the ocean area suitable for macroalgae cultivation13 and/or the area that might be required to reach gigaton-scale carbon removal via macroalgae cultivation13,19,36 has yielded a wide range of results, primarily due to differences in modelling methods. For example, Gao et al. (2022)36 estimate that 1.15 million km2 would be required to sequester 1 GtCO2 annually when considering carbon lost from seaweed biomass/sequestered as particulate organic carbon (POC) and refractory dissolved organic carbon (rDOC), and assume that the harvested seaweed is sold as food such that the carbon in the harvested seaweed is not sequestered. The area (0.31 million km2) required to sequester 1 GtCO2 in our study assumes that all harvested seaweed is sunk to the deep ocean to sequester carbon.Additionally, Wu et al.19 estimates that roughly 12 GtCO2 could be sequestered annually via macroalgae cultivation in approximately 20% of the world ocean area (that is, 1.67% ocean area per GtCO2), which is a much larger area per GtCO2 than our estimate of 0.085% ocean area. This notable difference arises for several reasons (including differences in yields, which in Wu et al. are around 500 tDW yr−1 in the highest-yield areas, whereas yields in our cheapest sequestration areas from G-MACMODS average 3,400 tDW km−2 yr−1) that arise from differences in model methodology. First, Wu et al. model temperate brown seaweeds, while our study considers different seaweed types, many of which have higher growth rates, and uses the most productive seaweed type for each ocean grid cell. The G-MACMODS seaweed growth model we use also has a highly optimized harvest schedule, includes luxury nutrient uptake (a key feature of macroalgal nutrient physiology) and does not directly model competition with phytoplankton during seaweed growth. Finally, tropical red seaweeds (the seaweed type in our cheapest sequestration areas) grow year-round, while others, such as the temperate brown seaweeds modelled by Wu et al., only grow seasonally. These differences all contribute to higher productivity in our model, leading to a smaller area required for gigaton-scale CO2 sequestration compared with Wu et al.Conversely, the ocean areas we model for seaweed-based CO2 sequestration or GHG emissions avoided are much larger than the 48 million km2 that Froehlich et al.13 estimate to be suitable for macroalgae farming globally. Although our maps show productivity and costs everywhere, the purpose of our modelling was to evaluate where different types of seaweed grow best and how production costs and product values vary over space, to highlight the lowest-cost areas (which are often the highest-producing areas) under various technoeconomic assumptions.Comparison of seaweed production costs to previous estimatesAlthough there are not many estimates of seaweed production costs in the scientific literature, our estimates for the lowest-cost 1% area of the ocean ($190–$2,790 tDW−1) are broadly consistent with previously published results: seaweed production costs reported in the literature range from $120 to $1,710 tDW−1 (refs. 40,41,61,62), but are highly dependent on assumed seaweed yields. For example, Camus et al.41 calculate a cost of $870 tDW−1 assuming a minimum yield of 12.4 kgDW m−1 of cultivation line (equivalent to 8.3 kgDW m−2 using 1.5 m spacing between lines). Using the economic values from Camus et al. but with our estimates of average yield for the cheapest 1% production cost areas (2.6 kgDW m−2) gives a much higher average cost of $2,730 tDW−1. Contrarily, van den Burg et al.40 calculate a cost of $1,710 tDW−1 using a yield of 20 tDW ha−1 (that is, 2.0 kg m−2). Instead assuming the average yield to be that from our lowest-cost areas (that is, 2.6 kgDW m−2 or 26 tDW ha−1) would decrease the cost estimated by van den Burg et al. (2016) to $1,290 tDW−1. Most recently, Capron et al.62 calculate an optimistic scenario cost of $120 tDW−1 on the basis of an estimated yield of 120 tDW ha−1 (12 kg m−2; over 4.5 times higher than the average yield in our lowest-cost areas). Again, instead assuming the average yield to be that in our lowest-cost areas would raise Capron et al.’s production cost to $540 tDW−1 (between the $190–$880 tDW−1 minimum to median production costs in the cheapest 1% areas from our model; Fig. 1a,b).Data sourcesSeaweed biomass harvestedWe used spatially explicit data for seaweed harvested globally under both ambient and limited-nutrient scenarios from the G-MACMODS seaweed growth model33.Fraction of deposited carbon sequestered for 100 yrWe used data from ref. 57 interpolated to our 1/12-degree grid resolution.Distance to the nearest portWe used the Distance from Port V1 dataset from Global Fishing Watch (https://globalfishingwatch.org/data-download/datasets/public-distance-from-port-v1) interpolated to our 1/12-degree grid resolution.Significant wave heightWe used data for annually averaged significant wave height from the European Center for Medium-range Weather Forecasts (ECMWF) interpolated to our 1/12-degree grid resolution.Ocean depthWe used data from the General Bathymetric Chart of the Oceans (GEBCO).Shipping lanesWe used data of Automatic Identification System (AIS) signal count per ocean grid cell, interpolated to our 1/12-degree grid resolution. We defined a major shipping lane grid cell as any cell with >2.25 × 108 AIS signals, a threshold that encompasses most major trans-Pacific and trans-Atlantic shipping lanes as well as major shipping lanes in the Indian Ocean, the North Sea, and coastal routes worldwide.Marine protected areas (MPAs)We used data from the World Database on Protected Areas (WDPA) and defined an MPA as any ocean MPA >20 km2.Reporting summaryFurther information on research design is available in the Nature Portfolio Reporting Summary linked to this article. More

  • in

    Maize and ancient Maya droughts

    Evans, N. P. et al. Quantification of drought during the collapse of the classic Maya civilization. Science 361, 498–501 (2018).Article 
    ADS 
    CAS 

    Google Scholar 
    Gill, R. B. The Great Maya Droughts: Water, Life, and Death (University of New Mexico Press, 2001).
    Google Scholar 
    Coe, M. D. The Maya (Thames and Hudson, 1993).
    Google Scholar 
    Douglas, P. M. J. et al. Drought, agricultural adaptation, and sociopolitical collapse in the Maya Lowlands. Proc. Natl. Acad. Sci. USA 112, 5607–5612 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    Haug, G. H. et al. Climate and the collapse of Maya civilization. Science 299, 1731–1735 (2003).Article 
    ADS 
    CAS 

    Google Scholar 
    Ford, A. & Nigh, R. Origins of the Maya forest garden: Maya resource management. J. Ethnobiol. 29, 213–236 (2009).Article 

    Google Scholar 
    Anderson, E. N. et al. Las Plantas de los Mayas: Etnobotánica en Quintana Roo, México (CONABIO-ECOSUR, 2005).
    Google Scholar 
    Fedick, S. L. Maya cornucopia: Indigenous food plants of the Maya lowlands. in The Real Business of Ancient Maya Economies (eds. Masson, M. A., Freidel, D. A. & Demarest, A. A.). 224–237 (University Press Florida, 2020).Ford, A. & Clarke, K. C. Linking the past and present of the ancient Maya: Lowland land use, population distribution, and density in the Late Classic period. in The Oxford Handbook of Historical Ecology and Applied Archaeology (eds. Isendahl, C. & Stump, D.) (Oxford Handbook of Historical Ecology and Applied Archaeology, 2015).Ford, A. & Nigh, R. The Maya Forest Garden: Eight Millennia of Sustainable Cultivation of the Tropical Woodlands (Routledge, 2016).Gómez-Pompa, A. On maya silviculture. Mexican Stud. (Estudios Mexicanos) 3, 1–17 (1987).Article 

    Google Scholar 
    Beach, T., Luzzadder-Beach, S., Krause, S. & Walling, S. ‘Mayacene’ floodplain and wetland formation in the Rio Bravo watershed of northwestern Belize. Holocene 25(10), 1612–1622 (2015).Pohl, M. D. et al. Early agriculture in the Maya lowlands. Lat. Am. Antiq. 7, 355–372 (1996).Article 

    Google Scholar 
    Fedick, S. L. The Managed Mosaic: Ancient Maya Agriculture and Resource Use (University of Utah Press, 1996).
    Google Scholar 
    Mueller, A. D. et al. Recovery of the forest ecosystem in the tropical lowlands of northern Guatemala after disintegration of Classic Maya polities. Geology 38, 523–526 (2010).Article 
    ADS 

    Google Scholar 
    Hodell, D. A., Curtis, J. H. & Brenner, M. Possible role of climate in the collapse of Classic Maya civilization. Nature 375, 391–394 (1995).Article 
    ADS 
    CAS 

    Google Scholar 
    Islebe, G. A., Hooghiemstra, H., Brenner, M., Curtis, J. H. & Hodell, D. A. A Holocene vegetation history from lowland Guatemala. Holocene 6, 265–271 (1996).Article 
    ADS 

    Google Scholar 
    Medina-Elizalde, M., Polanco-Martínez, J. M., Lases-Hernández, F., Bradley, R. & Burns, S. Testing the ‘tropical storm’ hypothesis of Yucatan Peninsula climate variability during the Maya Terminal Classic Period. Quat. Res. 86, 111–119 (2016).Aragón-Moreno, A. A., Islebe, G. A., Torrescano-Valle, N. & Arellano-Verdejo, J. Middle and late Holocene mangrove dynamics of the Yucatan Peninsula, Mexico. J. South Am. Earth Sci. 85, 307–311 (2018).Article 
    ADS 

    Google Scholar 
    Aragón-Moreno, A. A., Islebe, G. A., Roy, P. D., Torrescano-Valle, N. & Mueller, A. D. Climate forcings on vegetation of the southeastern Yucatán Peninsula (Mexico) during the middle to late Holocene. Palaeogeogr. Palaeoclimatol. Palaeoecol. 495, 214–226 (2018).Article 

    Google Scholar 
    Kennett, D. J. et al. Development and disintegration of Maya political systems in response to climate change. Science 338, 788–791 (2012).Article 
    ADS 
    CAS 

    Google Scholar 
    Conde, C. et al. El Niño y la agricultura. in Los impactos de El Niño en México (ed. Magaña, V.). 103–135 (Dirección General de Protección Civil, Secretaría de Gobernación, México, 1999).Magaña, V. O., Vázquez, J. L., Pérez, J. L. & Pérez, J. B. Impact of El Niño on precipitation in Mexico. Geofísica Int. 42, 313–330 (2003).
    Google Scholar 
    Wahl, D., Byrne, R. & Anderson, L. An 8700 year paleoclimate reconstruction from the southern Maya lowlands. Quat. Sci. Rev. 103, 19–25 (2014).Article 
    ADS 

    Google Scholar 
    Nooren, K. et al. Climate impact on the development of Pre-Classic Maya civilisation. Clim. Past 14, 1253–1273 (2018).Article 

    Google Scholar 
    Palomo-Kumul, J., Valdez-Hernández, M., Islebe, G. A., Cach-Pérez, M. J. & El Andrade, J. L. Niño-Southern oscillation affects the water relations of tree species in the Yucatan Peninsula. Mexico. Sci. Rep. 11, 10451 (2021).Article 
    ADS 
    CAS 

    Google Scholar 
    Rosenswig, R. M., VanDerwarker, A. M., Culleton, B. J. & Kennett, D. J. Is it agriculture yet? Intensified maize-use at 1000 cal BC in the Soconusco and Mesoamerica. J. Anthropol. Archaeol. 40, 89–108 (2015).Article 

    Google Scholar 
    Mueller, A. D. et al. Climate drying and associated forest decline in the lowlands of northern Guatemala during the late Holocene. Quat. Res. 71, 133–141 (2009).Article 

    Google Scholar 
    Aragón-Moreno, A. A., Islebe, G. A. & Torrescano-Valle, N. A ~3800-yr, high-resolution record of vegetation and climate change on the north coast of the Yucatan Peninsula. Rev. Palaeobot. Palynol. 178, 35–42 (2012).Article 

    Google Scholar 
    Carrillo-Bastos, A., Islebe, G. A. & Torrescano-Valle, N. 3800 Years of quantitative precipitation reconstruction from the Northwest Yucatan Peninsula. PLoS ONE 8, e84333 (2013).Article 
    ADS 

    Google Scholar 
    Berglund, B. E. Human impact and climate changes—Synchronous events and a causal link?. Quat. Int. 105, 7–12 (2003).Article 

    Google Scholar 
    Vela-Peláez, A. A., Torrescano-Valle, N., Islebe, G. A., Mas, J. F. & Weissenberger, H. Holocene precipitation changes in the Maya forest, Yucatán peninsula. Mexico. Palaeogeogr. Palaeoclimatol. Palaeoecol. 505, 42–52 (2018).Article 
    ADS 

    Google Scholar 
    Torrescano-Valle, N. & Islebe, G. A. Holocene paleoecology, climate history and human influence in the southwestern Yucatán Peninsula. Rev. Palaeobot. Palynol. 217, 1–8 (2015).Article 

    Google Scholar 
    Anselmetti, F. S., Hodell, D. A., Ariztegui, D., Brenner, M. & Rosenmeier, M. F. Quantification of soil erosion rates related to ancient Maya deforestation. Geology 35, 915–918 (2007).Article 
    ADS 

    Google Scholar 
    Beach, T. et al. A review of human and natural changes in Maya Lowland wetlands over the Holocene. Quat. Sci. Rev. 28, 1710–1724 (2009).Article 
    ADS 

    Google Scholar 
    Kerr, M. T. Holocene Precipitation Variability, Prehistoric Agriculture, and Natural and Human-Set Fires in Costa Rica (University of Tennessee, 2019).
    Google Scholar 
    Ebert, C. E., Peniche May, N., Culleton, B. J., Awe, J. J. & Kennett, D. J. Regional response to drought during the formation and decline of Preclassic Maya societies. Quat. Sci. Rev. 173, 211–235 (2017).Article 
    ADS 

    Google Scholar 
    De la Barreda, B., Metcalfe, S. E. & Boyd, D. S. Precipitation regionalization, anomalies and drought occurrence in the Yucatan Peninsula, Mexico. Int. J. Climatol. 40, 4541–4555 (2020).Article 

    Google Scholar 
    Islebe, G. A. et al. Holocene Paleoecology and Paleoclimatology of south and south-eastern Mexico: A palynological approach. in Mexico´s Environmental Holocene and Anthropocene History (eds. Torrescano-Valle, N., Islebe, G. A. & Roy, P.) (Springer, 2019).Tuxill, J., Reyes, L. A., Moreno, L. L., Uicab, V. C. & Jarvis, D. I. All maize is not equal: Maize variety choices and Mayan foodways in rural Yucatan, Mexico. in Pre-Columbian Foodways: Interdisciplinary Approaches to Food, Culture, and Markets in Ancient Mesoamerica (eds. Staller, J. & Carrasco, M.) 467–486 (Springer, 2010).Torrescano-Valle, N., Ramírez-Barajas, P. J., Islebe, G. A., Vela-Pelaez, A. A. & Folan, W. J. Human influence versus natural climate variability. in The Holocene and Anthropocene Environmental History of Mexico: A Paleoecological Approach on Mesoamerica (eds. Torrescano-Valle, N., Islebe, G. A. & Roy, P. D.). 171–194 (Springer, 2019).Faegri, K. & Iversen, J. Textbook of Pollen Analysis (Wiley, 1989).
    Google Scholar 
    Ford, A. The Maya forest: A domesticated landscape. in The Maya World (eds. Hutson, S. R. & Ardren, T.). 519–539 (Routledge, 2020).Fedick, S. L. & Santiago, L. S. Large variation in availability of Maya food plant sources during ancient droughts. Proc. Natl. Acad. Sci. USA 119, 2115657118 (2022).Article 

    Google Scholar 
    Puleston, D. E. The role of ramón in Maya subsistence. in Maya Subsistence. 353–366 (Elsevier, 1982).Atran, S. et al. Itza Maya tropical agro-forestry [and comments and replies]. Curr. Anthropol. 34, 633–700 (1993).Article 

    Google Scholar 
    Dussol, L., Elliott, M., Michelet, D. & Nondédéo, P. Ancient Maya sylviculture of breadnut (Brosimum alicastrum Sw.) and sapodilla (Manilkara zapota (L.) P. Royen) at Naachtun (Guatemala): A reconstruction based on charcoal analysis. Quat. Int. 457, 29–42 (2017).Ebel, R., de Jesús Méndez Aguilar, M. & Putnam, H. R. Milpa: One sister got climate-sick. The impact of climate change on traditional Maya farming systems. Int. J. Sociol. Agric. Food (Online) 24, 175–199 (2018).
    Google Scholar 
    Hernández-González, O. & Vergara-Yoisura, S. Studies on the productivity of Brosimum alicastrum a tropical tree used for animal feed in the Yucatan Peninsula. Bothalia 22, 7 (2014).
    Google Scholar 
    Martínez-Ruiz, N. del R. & Larqué-Saavedra, A. Semilla de Ramón. in Alimentos Vegetales Autóctonos Iberoamericanos Subutilizados (eds. Sonia, S.-A. & Álvarez-Parrilla, E.). 177–192 (Fabro Editores, 2018).Hatfield, J. L. & Dold, C. Water-use efficiency: Advances and challenges in a changing climate. Front. Plant Sci. 10, 103 (2019).Article 

    Google Scholar 
    Basso, B. & Ritchie, J. T. Evapotranspiration in high-yielding maize and under increased vapor pressure deficit in the US Midwest. Agric. Environ. Lett. 3, 170039 (2018).Article 

    Google Scholar 
    Gregory, P. J., Simmonds, L. P. & Pilbeam, C. J. Soil type, climatic regime, and the response of water use efficiency to crop management. Agron. J. 92, 814–820 (2000).Article 

    Google Scholar 
    Moy, C. M., Seltzer, G. O., Rodbell, D. T. & Anderson, D. M. Variability of El Niño/Southern Oscillation activity at millennial timescales during the Holocene epoch. Nature 420, 162–165 (2002).Article 
    ADS 
    CAS 

    Google Scholar 
    Revelle, W. psych: Procedures for Psychological, Psychometric, and Personality Research. R package at https://CRAN.R-project.org/package=psych (2022).Wickham, H. & Bryan, J. readxl: Read Excel Files. R package at https://readxl.tidyverse.org/ (2022).Wei, T. et al. Package ‘corrplot’. Statistician 56, e24 (2017).
    Google Scholar 
    QGIS Development Team. QGIS Geographic Information System. QGIS Association at https://www.qgis.org (2022)Instituto Nacional de Estadistica Geographia e Informatica (INEGI). 1:1000000 Merida, Carta de Precipitacion. Merida, Yucatán, Mexico (1981). More

  • in

    Commerson’s dolphin population structure: evidence for female phylopatry and male dispersal

    Waples, R. S. & Gaggiotti, O. INVITED REVIEW: What is a population? An empirical evaluation of some genetic methods for identifying the number of gene pools and their degree of connectivity. Mol. Ecol. 15, 1419–1439 (2006).Article 
    CAS 

    Google Scholar 
    Mendez, M., Rosenbaum, H. C., Subramaniam, A., Yackulic, C. & Bordino, P. Isolation by environmental distance in mobile marine species: Molecular ecology of franciscana dolphins at their southern range. Mol. Ecol. 19, 2212–2228 (2010).Article 
    CAS 

    Google Scholar 
    De Meeûs, T. et al. Population genetics and molecular epidemiology or how to “débusquer la bête”. Infect. Genet. Evol. 7, 308–332 (2007).Article 

    Google Scholar 
    Durigan, M. et al. Population genetic analysis of Giardia duodenalis: Genetic diversity and haplotype sharing between clinical and environmental sources. MicrobiologyOpen 6, e00424 (2017).Article 

    Google Scholar 
    Amaral, A. R. et al. Seascape genetics of a globally distributed, highly mobile marine mammal: The short-beaked common dolphin (genus Delphinus). PLoS ONE 7, e31482 (2012).Article 
    ADS 
    CAS 

    Google Scholar 
    Mendez, M. et al. Molecular ecology meets remote sensing: Environmental drivers to population structure of humpback dolphins in the Western Indian Ocean. Heredity 107, 349–361 (2011).Article 
    CAS 

    Google Scholar 
    de los Angeles Bayas-Rea, R., Félix, F. & Montufar, R. Genetic divergence and fine scale population structure of the common bottlenose dolphin (Tursiops truncatus, Montagu) found in the Gulf of Guayaquil. Ecuador. PeerJ 6, e4589 (2018).Article 

    Google Scholar 
    Natoli, A., Peddemors, V. M. & Rus Hoelzel, A. Population structure and speciation in the genus Tursiops based on microsatellite and mitochondrial DNA analyses. J. Evol. Biol. 17, 363–375 (2004).Article 
    CAS 

    Google Scholar 
    Oliveira, L. R., Loizaga De Castro, R., Cárdenas-Alayza, S. & Bonatto, S. L. Conservation genetics of South American aquatic mammals: An overview of gene diversity, population structure, phylogeography, non-invasive methods and forensics. Mammal Rev. 42, 275–303 (2012).Article 

    Google Scholar 
    Vollmer, N. L. & Rosel, P. E. Fine-scale population structure of common bottlenose dolphins (Tursiops truncatus) in offshore and coastal waters of the US Gulf of Mexico. Mar. Biol. 164, 1–15 (2017).Article 

    Google Scholar 
    MacLeod, C. D. Global climate change, range changes and potential implications for the conservation of marine cetaceans: A review and synthesis. Endanger. Species Res. 7, 125–136 (2009).Article 

    Google Scholar 
    Hartl, D. L., Clark, A. G. & Clark, A. G. Principles of Population Genetics, Vol. 116 (Sinauer associates Sunderland, 1997).Thomas, C. D. et al. Extinction risk from climate change. Nature 427, 145 (2004).Article 
    ADS 
    CAS 

    Google Scholar 
    Reeves, R. R., Smith, B. D., Crespo, E. A. & Notarbartolo di Sciara, G. Dolphins, whales and porpoises: 2002–2010 conservation action plan for the world’s cetaceans, Vol. 58 (IUCN, 2003).Crespo, E. A. & Hall, M. A. In Marine Mammals, 463–490 (Springer, 2002).Crespo, E. A. et al. Direct and indirect effects of highseas fisheries on the marine mammal populations in the northern and central Patagonian coast. J. Northwest Atl. Fish. Sci. 22, 189–207 (1997).Article 

    Google Scholar 
    Harlin-Cognato, A. D., Markowitz, T., Würsig, B. & Honeycutt, R. L. Multi-locus phylogeography of the dusky dolphin (Lagenorhynchus obscurus): Passive dispersal via the west-wind drift or response to prey species and climate change?. BMC Evol. Biol. 7, 1–17 (2007).Article 

    Google Scholar 
    Hoelzel, A. Evolution of population genetic structure in marine mammal species. In Population genetics for animal conservation, 294–318 (Cambridge University Press, Cambridge, 2009).Fraser, C. I., Nikula, R., Ruzzante, D. E. & Waters, J. M. Poleward bound: Biological impacts of Southern Hemisphere glaciation. Trends Ecol. Evol. 27, 462–471 (2012).Article 

    Google Scholar 
    Louis, M. et al. Influence of past climate change on phylogeography and demographic history of narwhals, Monodon monoceros. Proc. R. Soc. B 287, 20192964 (2020).Article 
    CAS 

    Google Scholar 
    Skovrind, M. et al. Circumpolar phylogeography and demographic history of beluga whales reflect past climatic fluctuations. Mol. Ecol. 30, 2543–2559 (2021).Article 

    Google Scholar 
    Foote, A. D. et al. Ancient DNA reveals that bowhead whale lineages survived Late Pleistocene climate change and habitat shifts. Nat. Commun. 4, 1–7 (2013).Article 

    Google Scholar 
    Crespo, E. A. et al. Status, population trend and genetic structure of South American fur seals, Arctocephalus australis, in southwestern Atlantic waters. Mar. Mamm. Sci. 31, 866–890 (2015).Article 

    Google Scholar 
    Feijoo, M., Lessa, E. P., De Castro, R. L. & Crespo, E. A. Mitochondrial and microsatellite assessment of population structure of South American sea lion (Otaria flavescens) in the Southwestern Atlantic Ocean. Mar. Biol. 158, 1857–1867 (2011).Article 

    Google Scholar 
    Túnez, J. I., Cappozzo, H. L., Nardelli, M. & Cassini, M. H. Population genetic structure and historical population dynamics of the South American sea lion, Otaria flavescens, in north-central Patagonia. Genetica 138, 831–841 (2010).Article 

    Google Scholar 
    Oliveira, L., Ott, P. H., Grazziotin, F. G., White, B. & Bonatto, S. In Paper (SC/S11/RW26) presented to the Southern Right Whale Assessment Workshop (Commission International Whaling).Loizaga de Castro, R., Dans, S. L. & Crespo, E. A. Spatial genetic structure of dusky dolphin, Lagenorhynchus obscurus, along the argentine coast: Preserve what scale?. Aquat. Conserv. Mar. Freshw. Ecosyst. 26, 173–183 (2016).Article 

    Google Scholar 
    Pimper, L. E., Goodall, R. N. P. & Remis, M. I. First mitochondrial DNA analysis of the spectacled porpoise (Phocoena dioptrica) from Tierra del Fuego, Argentina. Mamm. Biol. 77, 459–462 (2012).Article 

    Google Scholar 
    Pichler, F. B. et al. Origin and radiation of Southern Hemisphere coastal dolphins (genus Cephalorhynchus). Mol. Ecol. 10, 2215–2223 (2001).Article 
    CAS 

    Google Scholar 
    Dawson, S. M. In Encyclopedia of Marine Mammals, 166–172 (Elsevier, 2018).Robineau, D., Goodall, R. N. P., Pichler, F. & Baker, C. S. Description of a new subspecies of Commerson’s dolphin, Cephalorhynchus commersonii (Lacépède, 1804), inhabiting the coastal waters of the Kerguelen Islands. Mammalia 71, 172–180 (2007).Article 

    Google Scholar 
    Crespo, E. A. et al. Cephalorhynchus commersonii, Commerson’s Dolphin. IUCN; The IUCN Red List of Threatened Species; 10-2017; 1-14 (2017).Goodall, R. Commerson’s dolphin Cephalorhynchus commersonii (Lacépède 1804). Handb. Mar. Mamm. 5, 241–267 (1994).
    Google Scholar 
    Coscarella, M. A. Ecologıa, comportamiento y evaluación del impacto de embarcaciones sobre manadas de tonina overa Cephalorhynchus commersonii en Bahıa Engano, Chubut (Universidad de Buenos Aires, Buenos Aires, 2005).Dellabianca, N. A. et al. Spatial models of abundance and habitat preferences of commerson’s and peale’s dolphin in southern patagonian waters. PLoS ONE 11, e0163441 (2016).Article 

    Google Scholar 
    Goodall, R. et al. Studies of Commerson’s dolphins, Cephalorhynchus commersonii, off Tierra del Fuego, 1976–1984. Report of the International Whaling Commission (Special Issue 9), 143–160 (1988).White, R. The Distribution of Seabirds and Marine Mammals in Falkland Islands Waters (Joint Nature Conservation Committee, 2002).Loizaga de Castro, R., Dans, S. L., Coscarella, M. A. & Crespo, E. A. Living in an estuary: Commerson’s dolphin (Cephalorhynchus commersonii (Lacépède, 1804)), habitat use and behavioural pattern at the Santa Cruz River, Patagonia, Argentina. Latin Am. J. Aquat. Res. 41, 985–991 (2013).Article 

    Google Scholar 
    Pedraza, S. Ecología poblacional de la tonina overa, Cephalorhynchus commersonii, (Lacépède, 1804) en el litoral patagónico. Unpublished PhD thesis, Universidad de Buenos Aires, Buenos Aires, Argentina (2008).Garaffo, G. V. et al. Modeling habitat use for dusky dolphin and Commerson’s dolphin in Patagonia. Mar. Ecol. Prog. Ser. 421, 217–227 (2011).Article 
    ADS 

    Google Scholar 
    Cipriano, F., Hevia, M. & Iñíguez, M. Genetic divergence over small geographic scales and conservation implications for Commerson’s dolphins (Cephalorhynchus commersonii) in southern Argentina. Mar. Mamm. Sci. 27, 701–718 (2011).Article 
    CAS 

    Google Scholar 
    Pimper, L. E., Baker, C. S., Goodall, R. N. P., Olavarría, C. & Remis, M. I. Mitochondrial DNA variation and population structure of Commerson’s dolphins (Cephalorhynchus commersonii) in their southernmost distribution. Conserv. Genet. 11, 2157–2168 (2010).Article 

    Google Scholar 
    O’Brien, S. J. A role for molecular genetics in biological conservation. Proc. Natl. Acad. Sci. 91, 5748–5755 (1994).Article 
    ADS 
    CAS 

    Google Scholar 
    Loizaga de Castro, R., Hoelzel, A. & Crespo, E. Behavioural responses of Argentine coastal dusky dolphins (Lagenorhynchus obscurus) to a biopsy pole system. Anim. Welf. 22, 13–23 (2013).Article 
    CAS 

    Google Scholar 
    Elphinstone, M. S., Hinten, G. N., Anderson, M. J. & Nock, C. J. An inexpensive and high-throughput procedure to extract and purify total genomic DNA for population studies. Mol. Ecol. Notes 3, 317–320 (2003).Article 
    CAS 

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

    Google Scholar 
    Hoelzel, A., Hancock, J. & Dover, G. Evolution of the cetacean mitochondrial D-loop region. Mol. Biol. Evol. 8, 475–493 (1991).CAS 

    Google Scholar 
    Kumar, S., Stecher, G. & Tamura, K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33, 1870–1874 (2016).Article 
    CAS 

    Google Scholar 
    Ruzzante, D. E. et al. Validation of close-kin mark–recapture (CKMR) methods for estimating population abundance. Methods Ecol. Evol. 10, 1445–1453 (2019).Article 

    Google Scholar 
    Faircloth, B. C., Branstetter, M. G., White, N. D. & Brady, S. G. Target enrichment of ultraconserved elements from arthropods provides a genomic perspective on relationships among H ymenoptera. Mol. Ecol. Resour. 15, 489–501 (2015).Article 
    CAS 

    Google Scholar 
    Faircloth, B. C. MSATCOMMANDER: Detection of microsatellite repeat arrays and automated, locus-specific primer design. Mol. Ecol. Resour. 8, 92–94 (2008).Article 
    CAS 

    Google Scholar 
    Zhan, L. et al. MEGASAT: Automated inference of microsatellite genotypes from sequence data. Mol. Ecol. Resour. 17, 247–256 (2017).Article 
    CAS 

    Google Scholar 
    Nei, M. Molecular Evolutionary Genetics (Columbia University Press, 1987).Librado, P. & Rozas, J. DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 25, 1451–1452 (2009).Article 
    CAS 

    Google Scholar 
    Schneider, S., Roessli, D. & Excoffier, L. Arlequin: A software for population genetics data analysis, version 2.000. Genetics Biometry Laboratory, Department of Anthropology, University of Geneva, Switzerland (2000).Excoffier, L., Smouse, P. E. & Quattro, J. M. Analysis of molecular variance inferred from metric distances among DNA haplotypes: Application to human mitochondrial DNA restriction data. Genetics 131, 479–491 (1992).Article 
    CAS 

    Google Scholar 
    Darriba, D., Taboada, G. L., Doallo, R. & Posada, D. jModelTest 2: More models, new heuristics and parallel computing. Nat. Methods 9, 772 (2012).Article 
    CAS 

    Google Scholar 
    Excoffier, L. & Lischer, H. E. Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Mol. Ecol. Resour 10, 564–567 (2010).Article 

    Google Scholar 
    Bandelt, H.-J., Forster, P. & Röhl, A. Median-joining networks for inferring intraspecific phylogenies. Mol. Biol. Evol. 16, 37–48 (1999).Article 
    CAS 

    Google Scholar 
    Fu, Y.-X. Statistical tests of neutrality of mutations against population growth, hitchhiking and background selection. Genetics 147, 915–925 (1997).Article 
    CAS 

    Google Scholar 
    Rogers, A. R. & Harpending, H. Population growth makes waves in the distribution of pairwise genetic differences. Mol. Biol. Evol. 9, 552–569 (1992).CAS 

    Google Scholar 
    Peakall, R. & Smouse, P. E. GENALEX 6: Genetic analysis in Excel. Population genetic software for teaching and research. Mol. Ecol. Notes 6, 288–295 (2006).Article 

    Google Scholar 
    Mantel, N. The detection of disease clustering and a generalized regression approach. Can. Res. 27, 209–220 (1967).CAS 

    Google Scholar 
    Drummond, A. J. & Rambaut, A. BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evol. Biol. 7, 214 (2007).Article 

    Google Scholar 
    Harlin, A. D., Markowitz, T., Baker, C. S., Würsig, B. & Honeycutt, R. L. Genetic structure, diversity, and historical demography of New Zealand’s dusky dolphin (Lagenorhynchus obscurus). J. Mammal. 84, 702–717 (2003).Article 

    Google Scholar 
    Rambaut, A., Suchard, M., Xie, D. & Drummond, A. Tracer v1. 6. http://beast.bio.ed.ac.uk/Tracer (2014).Van Oosterhout, C., Hutchinson, W. F., Wills, D. P. & Shipley, P. MICRO-CHECKER: Software for identifying and correcting genotyping errors in microsatellite data. Mol. Ecol. Notes 4, 535–538 (2004).Article 

    Google Scholar 
    Rice, W. R. Analyzing tables of statistical tests. Evolution 43, 223–225 (1989).
    Google Scholar 
    Goudet, J. FSTAT, a program to estimate and test gene diversities and fixation indices, version 2.9. 3. http://www2.unil.ch/popgen/softwares/fstat.htm (2001).Waples, R. S. & Do, C. LDNE: A program for estimating effective population size from data on linkage disequilibrium. Mol. Ecol. Resour. 8, 753–756 (2008).Article 

    Google Scholar 
    Pritchard, J. K., Stephens, M. & Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 155, 945–959 (2000).Article 
    CAS 

    Google Scholar 
    Evanno, G., Regnaut, S. & Goudet, J. Detecting the number of clusters of individuals using the software STRUCTURE: A simulation study. Mol. Ecol. 14, 2611–2620 (2005).Article 
    CAS 

    Google Scholar 
    Earl, D. A. STRUCTURE HARVESTER: A website and program for visualizing STRUCTURE output and implementing the Evanno method. Conserv. Genet. Resour. 4, 359–361 (2012).Article 

    Google Scholar 
    Queller, D. C. & Goodnight, K. F. Estimating relatedness using genetic markers. Evolution 43, 258–275 (1989).
    Google Scholar 
    Wilson, G. A. & Rannala, B. Bayesian inference of recent migration rates using multilocus genotypes. Genetics 163, 1177–1191 (2003).Article 

    Google Scholar 
    Milinkovitch, M. C., Leduc, R., Tiedemann, R. & Dizon, A. In Marine Mammals: Biology and Conservation (ed Evans, P. G. H. & Raga, J. A.) 325–359 (Springer, 2002).Pichler, F. Population structure and genetic variation in Hector’s dolphin (Cephalorhynchus hectori), ResearchSpace@ Auckland (2001).Pichler, F. & Baker, C. Loss of genetic diversity in the endemic Hector’s dolphin due to fisheries-related mortality. Proc. R. Soc. Lond. Ser. B Biol. Sci. 267, 97–102 (2000).Article 
    CAS 

    Google Scholar 
    Greenwood, P. J. Mating systems, philopatry and dispersal in birds and mammals. Anim. Behav. 28, 1140–1162 (1980).Article 

    Google Scholar 
    Chilvers, B. L. & Wilkinson, I. S. Philopatry and site fidelity of New Zealand sea lions (Phocarctos hookeri). Wildl. Res. 35, 463–470 (2008).Article 

    Google Scholar 
    Engelhaupt, D. et al. Female philopatry in coastal basins and male dispersion across the North Atlantic in a highly mobile marine species, the sperm whale (Physeter macrocephalus). Mol. Ecol. 18, 4193–4205 (2009).Article 
    CAS 

    Google Scholar 
    Möller, L. M. & Beheregaray, L. B. Genetic evidence for sex-biased dispersal in resident bottlenose dolphins (Tursiops aduncus). Mol. Ecol. 13, 1607–1612 (2004).Article 

    Google Scholar 
    Jansen van Vuuren, B., Best, P., Roux, J. P. & Robinson, T. Phylogeographic population structure in the Heaviside’s dolphin (Cephalorhynchus heavisidii): Conservation implications. Anim. Conserv. 5, 303–307 (2002).Article 

    Google Scholar 
    Pérez-Alvarez, M. J. et al. Microsatellite markers reveal strong genetic structure in the endemic Chilean dolphin. PLoS ONE 10, e0123956 (2015).Article 

    Google Scholar 
    Hamner, R. M., Pichler, F. B., Heimeier, D., Constantine, R. & Baker, C. S. Genetic differentiation and limited gene flow among fragmented populations of New Zealand endemic Hector’s and Maui’s dolphins. Conserv. Genet. 13, 987–1002 (2012).Article 

    Google Scholar 
    Pichler, F., Dawson, S., Slooten, E. & Baker, C. Geographic isolation of Hector’s dolphin populations described by mitochondrial DNA sequences. Conserv. Biol. 12, 676–682 (1998).Article 

    Google Scholar 
    Kraft, S. et al. From settlers to subspecies: Genetic differentiation in commerson’s Dolphins between South America and the Kerguelen Islands. Front. Mar. Sci. 8, 782512 (2021).Article 

    Google Scholar 
    Grant, W. & Bowen, B. W. Shallow population histories in deep evolutionary lineages of marine fishes: Insights from sardines and anchovies and lessons for conservation. J. Hered. 89, 415–426 (1998).Article 

    Google Scholar 
    Ponce, J. F., Rabassa, J., Coronato, A. & Borromei, A. M. Palaeogeographical evolution of the Atlantic coast of Pampa and Patagonia from the last glacial maximum to the Middle Holocene. Biol. J. Lin. Soc. 103, 363–379 (2011).Article 

    Google Scholar 
    Wright, S. Isolation by distance. Genetics 28, 114 (1943).Article 
    CAS 

    Google Scholar 
    Meirmans, P. G. Nonconvergence in B ayesian estimation of migration rates. Mol. Ecol. Resour. 14, 726–733 (2014).Article 

    Google Scholar  More

  • in

    A sensitive soil biological indicator to changes in land-use in regions with Mediterranean climate

    Takoutsing, B. et al. Assessment of soil health indicators for sustainable production of maize in smallholder farming systems in the highlands of Cameroon. Geoderma 276, 64–73 (2016).Article 
    ADS 
    CAS 

    Google Scholar 
    Wenzel, W. W. et al. Soil and land use factors control organic caron status and accumulation in agricultural soils of Lower Austria. Geoderma 409, 115595. https://doi.org/10.1016/j.geoderma.2021.115595 (2022).Article 
    ADS 
    CAS 

    Google Scholar 
    Rinot, O., Levy, G. J., Steinberger, Y., Svoray, T. & Eshel, G. Soil health assessment: A critical review of current methodologies and a proposed new approach. Sci. Total. Environ. 648, 1484–1491 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    Veum, K. S., Sudduth, K. A., Kremer, R. J. & Kitchen, R. (2017) Sensor data fusion for soil health assessment. Geoderma 305, 53–61 (2017).Article 
    ADS 

    Google Scholar 
    Nunes, M. R., Van Es, H. M., Schindelbeck, R., Ristow, A. J. & Ryan, M. No-till and cropping system diversification improve soil health and crop yield. Geoderma 328, 30–43 (2018).Article 
    ADS 
    CAS 

    Google Scholar 
    Chhipaa, V., Stein, A., Shankar, H., George, K. J. & Alidoost, F. Assessing and transferring soil health information in a hilly terrain. Geoderma 343, 130–138 (2019).Article 
    ADS 

    Google Scholar 
    Oliver, D. P., Bramley, R. G. V., Riches, D., Porter, I. & Edwards, J. A review of soil physical and chemical properties as indicators of soil quality in Australian viticulture. Aust. J. Grape Wine Res. 19, 129–139 (2013).Article 
    CAS 

    Google Scholar 
    Riches, D. et al. Review: soil biological properties as indicators of soil quality in Australian viticulture. Aust. J. Grape Wine Res. 19, 311–323 (2013).CAS 

    Google Scholar 
    Ritz, K., Black, H. I. J., Campbell, C. D., Harris, J. A. & Wood, C. Selecting biological indicators for monitoring soils: a framework for balancing scientific opinion to assist policy development. Ecol. Ind. 9, 1212–1221 (2009).Article 
    CAS 

    Google Scholar 
    Zhuo, Z., Kirchner, I., Pfahl, S. & Cubasch, U. Climate impact of volcanic eruptions: the sensitivity to eruption season and latitude in MPI-ESM ensemble experiments. Atmos. Chem. Phys. 21, 13425–13442 (2021).Article 
    ADS 
    CAS 

    Google Scholar 
    Griffiths, B. S., Bonkowski, M., Roy, J. & Ritz, K. Functional stability, substrate utilisation and biological indicators of soils following environmental impacts. Appl. Soil Ecol. 16(1), 49–61 (2001).Article 

    Google Scholar 
    Avidano, L., Gamalero, E., Cossa, G. P. & Carraro, E. Characterization of soil health in an Italian polluted site by using microorganisms as bioindicators. Appl. Soil Ecol. 30(1), 21–33 (2005).Article 

    Google Scholar 
    Pattison, A. B. et al. Development of key soil health indicators for the Australian banana industry. Appl. Soil Ecol. 40(1), 155–164 (2008).Article 

    Google Scholar 
    Damsma, K. M., Rose, M. T. & Cavagnaro, T. R. Landscape scale survey of indicators of soil health in grazing systems. Soil Res. 53(2), 154–167 (2015).Article 

    Google Scholar 
    Fine, A. K., van Es, H. M. & Schindelbeck, R. R. Statistics, scoring functions, and regional analysis of a comprehensive soil health database. Soil Sci. Soc. Am. J. 81(3), 589–601 (2016).Article 

    Google Scholar 
    Roper, W. R., Osmond, D. L., Heitman, J. L., Wagger, M. G. & Reberg-Horton, S. C. Soil Health indicators do not differentiate among agronomic management systems in North Carolina soils. Soil Sci. Soc. Am. J. 81(4), 828–843 (2016).Article 

    Google Scholar 
    Li, Z. et al. Rapid diagnosis of agricultural soil health: A novel soil health index based on natural soil productivity and human management. J. Environ. Manage. 277, 111402. https://doi.org/10.1016/j.jenvman.2020.111402 (2021).Article 

    Google Scholar 
    Oren, A. & Steinberger, Y. Catabolic profiles of soil fungal communities along a geographic climatic gradient in Israel. Soil Biol. Biochem. 40, 2578–2587 (2008).Article 
    CAS 

    Google Scholar 
    Yu, J., Glazer, N. & Steinberger, Y. Carbon utilization, microbial biomass, and respiration in biological soil crusts in the Negev Desert. Biol. Fert. Soils 50, 285–293 (2014).Article 
    CAS 

    Google Scholar 
    Van der Putten, W. H. et al. Plant–soil feedbacks: the past, the present and future challenges. J. Ecol. 101, 265–276 (2013).Article 

    Google Scholar 
    Sherman, C. & Steinberger, Y. Microbial functional diversity associated with plant litter decomposition along a climatic gradient. Microb. Ecol. 64, 399–415 (2012).Article 
    CAS 

    Google Scholar 
    Dwivedi, V. & Soni, P. A review on the role of soil microbial biomass in eco-restoration of degraded ecosystem with special reference to mining areas. J. Appl. Nat. Sci. 3(1), 151–158 (2011).Article 

    Google Scholar 
    Barreiro, A., Martín, A., Carballas, T. & Díaz-Raviña, M. Long-term response of soil microbial communities to fire and fire-fighting chemicals. Biol. Fertil. Soils 52, 963–975 (2016).Article 
    CAS 

    Google Scholar 
    Soil Science Division Staff. Soil survey manual. In USDA Handbook 18 (ed. Ditzler, C., Scheffe, K. & Monger, H.C.). (Washington, G. P. O., 2017).Campbell, C. D., Chapman, S. J., Cameron, C. M., Davidson, M. S. & Potts, J. M. A rapid microtiter plate method to measure carbon dioxide evolved from carbon substrate amendments so as to determine the physiological profiles of soil microbial communities by using whole soil. Appl. Environ. Microbiol. 69, 3593–3599 (2003).Article 
    ADS 
    CAS 

    Google Scholar 
    Anderson, J. P. E. & Domsch, K. H. Physiological method for quantitative measurement of microbial biomass in soils. Soil Biol. Biochem. 10, 215–221 (1978).Article 
    CAS 

    Google Scholar 
    Creamer, R. E., Stone, D., Berry, P. & Kuiper, I. Measuring respiration profiles of soil microbial communities across Europe using MicroResp™ method. Appl. Soil Ecol. 97, 36–43 (2016).Article 

    Google Scholar 
    Oren, A. & Steinberger, Y. Coping with artifacts induced by CaCO3–CO2–H2O equilibria in substrate utilization profiling of calcareous soils. Soil Biol. Biochem. 40, 2569–2577 (2008).Article 
    CAS 

    Google Scholar 
    Zak, J. C., Willig, M. R., Howard, D. L. & Wildman, G. Functional diversity of microbial communities: A quantitative approach. Soil Biol. Biochem. 26(9), 1101–1108 (1994).Article 

    Google Scholar 
    Hotelling, H. The most predictable criterion. J. Educ. Psychol. 26, 139–142 (1935).Article 

    Google Scholar 
    Morrison, D. F. Multivariate Statistical Methods 2nd edn. (McGraw-Hill, 1976).MATH 

    Google Scholar 
    Rencher, A. C. Methods of Multivariate Analysis (Wiley, Uk, 1995).MATH 

    Google Scholar 
    IBM Corp. Released 2020. IBM SPSS Statistics for Windows, Version 27.0. (Armonk, NY: IBM Corp., 2020)R Core Team. A language and environment for statistical computing (R Foundation for Statistical Computing, 2021).
    Google Scholar 
    Bartoń K. MuMIn: Multi-Model Inference. R package version 1.46.0, https://CRAN.R-project.org/package=MuMIn, 2022.Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer-Verlag, 2016).Book 
    MATH 

    Google Scholar 
    Johnson, J. B. & Omland, K. S. Model selection in ecology and evolution. Trends Ecol. Evol. 19, 101–108 (2004).Article 

    Google Scholar 
    Kiryushin, V. I. The management of soil fertility and productivity of agrocenoses in adaptive-landscape farming systems. Eurasian Soil Sci. 52, 1137–1145 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    Hermans, S. M. et al. Using soil bacterial communities to predict physic-chemical variables and soil quality. Microbiome 8, 79 (2020).Article 
    CAS 

    Google Scholar 
    Fierer, N. & Jackson, R. B. The diversity and biogeography of soil bacterial communities. Proc. Natl. Acad. Sci. U. S. A. 103, 626–631 (2006).Article 
    ADS 
    CAS 

    Google Scholar 
    Frey, S. D., Drijber, R., Smith, H. & Melillo, J. M. Microbial biomass, functional capacity, and community structure after 12 years of soil warming. Soil Biol. Biochem. 40, 2904–2907 (2008).Article 
    CAS 

    Google Scholar 
    Powlson, D. S., Brookes, P. C. & Christensen, B. T. Measurement of soil microbial biomass provides an early indication of changes in total soil organic matter due to straw incorporation. Soil Biol. Biochem. 19, 159–164 (1987).Article 
    CAS 

    Google Scholar 
    Brookes, P. C. The use of microbial parameters in monitoring soil pollution by heavy metals. Biol. Fertil. Soils. 19, 269–279 (1995).Article 
    CAS 

    Google Scholar 
    Hermans, S. M. et al. Bacteria as emerging indicators of soil condition. Appl. Environ. Microbiol. 83, e02826-e2916. https://doi.org/10.1128/AEM.02826-16 (2016).Article 

    Google Scholar 
    Liddicoat, C. et al. Can bacterial indicators of a grassy woodland restoration inform ecosystem assessment and microbiota-mediated human health?. Environ. Int. 129, 105–117 (2019).Article 

    Google Scholar 
    Jeanne, T., Parent, S. -É. & Hogue, R. Using a soil bacterial species balance index to estimate potato crop productivity. PLoS ONE 14, e0214089. https://doi.org/10.1371/journal.pone.0214089 (2019).Article 
    CAS 

    Google Scholar 
    Taylor, B. R. & Parkinson, D. Respiration and mass loss rates of aspen and pine leaf litter decomposing in laboratory microcosms. Can. J. Bot. 66, 1948–1959 (1988).Article 

    Google Scholar 
    Wardle, D. A. & Parkinson, D. Response of the soil microbial biomass to glucose, and selective inhibitors, across a soil moisture gradient. Soil Biol. Biochem. 22, 825–834 (1990).Article 
    CAS 

    Google Scholar 
    Baldrian, P., Merhautová, V., Petránková, M., Cajthaml, T. & Snajdr, J. Distribution of microbial biomass and activity of extracellular enzymes in a hardwood forest soil reflect soil moisture content. Appl. Soil Ecol. 46, 177–182 (2010).Article 

    Google Scholar 
    Holland, T. C. et al. The response of soil biota to water availability in vineyards. Pedobiol. Int. J. Soil Biol. 56, 9–14 (2013).
    Google Scholar 
    Yu, J. & Steinberger, Y. Vertical distribution of microbial-community functionality under the canopies of Zygophyllum dumosum and Hammada scoparia in the Negev Desert. Microb. Ecol. 62, 218–227 (2011).Article 

    Google Scholar 
    Wardle, D. A. & Parkinson, D. Interaction between microclimatic variables and the soil microbial biomass. Biol. Fertil. Soils. 9, 273–280 (1990).Article 

    Google Scholar  More

  • in

    Ocean acidification causes fundamental changes in the cellular metabolism of the Arctic copepod Calanus glacialis as detected by metabolomic analysis

    Using a targeted metabolomics approach, we showed that late copepodite stages of the keystone Arctic copepod Calanus glacialis experience important changes in several central energetic pathways following exposure to decreasing pH. These findings shed light on the physiological changes underpinning the effects of OA on fitness related traits such as ingestion rate and metabolic rate previously observed in this species17,18,20.Cellular energy metabolismCellular energy production was altered consistently in both stage CIV and CV, with concentrations of higher energy adenosine phosphates (ATP and ADP) increasing, and concentrations of the lower energy, less-phosphorylated AMP decreasing, with decreasing seawater pH. Moreover, Phospho-L-arginine, which in crustaceans functions as phosphagen in the replenishment of ATP from ADP during transient energy demands32, increased significantly in stage CV. These changes strongly suggest that exposure to low pH affects energy production and expenditure in both developmental stages, although with nuanced differences.NAD+ increased significantly in stage CIV. NAD+ is an essential redox carrier receiving electrons from oxidative processes in the glycolysis, the TCA cycle, and fatty acid oxidation to form NADH. A high NAD+/NADH ratio facilitates higher rates of these reactions and thus potentially higher rates of ATP production (unfortunately, the LC-HRMS could not detect NADH). But most importantly, the produced NADH serves as electron donors to ATP synthesis in the oxidative phosphorylation. For every ATP produced in the oxidative phosphorylation one NADH is oxidised back to NAD+. High rates of ATP production in the oxidative phosphorylation would therefore amass NAD+, as observed in stage CIV. Conversely, ATP production in the glycolysis and TCA cycle consumes NAD+ (9 NAD+ per 4 ATP) and glycolytic ATP production would decrease the NAD+ concentration with decreasing pH.Heterotrophic organisms generally face a trade-off between rate and yield of ATP production. The efficient low rate/high yield production in the TCA cycle/oxidative phosphorylation may prevail under certain circumstances, whereas under other circumstances, the less efficient high rate/low yield production in the glycolysis may predominate33. Because glycolysis and oxidative phosphorylation compete for ADP, the one dominate over the other in terms of rates depending on the substrate being metabolised. In stage CIV copepodites, the TCA cycle pathway was enriched in the MetPA, and metabolites associated with glycolysis and the TCA cycle showed significant changes in their concentrations at decreasing seawater pH. Glucose, the entry point to glycolysis, increased significantly with decreasing pH. High levels of blood glucose (hyperglycemia) have been observed as a general stress response in decapod crustaceans34. Copepods have no circulatory system (although they have a dorsal heart) but might nevertheless react similarly on the cellular level. Along with the significant increase in glucose, lactate decreased significantly with pH in stage CIV. Lactate is an inevitable end product of glycolysis, because lactate dehydrogenase has the highest Vmax of any enzymes in the glycolytic pathway and the Keq for pyruvate to lactate is far in the direction of lactate35. Accordingly, although the glycolysis was not enriched in the MetPA, conceivably because none of its intermediate metabolites were included in the analyses (the protocol did not allow for it), we hypothesise that stage CIV copepodites experience a general down-regulation of glycolysis under decreasing pH. Alternatively, the amassing of glucose and depletion of lactate could also indicate increased gluconeogenesis. Gluconeogenesis occurs during starvation to replenish glycogen stores and ingestion rates did decrease in stage CIV20. But again, we did not target any intermediates in our analyses, and thus cannot firmly conclude on this.Phosphofructokinase-1 is a key regulatory enzyme of glycolysis36. This enzyme is allosterically inhibited by ATP and activated by AMP, and interestingly this regulation is augmented by low pH37,38. Thus, phosphofructokinase-1 could be key to the down-regulation of glycolysis we hypothesise. The fact that we found increasing oxygen consumption with decreasing pH in stage CIV copepodites from the same experiment adds further momentum to this line of thought20. It seems that stage CIV copepodites might experience the so-called Pasteur effect—a decrease in glycolysis at increased levels of oxygen uptake—when exposed to decreasing pH39. Although ATP and AMP were significantly affected also in stage CV, glucose, pyruvate and lactate did not change with decreasing pH, which perhaps indicate absence of the down-regulation of glycolysis we hypothesise for stage CIV. There is, nevertheless, one indication that down-regulation may in fact occur also in this developmental stage. Alpha-glycerophosphate decreased significantly with decreasing pH in stage CV. This molecule is an intermediate in the transfer of electrons from NADH produced by glycolysis in the cytosol to the oxidative phosphorylation in the mitochondria, and decreased concentrations could result from down-regulation of the glycolysis also in stage CV copepodites.The TCA cycle was enriched for stage CIV and most of the measured TCA cycle metabolites (alpha-ketoglutarate, succinate, fumarate, and malate) showed increasing concentrations at decreasing pH. Trigg et al.40 observed a similar increase in concentrations of TCA cycle-related metabolites in the Dungeness crab, Cancer magister (Dana, 1852), at decreased pH and concluded that TCA cycle activity is upregulated under OA. Since NAD+ is the product of the transport of electrons from the TCA cycle to the oxidative phosphorylation in the mitochondria,  the increase in NAD+ concentration we observed in stage CIV could reflect an increase in the flow of electrons from the TCA cycle to the oxidative phosphorylation, and by extension an increase in the energy production by the TCA cycle and the oxidative phosphorylation. There is negative feedback from the TCA cycle to glycolysis through inhibition of phosphofructokinase-1 by citrate, a metabolite of the TCA cycle38. Unfortunately, we did not target citrate in our targeted approach to specifically test this hypothesis, but the amassing of NAD+ do provide additional support to the idea that glycolysis is down-regulated at decreasing pH. Again, there is a less clear picture of how cellular energy metabolism is affected by decreasing pH in stage CV when compared to stage CIV. There was no clear pattern of regulation of TCA metabolites, and the TCA cycle was not enriched in the MetPA. Nevertheless, alpha-ketoglutarate concentrations did increase with decreasing pH in CVs.The glyoxylate/dicarboxylate cycle was also enriched in the pathway analysis, but this is probably also a result of the increases in concentrations of alpha-ketoglutarate, succinate, fumarate, and malate, and we are unable to distinguish it from the TCA cycle based on the set of metabolites analysed.Conclusively, lowered glycolysis due to inhibition of phosphofructokinase-1 and upregulation of the TCA cycle and oxidative phosphorylation at low pH in stage CIV appear plausible causes for the changes in ATP, ADP and AMP concentrations we observed. Alongside these effects, down-regulation of transcription of genes involved in the glycolysis were also present in nauplii of C. glacialis exposed to 35–38 days of low pH conditions16. On the other hand, studies on the acclimatisation and adaptation to OA in another calanoid copepod species, Pseudocalanus acuspes (Giesbrecht, 1881), showed no increase in expression of mitochondrial genes at pHT 7.54, which would have been expected if the TCA cycle or oxidative phosphorylation is upregulated41. Interestingly, De Wit et al.41 also showed natural selection in a large fraction of mitochondrial genes under OA conditions. Even evolutionarily conserved sequences, such as cytochrome oxidase subunit I, were under selection and it was hypothesised that the mitochondrial function of oxidative phosphorylation is a target for natural selection in copepods at low pH41.Besides its role in the transfer of energy from the mitochondria to the cell, ATP is also used to fuel cell homeostasis and active cellular acid–base regulation by activation of ATP-dependent enzymes involved in osmo-ionic- and acid–base regulation. In crustaceans, acid–base status is linked to ion regulation, and is maintained primarily through ion transport mechanisms moving acid and/or base equivalents between the extracellular fluid and the ambient water42. One prominent process in this respect is regulation by Na+/K+-ATPase42,43. While this regulation takes place in the gills of decapod crustaceans43, it is located in the maxillary glands and other specialised organs on the swimming legs of copepods44. Any extensive ATPase mediated pH regulation could have manifested itself by decreasing ATP concentrations, but this is contrary to what we report here. Interestingly, while the pCO2-sensitive isopod Cymodoce truncata (Leach, 1814) is able to maintain its cellular ATP concentration at the expense of the concentration of carbonate anhydrase (an enzyme involved in the cellular transformation of water and CO2 to bicarbonate ions and H+ prior to the ATPase mediated transport of H+ across the cell membrane), the pCO2-tolerant isopod Dynamene bifida (Torelli, 1930) upregulates ATP with no functional compromise to CA concentrations45. Finally, C. glacialis nauplii have shown upregulation of Na+/H+-antiporters independent of ATPase as a response to OA16, which one could hypothesise also may be the case in the copepodites. Arctic populations of the amphipod Gammarus setosus also do not experience increased ATPase activity during OA conditions46. It seems that C. glacialis faces OA without any ATP dependent acid/base regulation activity.Glycolysis is the first step of catabolism of carbohydrates for the production of energy. When down-regulating glycolysis the copepods may be increasingly dedicated to catabolism of amino acids e.g. through oxidative deamination of glutamate and/or catabolism of fatty acids through beta-oxidation to produce the energy they require21. Both lead to the production of molecules entering the TCA cycle and ultimately the oxidative phosphorylation for energy production in the mitochondria.Amino acid metabolismOf the free amino acids which were significantly affected by decreasing pH, the majority decreased in concentration, for both stage CIV and CV copepodites. This could be an indication of changes in protein synthesis at decreasing pH. Supporting this idea, biosynthesis of aminoacyl-tRNA was indicated as significantly enriched in the MetPA in both stage CIV and CV. Aminoacyl-tRNA partakes in the elongation of the protein amino acid chain during protein synthesis and the enrichment was most likely due to the changes in concentration of the many amino acids tested. One probable cause of protein synthesis is the increased demands of enzymes needed to handle stress at low pH, including for example enzymes involved in acid–base- and osmo-regulation or regulation of energy production. Increased protein synthesis caused by OA conditions has been observed in larvae of the purple sea urchin Strongylocentrotus purpuratus (O.F. Müller, 1776), where in vivo rates of protein synthesis and ion transport increased ∼50%47. Costs of protein synthesis are high and have shown to constitute a major part of copepod metabolic demand48 and we did observe significant increases in metabolic rate in copepodite stage CIV from the same experiment20 giving further credit to the idea that protein synthesis was upregulated.An alternate but not mutually exclusive explanation is that the copepods experience increased amino acid catabolism under OA. Glutamate increased in stage CIV accompanied by a significant increase in alpha-ketoglutarate in both stage CIV and CV. Alpha-ketoglutarate is part of the metabolic pathway of glutamine, glutamate and arginine in which glutamate acts as an intermediate in catabolism of these amino acids when it is deaminated to alpha-ketoglutarate to enter the TCA cycle49. Glutamate metabolism (in conjunction with alanine and aspartate metabolism) was significantly enriched in the MetPA in both stage CIV and CV, and these changes could be taken as an indication of a shift towards amino acid catabolism with decreasing pH. The key enzyme catalysing the oxidative deamination of glutamate is glutamate dehydrogenase (GDH), which functions in both directions: deamination of glutamate to form alpha-ketoglutarate or formation of glutamate from alpha-ketoglutarate. Studies on the ribbed mussel, Modiolus dernissus (Dillwyn, 1817), have shown that the balance of this action is strongly pushed towards deamination when pH decreases from 8.0 to 7.550. GDH is activated by ADP, and one could argue that the increase in ADP we observed would work against this shift, but ADP activates GDH mainly in the glutamate forming direction51. The other measured amino acids enter the TCA cycle at different positions we unfortunately could not target in our analyses. Glutamate also partakes in the arginine biosynthesis pathway in which it is transformed to ornithine to enter the urea cycle. Arginine biosynthesis was enriched in the MetPA and it is therefore possible that decreasing pH also changes amino acid catabolism to increase urea excretion. Decreasing pH has a similar depressing effect on amino acid concentration in the gills of the shore crab Carcinus maenas (Linnaeus, 1758) which also has been interpreted as a sign of increased protein catabolism52. Hammer and colleagues52 argued that this increase in catabolism served to buffer H+ by supplying nitrogen to NH4 formation in the cells. All in all, we hypothesise that increased amino acid catabolism, possibly driven by changes in GDH activity, and the down-regulation of glycolysis by inhibition of phosphofructokinase-1 may be major drivers of a shift from carbohydrate metabolism towards catabolism of amino acids.D-glutamine/D-glutamate metabolism was highly enriched in the MetPA in both developmental stages. Several studies show enriched D-glutamine/D-glutamate metabolism in crustaceans [e.g. 53], but they offer no explanation of its function or the reason why it is enriched. While D-glutamate act in neurotransmission, this action is evolutionarily restricted to ctenophores, and biochemical measurements of D-amino acid concentrations have shown absence of D-glutamate in crustaceans54,55.We observed no changes in concentrations of 8-oxy-2-deoxyguanosine, a product of DNA oxidation. Furthermore, regulation of cellular response to oxidative stress is down-regulated in C. glacialis nauplii16, and OA may not induce oxidative stress in C. glacialis.Fatty acid metabolismBesides their importance in energy storage as wax esters, fatty acids are involved in many central processes in cells, most prominently through their function as cell membrane building blocks. Many fatty acids are obtained from the diet but some longer chain fatty acids, such as 20:1n-9 are synthesised de novo in copepods56. Stage CV copepodites experienced increases in most of the targeted free fatty acids (18 of 21) with decreasing pH. Only one of those 18 increased significantly, but since the direction of change were the same in all, we argue that the pattern of change does merit consideration. Conspicuous exceptions were eicosapentaenoic acid (EPA) 20:5n-3 and docosahexaenoic acid (DHA) 22:6n-3, which both decreased significantly. The only other study (to our knowledge) of metabolomic effects of environmental changes in copepods showed the exact same response to starvation in a mix of C. finmarchicus and C. helgolandicus stage CV copepodites, with most fatty acids increasing while EPA and DHA decreased in concentration57. EPA and DHA are key marine polyunsaturated fatty acids (PUFAs) exclusively produced by marine algae. They contribute a major fraction of the fatty acids of cell membrane phospholipids58, and zooplankton reproductive production is highly dependent on especially EPA59. EPA and DHA are key for cell membrane fluidity, which for calanoid copepods is especially important during diapause in the deep during copepodite stage CV60. They have also been linked to diapause buoyancy control, and are selectively metabolized in diapausing copepodites61. The importance of EPA and DHA for cell membrane integrity may be central for the changes we observed. Glycerol-3-phosphate, the precursor for the glycerol backbone of cell membrane phospholipids also decreased significantly and it seems decreasing pH could affect cell membrane turnover.Changing fatty acid concentration could be due to either a change in lipid intake from feeding or increased fatty acid catabolism. While ingestion rates decreased in stage CIV, they were unchanged in stage CV with decreasing pH20. Also, Thalassiosira weissflogii (Grunow) G.Fryxell & Hasle, 1977, the diatom we fed to the copepods, is rich in 16:0, 16:1n-7 and EPA59. The concentrations of 16:0 and 16:1n-7 increased, whereas EPA concentration decreased. If fatty acid concentrations reflected feeding, we would have seen increased concentrations of all three. We therefore believe that the general increases in concentrations of free fatty acids were caused by increasing catabolism of the wax esters stored in stage CV. It may be that due to the metabolic reconfiguration to enter hibernation, stage CV copepodites are already committed to the catabolism of fatty acids through beta-oxidation, and stored wax esters are being hydrolysed to increase the availability of free fatty acids for energy production. Mayor and colleagues57 arrived at the same conclusion. We hypothesise that stress due to low pH increases the organism’s energetic demands, but carbohydrates are not used to accommodate these demands due to the down-regulation of the glycolysis, rather demands are met by hydrolysing and metabolising wax esters in stage CV. The further ramifications of future OA could therefore be a less efficient build-up of wax esters so important for hibernation in this species.Finally, besides their importance for cell membrane fluidity, EPA and DHA are important precursors for eicosanoid endocrine hormones. These hormones are important regulators of, among other processes, ion flux62. As mentioned above, acid base regulation is coupled to osmoregulation in crustaceans42, and the decrease in concentrations of these two specific fatty acids, when all other fatty acid concentrations increased might represent an indication for changing endocrine hormone production to counter adverse whole-organism effects of OA.Changes in metabolite concentrations cannot be directly translated into changes in the rate of the processes they are involved in. However, they do pin-point processes which are affected by the imposed environmental changes. Also, in our analyses we targeted a limited range of molecules. In that respect OA could inflict changes in other important metabolic pathways we did not investigate. The absence of specific biochemical pathways in our analyses and discussion should therefore not be taken as indication that these are not implicated in this species responses to OA.From our previously published study on copepodites from the same incubations, we know that high pCO2/low pH conditions have detrimental effects on the balance between energy input (ingestion) and energy expenditure (metabolism) in stage CIV copepodites but not in stage CV copepodites20. The effects we report here help in this sense to shed light on the metabolic origin of the rather severe effects on energy balance we observed in stage CIV copepodites and the difference in response between stage CIV and CV20. Copepods develop through six nauplii and five copepodite stages before maturation, and while previous studies show negligible effects in stage CV and adults17,18,20, any effects in any developmental stage along the way will affect the fitness of the individual and the recruitment to the population as a whole. In addition, the enhanced fatty acid metabolism observed in stage CV needs further investigation, to determine the magnitude of the fitness implications of the energy diverted away from energy storage for hibernation. More

  • in

    Effects of phytoplankton, viral communities, and warming on free-living and particle-associated marine prokaryotic community structure

    Azam, F. et al. The ecological role of water-column microbes in the sea. Marine Ecol. Prog. Ser. 10, 257–263 (1983).Fuhrman, J. A. & Caron D. A. in Manual of Environmental Microbiology (eds Yates, M. V. et al.) 4.2.2–4.2.2.-34 (ASM Press, 2016).Gasol, J. M. & Kirchman, D. L. Microbial Ecology of the Oceans (John Wiley & Sons, 2018).Fuhrman, J. A. et al. A latitudinal diversity gradient in planktonic marine bacteria. Proc. Natl Acad. Sci. 105, 7774–7778 (2008).Article 
    ADS 
    CAS 

    Google Scholar 
    Gilbert, J. A. et al. The seasonal structure of microbial communities in the Western English Channel. Environ. Microbiol. 11, 3132–3139 (2009).Article 
    CAS 

    Google Scholar 
    Gilbert, J. A. et al. Defining seasonal marine microbial community dynamics. ISME J. 6, 298–308 (2012).Article 
    CAS 

    Google Scholar 
    Hatosy, S. M. et al. Beta diversity of marine bacteria depends on temporal scale. Ecology 94, 1898–1904 (2013).Article 

    Google Scholar 
    Ward, C. S. et al. Annual community patterns are driven by seasonal switching between closely related marine bacteria. ISME J. 11, 1412–1422 (2017).Article 

    Google Scholar 
    Fuhrman, J. A. et al. Annually reoccurring bacterial communities are predictable from ocean conditions. Proc. Natl Acad. Sci. 103, 13104–13109 (2006).Article 
    ADS 
    CAS 

    Google Scholar 
    Gonzalez, J. M., Sherr, E. B. & Sherr, B. F. Size-selective grazing on bacteria by natural assemblages of estuarine flagellates and ciliates. Appl. Environ. Microbiol. 56, 583–589 (1990).Article 
    ADS 
    CAS 

    Google Scholar 
    Guixa-Boixereu, N., Vaque, D., Gasol, J. M. & Pedros-Alio, C. Distribution of viruses and their potential effect on bacterioplankton in an oligotrophic marine system. Aquat. Microb. Ecol. 19, 205–213 (1999).Article 

    Google Scholar 
    Šimek, K. et al. Shifts in bacterial community composition associated with different microzooplankton size fractions in a eutrophic reservoir. Limnol. Oceanogr. 44, 1634–1644 (1999).Article 
    ADS 

    Google Scholar 
    Hewson, I., Vargo, G. & Fuhrman, J. Bacterial diversity in shallow oligotrophic marine benthos and overlying waters: effects of virus infection, containment, and nutrient enrichment. Microb. Ecol. 46, 322–336 (2003).Article 
    CAS 

    Google Scholar 
    Schwalbach, M. S., Hewson, I. & Fuhrman, J. A. Viral effects on bacterial community composition in marine plankton microcosms. Aquat. Microb. Ecol. 34, 117–127 (2004).Article 

    Google Scholar 
    Winter, C., Smit, A., Herndl, G. J. & Weinbauer, M. G. Linking bacterial richness with viral abundance and prokaryotic activity. Limnol. Oceanogr. 50, 968–977 (2005).Article 
    ADS 

    Google Scholar 
    Chow, C.-E. T., Kim, D. Y., Sachdeva, R., Caron, D. A. & Fuhrman, J. A. Top-down controls on bacterial community structure: microbial network analysis of bacteria, T4-like viruses and protists. ISME J. 8, 816–829 (2014).Article 
    CAS 

    Google Scholar 
    Suzuki, S. et al. Comparison of community structures between particle-associated and free-living prokaryotes in tropical and subtropical Pacific Ocean surface waters. J. Oceanogr. 73, 383–395 (2017).Article 
    CAS 

    Google Scholar 
    Milici, M. et al. Diversity and community composition of particle‐associated and free‐living bacteria in mesopelagic and bathypelagic Southern Ocean water masses: evidence of dispersal limitation in the Bransfield Strait. Limnol. Oceanogr. 62, 1080–1095 (2017).Article 
    ADS 

    Google Scholar 
    D’ambrosio, L., Ziervogel, K., MacGregor, B., Teske, A. & Arnosti, C. Composition and enzymatic function of particle-associated and free-living bacteria: a coastal/offshore comparison. ISME J. 8, 2167–2179 (2014).Article 

    Google Scholar 
    Rieck, A., Herlemann, D. P., Jürgens, K. & Grossart, H.-P. Particle-associated differ from free-living bacteria in surface waters of the Baltic Sea. Front. Microbiol. 6, 1297 (2015).Article 

    Google Scholar 
    Yung, C.-M., Ward, C. S., Davis, K. M., Johnson, Z. I. & Hunt, D. E. Insensitivity of diverse and temporally variable particle-associated microbial communities to bulk seawater environmental parameters. Appl. Environ. Microbiol. 82, 3431–3437 (2016).Article 
    ADS 
    CAS 

    Google Scholar 
    Buchan, A., LeCleir, G. R., Gulvik, C. A. & González, J. M. Master recyclers: features and functions of bacteria associated with phytoplankton blooms. Nat. Rev. Microbiol. 12, 686–698 (2014).Article 
    CAS 

    Google Scholar 
    Duret, M. T., Lampitt, R. S. & Lam, P. Prokaryotic niche partitioning between suspended and sinking marine particles. Environ. Microbiol. Rep. 11, 386–400 (2019).Article 
    CAS 

    Google Scholar 
    Crespo, B. G., Pommier, T., Fernández‐Gómez, B. & Pedrós‐Alió, C. Taxonomic composition of the particle‐attached and free‐living bacterial assemblages in the Northwest Mediterranean Sea analyzed by pyrosequencing of the 16S rRNA. Microbiologyopen 2, 541–552 (2013).Article 
    CAS 

    Google Scholar 
    Mestre, M., Borrull, E., Sala, M. & Gasol, J. M. Patterns of bacterial diversity in the marine planktonic particulate matter continuum. ISME J. 11, 999–1010 (2017).Yeh, Y. C. et al. Comprehensive single‐PCR 16S and 18S rRNA community analysis validated with mock communities, and estimation of sequencing bias against 18S. Environ. Microbiol. 23, 3240–3250 (2021).Parada, A. E., Needham, D. M. & Fuhrman, J. A. Every base matters: assessing small subunit rRNA primers for marine microbiomes with mock communities, time series and global field samples. Environ. Microbiol. 18, 1403–1414 (2016).Article 
    CAS 

    Google Scholar 
    Needham, D. M. et al. Dynamics and interactions of highly resolved marine plankton via automated high-frequency sampling. ISME J. 12, 2417 (2018).Article 
    CAS 

    Google Scholar 
    McNichol, J., Berube, P. M., Biller, S. J. & Fuhrman, J. A. Evaluating and improving small subunit rRNA PCR primer coverage for bacteria, archaea, and eukaryotes using metagenomes from global ocean surveys. Msystems 6, e00565–00521 (2021).Article 
    CAS 

    Google Scholar 
    Chow, C. E. T. & Fuhrman, J. A. Seasonality and monthly dynamics of marine myovirus communities. Environ. Microbiol. 14, 2171–2183 (2012).Article 

    Google Scholar 
    Filée, J., Tétart, F., Suttle, C. A. & Krisch, H. Marine T4-type bacteriophages, a ubiquitous component of the dark matter of the biosphere. Proc. Natl Acad. Sci. 102, 12471–12476 (2005).Article 
    ADS 

    Google Scholar 
    Pagarete, A. et al. Strong seasonality and interannual recurrence in marine myovirus communities. Appl. Environ. Microbiol. 79, 6253–6259 (2013).Article 
    ADS 
    CAS 

    Google Scholar 
    Comeau, A. M. & Krisch, H. M. The capsid of the T4 phage superfamily: the evolution, diversity, and structure of some of the most prevalent proteins in the biosphere. Mol. Biol. Evolution 25, 1321–1332 (2008).Article 
    CAS 

    Google Scholar 
    Needham, D. M. et al. Short-term observations of marine bacterial and viral communities: patterns, connections and resilience. ISME J. 7, 1274–1285 (2013).Article 
    CAS 

    Google Scholar 
    Needham, D. M., Sachdeva, R. & Fuhrman, J. A. Ecological dynamics and co-occurrence among marine phytoplankton, bacteria and myoviruses shows microdiversity matters. ISME J. 11, 1614–1629 (2017).Article 

    Google Scholar 
    Ahlgren, N. A., Perelman, J. N., Yeh, Y. C. & Fuhrman, J. A. Multi‐year dynamics of fine‐scale marine cyanobacterial populations are more strongly explained by phage interactions than abiotic, bottom‐up factors. Environ. Microbiol. 21, 2948–2963 (2019).Article 
    CAS 

    Google Scholar 
    Ren, J., Ahlgren, N. A., Lu, Y. Y., Fuhrman, J. A. & Sun, F. VirFinder: a novel k-mer based tool for identifying viral sequences from assembled metagenomic data. Microbiome 5, 1–20 (2017).Article 

    Google Scholar 
    Roux, S., Enault, F., Hurwitz, B. L. & Sullivan, M. B. VirSorter: mining viral signal from microbial genomic data. PeerJ 3, e985 (2015).Article 

    Google Scholar 
    Ignacio-Espinoza, J. C., Ahlgren, N. A. & Fuhrman, J. A. Long-term stability and Red Queen-like strain dynamics in marine viruses. Nat. Microbiol. 5, 265–271 (2020).Article 
    CAS 

    Google Scholar 
    Brum, J. R. et al. Patterns and ecological drivers of ocean viral communities. Science 348, (2015).Brown, M. V. et al. Global biogeography of SAR11 marine bacteria. Mol. Syst. Biol. 8, 595 (2012).Article 
    ADS 

    Google Scholar 
    Johnson, Z. I. et al. Niche partitioning among Prochlorococcus ecotypes along ocean-scale environmental gradients. Science 311, 1737–1740 (2006).Article 
    ADS 
    CAS 

    Google Scholar 
    Zwirglmaier, K. et al. Global phylogeography of marine Synechococcus and Prochlorococcus reveals a distinct partitioning of lineages among oceanic biomes. Environ. Microbiol. 10, 147–161 (2008).
    Google Scholar 
    Martiny, A. C., Tai, A. P., Veneziano, D., Primeau, F. & Chisholm, S. W. Taxonomic resolution, ecotypes and the biogeography of Prochlorococcus. Environ. Microbiol. 11, 823–832 (2009).Article 

    Google Scholar 
    Bond, N. A., Cronin, M. F., Freeland, H. & Mantua, N. Causes and impacts of the 2014 warm anomaly in the NE Pacific. Geophys. Res. Lett. 42, 3414–3420 (2015).Article 
    ADS 

    Google Scholar 
    Di Lorenzo, E. & Mantua, N. Multi-year persistence of the 2014/15 North Pacific marine heatwave. Nat. Clim. Change 6, 1042–1047 (2016).Article 
    ADS 

    Google Scholar 
    Traving, S. J. et al. Prokaryotic responses to a warm temperature anomaly in northeast subarctic Pacific waters. Commun. Biol. 4, 1–12 (2021).Article 

    Google Scholar 
    Peña, M. A., Nemcek, N. & Robert, M. Phytoplankton responses to the 2014–2016 warming anomaly in the northeast subarctic Pacific Ocean. Limnol. Oceanogr. 64, 515–525 (2019).Article 
    ADS 

    Google Scholar 
    Yang, B., Emerson, S. R. & Peña, M. A. The effect of the 2013–2016 high temperature anomaly in the subarctic Northeast Pacific (the “Blob”) on net community production. Biogeosciences 15, 6747–6759 (2018).Article 
    ADS 
    CAS 

    Google Scholar 
    Cavole, L. M. et al. Biological impacts of the 2013–2015 warm-water anomaly in the Northeast Pacific: winners, losers, and the future. Oceanography 29, 273–285 (2016).Article 

    Google Scholar 
    Callahan, B. J. et al. DADA2: high-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).Article 
    CAS 

    Google Scholar 
    Needham, D. M. & Fuhrman, J. A. Pronounced daily succession of phytoplankton, archaea and bacteria following a spring bloom. Nat. Microbiol. 1, 16005 (2016).Article 
    CAS 

    Google Scholar 
    Grossart, H. P., Levold, F., Allgaier, M., Simon, M. & Brinkhoff, T. Marine diatom species harbour distinct bacterial communities. Environ. Microbiol. 7, 860–873 (2005).Article 
    CAS 

    Google Scholar 
    Teeling, H. et al. Substrate-controlled succession of marine bacterioplankton populations induced by a phytoplankton bloom. Science 336, 608–611 (2012).Article 
    ADS 
    CAS 

    Google Scholar 
    Chafee, M. et al. Recurrent patterns of microdiversity in a temperate coastal marine environment. ISME J. 12, 237–252 (2018).Article 

    Google Scholar 
    Teeling, H. et al. Recurring patterns in bacterioplankton dynamics during coastal spring algae blooms. elife 5, e11888 (2016).Article 

    Google Scholar 
    Unfried, F. et al. Adaptive mechanisms that provide competitive advantages to marine bacteroidetes during microalgal blooms. ISME J. 12, 2894–2906 (2018).Article 
    CAS 

    Google Scholar 
    Francis, T. B. et al. Changing expression patterns of TonB-dependent transporters suggest shifts in polysaccharide consumption over the course of a spring phytoplankton bloom. ISME J. 15, 2336–2350 (2021).Thingstad, T. F. & Lignell, R. Theoretical models for the control of bacterial growth rate, abundance, diversity and carbon demand. Aquat. Microb. Ecol. 13, 19–27 (1997).Article 

    Google Scholar 
    Thingstad, T. F., Våge, S., Storesund, J. E., Sandaa, R.-A. & Giske, J. A theoretical analysis of how strain-specific viruses can control microbial species diversity. Proc. Natl Acad. Sci. 111, 7813–7818 (2014).Article 
    ADS 
    CAS 

    Google Scholar 
    Thingstad, T. F., Pree, B., Giske, J. & Våge, S. What difference does it make if viruses are strain-, rather than species-specific? Front. Microbiol. 6, 320 (2015).Article 

    Google Scholar 
    Prokopowich, C. D., Gregory, T. R. & Crease, T. J. The correlation between rDNA copy number and genome size in eukaryotes. Genome 46, 48–50 (2003).Article 
    CAS 

    Google Scholar 
    Zhu, F., Massana, R., Not, F., Marie, D. & Vaulot, D. Mapping of picoeucaryotes in marine ecosystems with quantitative PCR of the 18S rRNA gene. FEMS Microbiol. Ecol. 52, 79–92 (2005).Article 
    CAS 

    Google Scholar 
    Sintes, E. & Del Giorgio, P. A. Feedbacks between protistan single-cell activity and bacterial physiological structure reinforce the predator/prey link in microbial foodwebs. Front. Microbiol. 5, 453 (2014).Article 

    Google Scholar 
    Del Giorgio, P. A. et al. Bacterioplankton community structure: protists control net production and the proportion of active bacteria in a coastal marine community. Limnol. Oceanogr. 41, 1169–1179 (1996).Article 
    ADS 

    Google Scholar 
    Andersson, A., Larsson, U. & Hagström, Å. Size-selective grazing by a microflagellate on pelagic bacteria. Marine Ecol. Prog. Ser. 33, 51–57 (1986).Pernthaler, J. Predation on prokaryotes in the water column and its ecological implications. Nat. Rev. Microbiol. 3, 537–546 (2005).Article 
    CAS 

    Google Scholar 
    Baltar, F. et al. Marine bacterial community structure resilience to changes in protist predation under phytoplankton bloom conditions. ISME J. 10, 568–581 (2016).Article 

    Google Scholar 
    Suzuki, M. T. Effect of protistan bacterivory on coastal bacterioplankton diversity. Aquat. Microb. Ecol. 20, 261–272 (1999).Article 

    Google Scholar 
    Yokokawa, T. & Nagata, T. Growth and grazing mortality rates of phylogenetic groups of bacterioplankton in coastal marine environments. Appl. Environ. Microbiol. 71, 6799–6807 (2005).Article 
    ADS 
    CAS 

    Google Scholar 
    Eren, A. M. et al. Oligotyping: differentiating between closely related microbial taxa using 16S rRNA gene data. Methods Ecol. Evolution 4, 1111–1119 (2013).Article 

    Google Scholar 
    Coleman, M. L. & Chisholm, S. W. Code and context: Prochlorococcus as a model for cross-scale biology. Trends Microbiol. 15, 398–407 (2007).Article 
    CAS 

    Google Scholar 
    Scanlan, D. J. et al. Ecological genomics of marine picocyanobacteria. Microbiol. Mol. Biol. Rev. 73, 249–299 (2009).Article 
    CAS 

    Google Scholar 
    Moore, L. R., Rocap, G. & Chisholm, S. W. Physiology and molecular phylogeny of coexisting Prochlorococcus ecotypes. Nature 393, 464–467 (1998).Article 
    ADS 
    CAS 

    Google Scholar 
    Rusch, D. B., Martiny, A. C., Dupont, C. L., Halpern, A. L. & Venter, J. C. Characterization of Prochlorococcus clades from iron-depleted oceanic regions. Proc. Natl Acad. Sci. 107, 16184–16189 (2010).Article 
    ADS 
    CAS 

    Google Scholar 
    Larkin, A. A. et al. Persistent El Niño driven shifts in marine cyanobacteria populations. PloS ONE 15, e0238405 (2020).Article 
    CAS 

    Google Scholar 
    Arandia‐Gorostidi, N. et al. Warming the phycosphere: differential effect of temperature on the use of diatom‐derived carbon by two copiotrophic bacterial taxa. Environ. Microbiol. 22, 1381–1396 (2020).Article 

    Google Scholar 
    Arandia‐Gorostidi, N., Huete‐Stauffer, T. M., Alonso‐Sáez L, G. & Morán, X. A. Testing the metabolic theory of ecology with marine bacteria: different temperature sensitivity of major phylogenetic groups during the spring phytoplankton bloom. Environ. Microbiol. 19, 4493–4505 (2017).Article 

    Google Scholar 
    Fagan, A. J., Moreno, A. R. & Martiny, A. C. Role of ENSO conditions on particulate organic matter concentrations and elemental ratios in the Southern California Bight. Front. Mar. Sci. 6, 386 (2019).Article 

    Google Scholar 
    Chang, C. W. et al. Reconstructing large interaction networks from empirical time series data. Ecol. Lett. 24, 2763–2774 (2021).Article 

    Google Scholar 
    Lie, A. A., Kim, D. Y., Schnetzer, A. & Caron, D. A. Small-scale temporal and spatial variations in protistan community composition at the San Pedro Ocean Time-series station off the coast of southern California. Aquat. Microb. Ecol. 70, 93–110 (2013).Article 

    Google Scholar 
    Yeh, Y.-C., Needham, D. M., Sieradzki, E. T. & Fuhrman, J. A. Taxon disappearance from microbiome analysis reinforces the value of mock communities as a standard in every sequencing run. MSystems 3, e00023–00018 (2018).Article 

    Google Scholar 
    Caporaso, J. G. et al. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 7, 335–336 (2010).Article 
    CAS 

    Google Scholar 
    Quast, C. et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 41, D590–D596 (2013).Article 
    CAS 

    Google Scholar 
    Guillou, L. et al. The Protist Ribosomal Reference database (PR2): a catalog of unicellular eukaryote small sub-unit rRNA sequences with curated taxonomy. Nucleic Acids Res. 41, D579–D604 (2013).
    Google Scholar 
    Amir, A. et al. Deblur rapidly resolves single-nucleotide community sequence patterns. MSystems 2, (2017).Decelle, J. et al. Phyto REF: a reference database of the plastidial 16S rRNA gene of photosynthetic eukaryotes with curated taxonomy. Mol. Ecol. Resour. 15, 1435–1445 (2015).Article 
    CAS 

    Google Scholar 
    Amin, S. et al. Interaction and signalling between a cosmopolitan phytoplankton and associated bacteria. Nature 522, 98–101 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    Legendre, P. & Gallagher, E. D. Ecologically meaningful transformations for ordination of species data. Oecologia 129, 271–280 (2001).Article 
    ADS 

    Google Scholar 
    Hill, M. O. & Gauch, H. G. J. Detrended correspondence analysis: an improved ordination technique. Vegetatio 42, 47–58 (1980).Ter Braak, C. J. Canonical correspondence analysis: a new eigenvector technique for multivariate direct gradient analysis. Ecology 67, 1167–1179 (1986).Article 

    Google Scholar 
    Peres-Neto, P. R., Legendre, P., Dray, S. & Borcard, D. Variation partitioning of species data matrices: estimation and comparison of fractions. Ecology 87, 2614–2625 (2006).Article 

    Google Scholar  More

  • in

    A regulatory hydrogenase gene cluster observed in the thioautotrophic symbiont of Bathymodiolus mussel in the East Pacific Rise

    Sogin, E. M., Leisch, N. & Dubilier, N. Chemosynthetic symbioses. Curr. Biol. 30, R1137–R1142 (2020).Article 
    CAS 

    Google Scholar 
    Dubilier, N., Bergin, C. & Lott, C. Symbiotic diversity in marine animals: The art of harnessing chemosynthesis. Nat. Rev. Microbiol. 6, 725–740 (2008).Article 
    CAS 

    Google Scholar 
    Barry, J. P. et al. Methane-based symbiosis in a mussel, Bathymodiolus platifrons, from cold seeps in Sagami Bay Japan. Invertebr. Biol. 121, 47–54 (2002).Article 

    Google Scholar 
    Le Pennec, M., Donval, A. & Herry, A. Nutritional strategies of the hydrothermal ecosystem bivalves. Prog. Oceanogr. 24, 71–80 (1990).Article 
    ADS 

    Google Scholar 
    Rau, G. H. & Hedges, J. I. Carbon-13 depletion in a hydrothermal vent mussel: Suggestion of a chemosynthetic food source. Science 203, 648–649 (1979).Article 
    ADS 
    CAS 

    Google Scholar 
    Wentrup, C., Wendeberg, A., Schimak, M., Borowski, C. & Dubilier, N. Forever competent: Deep-sea bivalves are colonized by their chemosynthetic symbionts throughout their lifetime. Environ. Microbiol. 16, 3699–3713 (2014).Article 

    Google Scholar 
    Dattagupta, S., Bergquist, D., Szalai, E., Macko, S. & Fisher, C. Tissue carbon, nitrogen, and sulfur stable isotope turnover in transplanted Bathymodiolus childressi mussels: Relation to growth and physiological condition. Limnol. Oceanogr. 49, 1144–1151 (2004).Article 
    ADS 
    CAS 

    Google Scholar 
    Ikuta, T. et al. Heterogeneous composition of key metabolic gene clusters in a vent mussel symbiont population. ISME J. 10, 990–1001 (2016).Article 

    Google Scholar 
    Takishita, K. et al. Genomic evidence that methanotrophic endosymbionts likely provide deep-sea Bathymodiolus mussels with a sterol intermediate in cholesterol biosynthesis. Genome Biol. Evol. 9, 1148–1160 (2017).Article 

    Google Scholar 
    Sayavedra, L. et al. Horizontal acquisition followed by expansion and diversification of toxin-related genes in deep-sea bivalve symbionts. BioRxiv 110, 330 (2019).
    Google Scholar 
    Ponnudurai, R. et al. Metabolic and physiological interdependencies in the Bathymodiolus azoricus symbiosis. ISME J. 11, 463–477 (2017).Article 
    CAS 

    Google Scholar 
    Ponnudurai, R. et al. Genome sequence of the sulfur-oxidizing Bathymodiolus thermophilus gill endosymbiont. Stand Genom. Sci. 12, 1–9 (2017).
    Google Scholar 
    Kiel, S. The Vent and Seep Biota: Aspects from Microbes to Ecosystems Vol. 33 (Springer Science & Business Media, 2010).
    Google Scholar 
    Lorion, J. et al. Adaptive radiation of chemosymbiotic deep-sea mussels. Proc. R. Soc. B 280, 20131243 (2013).Article 

    Google Scholar 
    Nussbaumer, A. D., Fisher, C. R. & Bright, M. Horizontal endosymbiont transmission in hydrothermal vent tubeworms. Nature 441, 345–348 (2006).Article 
    ADS 
    CAS 

    Google Scholar 
    Gros, O., Liberge, M., Heddi, A., Khatchadourian, C. & Felbeck, H. Detection of the free-living forms of sulfide-oxidizing gill endosymbionts in the lucinid habitat (Thalassia testudinum environment). Appl. Environ. Microbiol. 69, 6264–6267 (2003).Article 
    ADS 
    CAS 

    Google Scholar 
    Won, Y.-J. et al. Environmental acquisition of thiotrophic endosymbionts by deep-sea mussels of the genus Bathymodiolus. Appl. Environ. Microbiol. 69, 6785–6792 (2003).Article 
    ADS 
    CAS 

    Google Scholar 
    Laming, S. R., Gaudron, S. M. & Duperron, S. Lifecycle ecology of deep-sea chemosymbiotic mussels: A review. Front. Mar. Sci. 5, 282 (2018).Article 

    Google Scholar 
    Laming, S. R., Duperron, S., Cunha, M. R. & Gaudron, S. M. Settled, symbiotic, then sexually mature: Adaptive developmental anatomy in the deep-sea, chemosymbiotic mussel Idas modiolaeformis. Mar. Biol. 161, 1319–1333 (2014).Article 

    Google Scholar 
    Salerno, J. L. et al. Characterization of symbiont populations in life-history stages of mussels from chemosynthetic environments. Biol. Bull. 208, 145–155 (2005).Article 

    Google Scholar 
    Wentrup, C., Wendeberg, A., Huang, J. Y., Borowski, C. & Dubilier, N. Shift from widespread symbiont infection of host tissues to specific colonization of gills in juvenile deep-sea mussels. ISME J. 7, 1244–1247 (2013).Article 
    CAS 

    Google Scholar 
    Pennec, M. L. & Beninger, P. G. Ultrastructural characteristics of spermatogenesis in three species of deep-sea hydrothermal vent mytilids. Can. J. Zool. 75, 308–316 (1997).Article 

    Google Scholar 
    Eckelbarger, K. & Young, C. Ultrastructure of gametogenesis in a chemosynthetic mytilid bivalve (Bathymodiolus childressi) from a bathyal, methane seep environment (northern Gulf of Mexico). Mar. Biol. 135, 635–646 (1999).Article 

    Google Scholar 
    Ansorge, R. et al. Diversity matters: Deep-sea mussels harbor multiple symbiont strains. bioRxiv 99, 1039 (2019).
    Google Scholar 
    Petersen, J. M., Wentrup, C., Verna, C., Knittel, K. & Dubilier, N. Origins and evolutionary flexibility of chemosynthetic symbionts from deep-sea animals. Biol. Bull. 223, 123–137 (2012).Article 
    CAS 

    Google Scholar 
    Sayavedra, L. et al. Abundant toxin-related genes in the genomes of beneficial symbionts from deep-sea hydrothermal vent mussels. Elife 4, e07966 (2015).Article 

    Google Scholar 
    Ansorge, R. et al. Functional diversity enables multiple symbiont strains to coexist in deep-sea mussels. Nat. Microbiol. 4, 2487–2497 (2019).Article 

    Google Scholar 
    Petersen, J. M. et al. Hydrogen is an energy source for hydrothermal vent symbioses. Nature 476, 176–180 (2011).Article 
    ADS 
    CAS 

    Google Scholar 
    Nakamura, K. & Takai, K. Theoretical constraints of physical and chemical properties of hydrothermal fluids on variations in chemolithotrophic microbial communities in seafloor hydrothermal systems. Prog. Earth Planet Sci. 1, 1–24 (2014).Article 
    ADS 

    Google Scholar 
    Perez, M. & Juniper, S. K. Insights into symbiont population structure among three vestimentiferan tubeworm host species at eastern Pacific spreading centers. Appl. Environ. Microbiol. 82, 5197–5205 (2016).Article 
    ADS 
    CAS 

    Google Scholar 
    Wilbanks, E. G. et al. Metagenomic methylation patterns resolve bacterial genomes of unusual size and structural complexity. ISME J. https://doi.org/10.1038/s41396-022-01242-7 (2022).Article 

    Google Scholar 
    Rodriguez-Casariego, J. A., Cunning, R., Baker, A. C. & Eirin-Lopez, J. M. Symbiont shuffling induces differential DNA methylation responses to thermal stress in the coral Montastraea cavernosa. Mol. Ecol. 31, 588–602 (2022).Article 
    CAS 

    Google Scholar 
    Triant, D. A. & Whitehead, A. Simultaneous extraction of high-quality RNA and DNA from small tissue samples. J. Hered. 100, 246–250 (2009).Article 
    CAS 

    Google Scholar 
    Chin, C.-S. et al. Nonhybrid, finished microbial genome assemblies from long-read SMRT sequencing data. Nat. Methods 10, 563–569 (2013).Article 
    CAS 

    Google Scholar 
    Wick, R. R. et al. Trycycler: Consensus long-read assemblies for bacterial genomes. Genome Biol. 22, 1–17 (2021).Article 

    Google Scholar 
    Kolmogorov, M., Yuan, J., Lin, Y. & Pevzner, P. A. Assembly of long, error-prone reads using repeat graphs. Nat. Biotechnol. 37, 540–546 (2019).Article 
    CAS 

    Google Scholar 
    Wick, R. R., Judd, L. M., Gorrie, C. L. & Holt, K. E. Unicycler: Resolving bacterial genome assemblies from short and long sequencing reads. PLoS Comput. Biol. 13, e1005595 (2017).Article 
    ADS 

    Google Scholar 
    Krawczyk, P. S., Lipinski, L. & Dziembowski, A. PlasFlow: Predicting plasmid sequences in metagenomic data using genome signatures. Nucleic Acids Res. 46, e35–e35 (2018).Article 

    Google Scholar 
    Mikheenko, A., Prjibelski, A., Saveliev, V., Antipov, D. & Gurevich, A. Versatile genome assembly evaluation with QUAST-LG. Bioinformatics 34, i142–i150 (2018).Article 
    CAS 

    Google Scholar 
    Parks, D. H., Imelfort, M., Skennerton, C. T., Hugenholtz, P. & Tyson, G. W. CheckM: Assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 25, 1043–1055 (2015).Article 
    CAS 

    Google Scholar 
    Couvin, D. et al. CRISPRCasFinder, an update of CRISRFinder, includes a portable version, enhanced performance and integrates search for Cas proteins. Nucleic Acids Res. 46, W246–W251 (2018).Article 
    CAS 

    Google Scholar 
    Perez, M., Angers, B., Young, C. R. & Juniper, S. K. Shining light on a deep-sea bacterial symbiont population structure with CRISPR. Microbial. Genom. https://doi.org/10.1099/mgen.0.000625 (2021).Article 

    Google Scholar 
    Hyatt, D. et al. Prodigal: Prokaryotic gene recognition and translation initiation site identification. BMC Bioinform. 11, 1–11 (2010).Article 

    Google Scholar 
    Nielsen, H. Protein Function Prediction 59–73 (Springer, 2017).Book 

    Google Scholar 
    Krogh, A., Larsson, B., Von Heijne, G. & Sonnhammer, E. L. Predicting transmembrane protein topology with a hidden Markov model: Application to complete genomes. J. Mol. Biol. 305, 567–580 (2001).Article 
    CAS 

    Google Scholar 
    Lagesen, K. et al. RNAmmer: Consistent and rapid annotation of ribosomal RNA genes. Nucleic Acids Res. 35, 3100-31C08 (2007).Article 
    ADS 
    CAS 

    Google Scholar 
    Chan, P. P. & Lowe, T. M. Gene Prediction 1–14 (Springer, 2019).
    Google Scholar 
    Griffiths-Jones, S. et al. Rfam: Annotating non-coding RNAs in complete genomes. Nucleic Acids Res. 33, D121–D124 (2005).Article 
    CAS 

    Google Scholar 
    Kanehisa, M., Sato, Y. & Morishima, K. BlastKOALA and GhostKOALA: KEGG tools for functional characterization of genome and metagenome sequences. J. Mol. Biol. 428, 726–731 (2016).Article 
    CAS 

    Google Scholar 
    Siguier, P., Pérochon, J., Lestrade, L., Mahillon, J. & Chandler, M. ISfinder: the reference centre for bacterial insertion sequences. Nucleic Acids Res. 34, D32–D36 (2006).Article 
    CAS 

    Google Scholar 
    Bertelli, C. et al. IslandViewer 4: Expanded prediction of genomic islands for larger-scale datasets. Nucleic Acids Res. 45, W30–W35 (2017).Article 
    CAS 

    Google Scholar 
    Arndt, D. et al. PHASTER: A better, faster version of the PHAST phage search tool. Nucleic Acids Res. 44, W16–W21 (2016).Article 
    CAS 

    Google Scholar 
    Roeselers, G. et al. Complete genome sequence of Candidatus Ruthia magnifica. Stand Genomic Sci. 3, 163–173 (2010).Article 

    Google Scholar 
    Emms, D. M. & Kelly, S. OrthoFinder: Phylogenetic orthology inference for comparative genomics. Genome Biol. 20, 1–14 (2019).Article 

    Google Scholar 
    Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).Article 

    Google Scholar 
    Minh, B. Q. et al. IQ-TREE 2: New models and efficient methods for phylogenetic inference in the genomic era. Mol. Biol. Evol. 37, 1530–1534 (2020).Article 
    CAS 

    Google Scholar 
    Kumar, S., Stecher, G., Li, M., Knyaz, C. & Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 35, 1547 (2018).Article 
    CAS 

    Google Scholar 
    Letunic, I. & Bork, P. Interactive tree of life (iTOL) v4: Recent updates and new developments. Nucleic Acids Res. 47, W256–W259 (2019).Article 
    CAS 

    Google Scholar 
    Eren, A. M. et al. Community-led, integrated, reproducible multi-omics with anvi’o. Nat. Microbiol. 6, 3–6 (2021).Article 
    CAS 

    Google Scholar 
    Darling, A. E., Mau, B. & Perna, N. T. progressiveMauve: Multiple genome alignment with gene gain, loss and rearrangement. PLoS ONE 5, e11147 (2010).Article 
    ADS 

    Google Scholar 
    Tesler, G. GRIMM: Genome rearrangements web server. Bioinformatics 18, 492–493 (2002).Article 
    CAS 

    Google Scholar 
    Cabanettes, F. & Klopp, C. D-GENIES: Dot plot large genomes in an interactive, efficient and simple way. PeerJ 6, e4958 (2018).Article 

    Google Scholar 
    Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 215, 403–410 (1990).Article 
    CAS 

    Google Scholar 
    Gilchrist, C. L. & Chooi, Y.-H. Clinker & clustermap. js: Automatic generation of gene cluster comparison figures. Bioinformatics 37, 2473–2475 (2021).Article 
    CAS 

    Google Scholar 
    Taboada, B., Estrada, K., Ciria, R. & Merino, E. Operon-mapper: A web server for precise operon identification in bacterial and archaeal genomes. Bioinformatics 34, 4118–4120 (2018).Article 
    CAS 

    Google Scholar 
    Søndergaard, D., Pedersen, C. N. & Greening, C. HydDB: A web tool for hydrogenase classification and analysis. Sci. Rep. 6, 1–8 (2016).Article 

    Google Scholar 
    NCBI Genome Browser. https://www.ncbi.nlm.nih.gov/genome/browse/#!/prokaryotes/. Accessed 12 March 2022.Mcmullin, E. R., Hourdez, S., Schaeffer, S. W. & Fisher, C. R. Review article phylogeny and biogeography of deep sea vestimentiferan tubeworms and their bacterial symbionts. Symbiosis. 34, 1–41 (2003).
    Google Scholar 
    Won, Y.-J., Jones, W. J. & Vrijenhoek, R. C. Absence of cospeciation between deep-sea mytilids and their thiotrophic endosymbionts. J. Shellfish Res. 27, 129–138 (2008).Article 

    Google Scholar 
    Miyazaki, J.-I., Martins, Ld. O., Fujita, Y., Matsumoto, H. & Fujiwara, Y. Evolutionary process of deep-sea Bathymodiolus mussels. PLoS ONE 5, e10363 (2010).Article 
    ADS 

    Google Scholar 
    Bright, M. & Bulgheresi, S. A complex journey: Transmission of microbial symbionts. Nat. Rev. Microbiol. 8, 218–230 (2010).Article 
    CAS 

    Google Scholar 
    Raggi, L., Schubotz, F., Hinrichs, K. U., Dubilier, N. & Petersen, J. M. Bacterial symbionts of Bathymodiolus mussels and Escarpia tubeworms from Chapopote, an asphalt seep in the southern Gulf of Mexico. Environ. Microbiol. 15, 1969–1987 (2013).Article 
    CAS 

    Google Scholar 
    Goris, J. et al. DNA–DNA hybridization values and their relationship to whole-genome sequence similarities. Int. J. Syst. Evol. Microbiol. 57, 81–91 (2007).Article 
    CAS 

    Google Scholar 
    Meier-Kolthoff, J. P., Auch, A. F., Klenk, H.-P. & Göker, M. Genome sequence-based species delimitation with confidence intervals and improved distance functions. BMC Bioinform. 14, 1–14 (2013).Article 

    Google Scholar 
    Konstantinidis, K. T. & Tiedje, J. M. Genomic insights that advance the species definition for prokaryotes. Proc. Natl. Acad. Sci. 102, 2567–2572 (2005).Article 
    ADS 
    CAS 

    Google Scholar 
    Ho, P.-T. et al. Geographical structure of endosymbiotic bacteria hosted by Bathymodiolus mussels at eastern Pacific hydrothermal vents. BMC Evol. Biol. 17, 1–16 (2017).Article 

    Google Scholar 
    Romero Picazo, D. et al. Horizontally transmitted symbiont populations in deep-sea mussels are genetically isolated. ISME J. 13, 2954–2968 (2019).Article 

    Google Scholar 
    Perez, M. & Juniper, S. K. Is the trophosome of Ridgeia piscesae monoclonal?. Symbiosis 74, 55–65 (2018).Article 
    CAS 

    Google Scholar 
    Polzin, J., Arevalo, P., Nussbaumer, T., Polz, M. F. & Bright, M. Polyclonal symbiont populations in hydrothermal vent tubeworms and the environment. Proc. R. Soc. B 286, 20181281 (2019).Article 
    CAS 

    Google Scholar 
    Russell, S. L. & Cavanaugh, C. M. Intrahost genetic diversity of bacterial symbionts exhibits evidence of mixed infections and recombinant haplotypes. Mol. Biol. Evol. 34, 2747–2761 (2017).Article 
    CAS 

    Google Scholar 
    Breusing, C., Genetti, M., Russell, S. L., Corbett-Detig, R. B. & Beinart, R. A. Horizontal transmission enables flexible associations with locally adapted symbiont strains in deep-sea hydrothermal vent symbioses. Proc. Natl. Acad. Sci. 119, e2115608119 (2022).Article 
    CAS 

    Google Scholar 
    Lan, Y. et al. Endosymbiont population genomics sheds light on transmission mode, partner specificity, and stability of the scaly-foot snail holobiont. ISME J. https://doi.org/10.1038/s41396-022-01261-4 (2022).Article 

    Google Scholar 
    Anantharaman, K., Breier, J. A., Sheik, C. S. & Dick, G. J. Evidence for hydrogen oxidation and metabolic plasticity in widespread deep-sea sulfur-oxidizing bacteria. Proc. Natl. Acad. Sci. 110, 330–335 (2013).Article 
    ADS 
    CAS 

    Google Scholar 
    Fritsch, J. et al. Rubredoxin-related maturation factor guarantees metal cofactor integrity during aerobic biosynthesis of membrane-bound [NiFe] hydrogenase. J. Biol. Chem. 289, 7982–7993 (2014).Article 
    CAS 

    Google Scholar 
    Petersen, J. M. et al. Chemosynthetic symbionts of marine invertebrate animals are capable of nitrogen fixation. Nat. Microbiol. 2, 1–11 (2016).Article 

    Google Scholar 
    Nakagawa, S. et al. Allying with armored snails: The complete genome of gammaproteobacterial endosymbiont. ISME J. 8, 40–51 (2014).Article 
    CAS 

    Google Scholar 
    Vignais, P. M., Billoud, B. & Meyer, J. Classification and phylogeny of hydrogenases. FEMS Microbiol. Rev. 25, 455–501 (2001).Article 
    CAS 

    Google Scholar 
    Perez, M. et al. Divergent paths in the evolutionary history of maternally transmitted clam symbionts. Proc. R. Soc. B 289, 20212137 (2022).Article 
    CAS 

    Google Scholar 
    Li, S. et al. N 4-cytosine DNA methylation is involved in the maintenance of genomic stability in Deinococcus radiodurans. Front. Microbiol. 10, 1905 (2019).Article 

    Google Scholar 
    Casadesús, J. & Low, D. Epigenetic gene regulation in the bacterial world. Microbiol. Mol. Biol. Rev. 70, 830–856 (2006).Article 

    Google Scholar 
    De Oliveira, A. L., Srivastava, A., Espada-Hinojosa, S. & Bright, M. The complete and closed genome of the facultative generalist Candidatus Endoriftia persephone from deep-sea hydrothermal vents. Mol. Ecol. Resour. https://doi.org/10.1111/1755-0998.13668 (2022).Article 

    Google Scholar 
    Ponnudurai, R. et al. Comparative proteomics of related symbiotic mussel species reveals high variability of host–symbiont interactions. ISME J. 14, 649–656 (2020).Article 
    CAS 

    Google Scholar 
    Yu, N. Y. et al. PSORTb 3.0: Improved protein subcellular localization prediction with refined localization subcategories and predictive capabilities for all prokaryotes. Bioinformatics 26, 1608–1615 (2010).Article 
    CAS 

    Google Scholar  More