More stories

  • in

    Storm surge and ponding explain mangrove dieback in southwest Florida following Hurricane Irma

    A combination of airborne and satellite remote sensing data were used to quantify changes in mangrove forest structure and function from Hurricane Irma (Supplementary Fig. 1). Findings based on multi-sensor airborne data were scaled to the entire study area using estimates of forest structure and vegetation phenology derived from satellite data.G-LiHT Airborne campaignDuring April 2017, NASA Goddard’s Lidar, Hyperspectral, and Thermal (G-LiHT) airborne imager conducted an extensive airborne campaign in South Florida covering >130,000 ha. The same flight lines were resurveyed with G-LiHT eight months later, during November and December of 2017, to quantify structural changes in coastal forests of South Florida and Everglades National Park (ENP) following Hurricane Irma (Fig. 1). Lidar data was collected with two VQ-280i (Riegl USA) and synced during flight using RiACQUIRE version 2.3.7. The plane flew at a nominal height of 335 m above ground level at a pulse repetition frequency of 300 kHz to collect ~12 laser pulses per square meter. The analysis of pre- and post-hurricane conditions used 1-m resolution lidar data products (Supplementary Fig. 2) and 3-cm resolution stereo aerial and ground photos to estimate changes in vegetation structure, fractional cover, and terrain heights across the study domain. G-LiHT lidar canopy height models, digital terrain models, and estimates of fractional vegetation canopy cover (FVC) were produced using standard processing methodology21. All Level 1 through 3 lidar data products and fine-resolution imagery are openly shared through the G-LiHT webpage (https://gliht.gsfc.nasa.gov/).High resolution stereo maps of canopy heightStereo imagery from high-resolution commercial satellites can be used to estimate canopy and terrain surfaces42,43. Here, we derived digital surface models (DSMs) from DigitalGlobe’s WorldView 2 Level 1B imagery. DigitalGlobe provides these data to U.S. Government agencies and non-profit organizations that support U.S. interests via the NextView license agreement44. The spatial resolution of these data depends on the degree of off-nadir pointing for each acquisition. In this study, image resolution ranged from 0.5 to 0.7 m. We selected along-track stereopairs within the study domain to identify stereo image strips (each ~17 km × 110 km) that were nominally cloud-free over the forested domain of interest for years 2012–2013, the most recent cloud-free stereo data available for the study region prior to Hurricane Irma. The DSMs were produced using the Ames Stereo Pipeline (ASP) v. 2.5.1 on the NASA Center for Climate Simulation’s Advanced Data Analytics Platform at Goddard Space Flight Center (ADAPT, https://www.nccs.nasa.gov/services/adapt). The Worldview DSMs have been shown to accurately estimate mangrove canopy height when compared to airborne lidar and radar interferometry42,43. The processing workflow was adapted from ref. 45, and was implemented semi-global matching algorithms with a 5 × 5 correlation kernel, and a 3 × 3 median-filter applied to the output point cloud prior to producing a 1 m DSM using a weighted average gridding rule46. The ASP processing yielded five DSMs at 1-m resolution that were used to capture pre-storm canopy surface elevations.Each of the five Worldview DSMs were individually calibrated using overlapping pre-storm G-LiHT lidar data to estimate mangrove canopy heights across the study region (Supplementary Fig. 1). We sampled 1000 points within the mangrove forest cover (see mangrove classification, below) to develop a bias-correction equation between G-LiHT lidar-derived canopy heights and stereo DSM elevations (Supplementary Fig. 6). The bias-corrected canopy height models from high-resolution stereo imagery were mosaicked together to generate a 1-m resolution CHM for the entire study region (Supplementary Fig. 7). A pre-storm canopy volume was calculated by summing the 1 m × 1 m WorldView CHM for the entire region of interest. Similarly, a post-storm canopy volume was derived using the canopy damage model (see the section below), the relationship between the pre-storm CHM and the max wind speed. This analysis was conducted in ArcMap 10.7.1.Landsat mangrove forest classificationLandsat 8 Operational Land Imager (OLI) imagery was used to map mangrove cover for the southern Florida study region. The imagery was preprocessed to surface reflectance47 and clouds were masked following methods outlined in ref. 48. The Surface Reflectance Tier 1 product in Google Earth Engine was used to create a cloud-free image mosaic for 2016 based on the median values of all cloud-free images for the year for all bands (Supplementary Fig. 1).Training points were hand-selected using contemporary Google Earth imagery, field photos, and expert knowledge of the region. Twenty-four polygons covering a mangrove area of 1243 ha and 17 polygons covering a non-mangrove area of 2759 ha were identified for training regions. Within each of the two classes (i.e., mangrove and non-mangrove), 100,00 points were sampled and used for the training data in a Random Forest Classification implemented in Google Earth Engine49. The Random Forest model used 20 trees and a bag fraction of 0.5. The Landsat-based mangrove map was validated using the Region 3 species land cover map developed by the National Park Service for Everglades National Park50. The National Park species map was reclassified into mangrove and non-mangrove land cover, and 500 randomly generated points were sampled within each of the two land cover classes. The resulting error matrix indicated an overall accuracy of 90.6%.Post-storm canopy coverTime series of Landsat data were used to estimate hurricane damages of mangrove forest cover through December 31, 2017. We combined data from Landsat 7 ETM+ and Landsat 8 OLI to create a dense time series of cloud-free observations. All images were pre-processed to surface reflectance and masked for clouds using the same methods as the mangrove classification. Landsat 7 and Landsat 8 data were then harmonized to account for differences in the sensor specifications following51. We calculated the Normalized Difference Vegetation Index (NDVI) and Normalized Difference Water Index (NDWI) for each image in the collection. We calculated two reference maps from the time series of Landsat imagery (Supplementary Fig. 1). A pre-storm reference was calculated as the median value for each reflectance and index band for all cloud-free imagery in the two years prior Hurricane Irma, August 31, 2015 through August 31, 2017. Similarly, a post-storm median mosaic image was made using Landsat data between October 1, 2017 and December 31, 2017.Pre- and post-storm wall-to-wall Fractional Vegetation Cover (FVC) maps were generated using a combination of lidar-based FVC metrics and Landsat imagery (Supplementary Fig. 1). First, lidar-based FVC was binned into five classes; 0–20%, 20–40%, 40–60%, 60–80%, and 80–100% (Supplementary Fig. 7). We then collected 1000 randomly generated points in each of the five FVC classes, a total of 5000 points, to be used as training data in the Landsat classification. Here, we implemented a Random Forest Classifier using 100 trees and a bag fraction of 0.5. These steps were applied to both the pre-storm and post-storm lidar-derived FVC and Landsat mosaic image metrics. Changes between the pre- and post-storm FVC were then calculated based on the five different FVC classes (Supplementary Fig. 7). For example, a pixel with pre-storm FVC of 80–100% and a post-storm FVC of 20–40%, a reduction of three FVC classes, was assigned a drop in FVC of 40–60% (Fig. 1).Recovery times and resilienceWe estimated the time to full recovery of pre-storm mangrove green canopy cover using the time series of Landsat NDVI during the first 15-months following Hurricane Irma. The pre-storm mean NDVI layer was used as a reference, as described in the previous section. Next we calculated the NDVI anomaly for each image during the post-storm period, September 17, 2017 through December 31, 2018 (Supplementary Fig. 1). We then summed the individual anomaly values from each Landsat image and normalized by the total number of valid pixels (i.e., pixels meeting quality control measures) to estimate the average change in NDVI within the 15 months after the storm. We used anomaly values to identify mangrove forests with large decreases in the 15 months after the storm using a threshold of 0.2 for the 15-month NDVI average anomaly19,52. These areas suffered large losses of canopy material and limited new growth during the post-storm period. We used the slope in NDVI values for each pixel during 2018 to estimate the time in years to full recovery to pre-storm NDVI values, excluding data from October to December 2017 to remove delayed browning of damaged vegetation and spurious NDVI values from surface water features following the storm. Areas with a negative NDVI slope were not assigned a recovery time.We used a combination of the NDVI slope, estimated time to full NDVI recovery, and the average change in NDVI between the pre- and post-storm periods to categorize mangrove forest resilience, the potential for mangroves to rebound to pre-disturbance conditions. The specific criteria for mangrove recovery rates and mangrove damage thresholds were adapted from field and remote sensing studies, respectively6,19,25. Regions of high resilience (a combination of high resistance and resilience) were identified based on rapid recovery and/or little to no immediate impact from the storm: (1) areas that were observed to recover to pre-disturbance conditions during 2018, (2) areas that were predicted to recover within 5 years regardless of the post-storm drop in NDVI6, and (3) regions with a post-storm change in NDVI 15 years or a negative NDVI slope that occurred in regions with the largest ( >0.2) post-storm drop in average NDVI25 (Supplementary Fig. 9). The resilience class map is available online for download53.Mangrove species and elevationWe used species level maps developed by the National Park Service for Everglades National Park50 to characterize the impact of Hurricane Irma on different mangrove species. For that study, dominant species were identified through photo-interpretation of stereoscopic, color-infrared aerial imagery. Grid cells of 50 m × 50 m covering an area (Region 3) of ~100,000 ha in southwest Florida were interpreted based on the majority cover type and validated using field observations. A total of 169 vegetation cover classes were identified in this region, however, only five mangrove cover classes were considered for these analyses: Avicennia germinans (Black Mangrove), Laguncularia racemosa (White Mangrove), Rhizophora mangle (Red Mangrove), Conocarpus erectus (Buttonwood), and a single mixed species mangrove class. Mangrove forest communities were defined as the dominant diagnostic species in the upper-most stratum50. The mangrove species data were reprojected to match the Landsat resolution and the resilience maps. We used the intersection of the resilience and species extent maps to estimate the proportion of each resilience class by dominant species.The USGS National Elevation Dataset (NED) was used to estimate the soil elevation across southwest Florida. The 1/9 arc second (~3 m × 3 m) products were acquired from NED, and reprojected to Landsat resolution to estimate the proportion of each resilience class by soil elevation.Additional data and analysisModeled maximum storm surge data for Hurricane Irma were acquired from Coastal Emergency Risks Assessment data portal. Storm surge is derived from the ADCIRC Prediction System that solves for time dependent, circulation, and transport in multiple dimensions54. Maximum sustained hurricane wind speed was modeled hourly at a 5 km × 5 km resolution for 2017 by NASA’s Global Modeling and Assimilation Office (GMAO)55. The storm maximum wind speed for each 5 km × 5 km grid cell was calculated and binned into six discrete classes of wind speeds at 5 m s−1 increments: 26–30, 31–35, 36–40, 41–45, 46–50, and >50.Statistical analysesCanopy height losses measured from NASA G-LiHT data were grouped by five pre-storm canopy height classes (0–5 m, 5–10 m, 10–15 m, 15–20 m, and >20 m). All valid pixels within the lidar footprint was used to calculate the mean, standard error, and area (sum of 1 m × 1 m pixels) for each class (Supplementary Table 1). These results were then tested for significant differences between canopy height losses and pre-storm canopy height classes between using a one-way ANOVA analysis with a post-hoc Tukey test in R (version 4.0.3). For testing the significance between environmental variables (i.e., pre-storm canopy height, canopy height loss, percent canopy height loss, surface elevation, and storm surge water level above ground) we employed a two-sided Kolmogorov–Smirnov test56 implemented in R (version 4.0.3). First, we created a multi-band stacked image which included each of the variable layers. Within each resilience class (i.e., Low, Intermediate, and High) with randomly selected 10,000–20,000 points using Google Earth Engine to sample from the environmental variables images. From that sample set we then randomly selected 500 samples within each of the resilience classes. Each class combination (1) Low-Intermediate, (2) Low-High, and (3) Intermediate-High were compared using the Kolmogorov–Smirnov test. We repeated this procedure using 5000 iterations in order to provide a robust estimate of the Kolmogorow–Smirnov statistic, including the mean and first and third quartiles, which were then compared to the critical value (Supplementary Table 2).Reporting summaryFurther information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    The impact of large and small dams on malaria transmission in four basins in Africa

    Study areaFour major river basins, located across different sub-regions of SSA, were selected for this study: Limpopo, Omo-Turkana, Volta, and Zambezi (Fig. 1). These basins were selected to (i) foster inclusion of enable different African regions and (ii) ensure focus on basins with sufficient data availability.Figure 1source malaria data23 on ArcGIS software (version 10.5. 1, Environmental Systems Research Institute Inc, Redlands, CA, USA, 2016)].Distribution of large and small dams in Limpopo, Volta, Zambezi and Omo-Turkana basins by malaria stability zone. [The figure was made using open-Full size imageThe Limpopo River basin is located in southern Africa. Draining an area of approximately 408,000 km2, the Limpopo River basin is distributed among South Africa (45%), Botswana (20%), Zimbabwe (15%) and Mozambique (20%). About 14 million people live in this basin. The climate of the Limpopo River basin varies along the path of the river from a temperate climate in the west to a subtropical climate at the river mouth in Mozambique. The hydrology of the Limpopo River basin is influenced by the highly seasonal distribution of rainfall over the catchment. About 95% of rain falls between October and April with a peak normally in February. Temperature varies from 30 to 34 °C in summer and 22–26 °C in winter15.The Volta River basin is located in West Africa with a population of over 23 million. Draining an area of 409,000 km2 the basin is spread across six countries: Benin (4%), Burkina Faso (42%), Cote d’Ivoire (3%), Ghana (41%), Mali (4%) and Togo (6%). Average annual rainfall varies across the basin from approximately 1600 mm in the southeast, to about 360 mm in the north. Annual mean temperatures in the basin vary from 27 to 30 °C16. The main rainy season is between March and October.The Zambezi River basin is located in southern Africa. Draining an area of 1.34 million km2, the basin is spread across eight countries: Angola (19%), Botswana (1%), Namibia (1%) Benin (4%), Zimbabwe (16%), Zambia (42%), Tanzania (2%), Malawi (8%) and Mozambique (12%). The population of the Zambezi basin is estimated to be about 32 million. Annual rainfall in the basin ranges from 550 mm in the south to 1800 mm in the north. The annual mean temperatures ranges from 18 °C at higher elevations in the south of the basin to 26 °C for low elevations in the delta in Mozambique17.The Omo-Turkana Basin covers approximately 131,000 km2, stretching from southern Ethiopia to northern Kenya. Hydrologically, the basin is dominated by Lake Turkana, with the Omo River, which drains the Ethiopian portion of the basin, supplying 90% of the inflow to the lake. The basin is home to approximately 15 million people, the majority of whom live in the Ethiopian highlands, in the north. The annual mean temperature ranges from 24 °C in the north to 29 °C in the south. The mean annual rainfall ranges from 250 mm in the south to 500 mm in the north18.Data sourcesDam dataSmall damsData on location and size of small dams are not readily available in either global or regional data sets. The European Commission’s Joint Research Center (JRC) Yearly Water Classification History v1.0 data set was used to identify water bodies in each of the four basins19. Water bodies less than 100 ha and greater than 2 ha were identified. All were checked with Google Earth images to distinguish between reservoirs and natural water bodies (Supplementary Fig. S1). Ultimately, a total of 4907 small dams located in the four basins were identified and included in the analyses.Large damsFor large dams, the FAO African Dams Database20, International Commission for Large dams (ICOLD)21 and the International Rivers Database22, which together contain 1286 georeferenced African large dams, were utilized. The accuracy of dam locations was first verified with Google Earth. When the location of a dam did not precisely match the coordinates stipulated in either of the two databases, manual corrections were made by adjusting the coordinates of a dam to its location as shown in Google Earth (see Supplementary Information). Dams for which precise locations could not be determined, as well as dams without reservoirs (i.e., run-of-river schemes), were removed. Ultimately, across the four basins, a total of 258 large dams with confirmed georeferenced locations were identified and included in the analyses.Perimeters of large and small dam reservoirsReservoir perimeters of both large and small dams were extracted from the European Commission’s Joint Research Center (JRC) global surface water datasets19, published through the Google Earth Engine. This dataset includes maps of the location and temporal variability in maximum perimeter records of the global surface water coverage from 1984 to 2015. In this study, the maximum perimeter records were used in each year of 2000, 2005, 2010 and 2015. The data were exported to ArcGIS.Data on anopheles mosquito distributionData for vector distribution were obtained from the Malaria Atlas Project (MAP) database23. The MAP database contains a georeferenced illustration of the major malaria vector species in different malaria-endemic areas in Africa.Malaria dataAnnual malaria incidence data were obtained from the MAP database. We acquired data for the years 2000, 2005, 2010 and 2015. These years were selected to align with updates to Worldpop population data24, which are recomputed every five years. MAP produced a 1 km resolution continuous map of annual malaria incidence for Africa based on 33,761 studies across the region. We imported these data to ArcGIS for analyses. Annual malaria incidence was determined as the number of cases per 1000 population. To ascertain the impact of dams on malaria incidence rates as a function of distance from the reservoir perimeter, we created two distance zones: 0–5 km (at risk) and 5–10 km (control). When distance zones were overlapping for two or more nearby dams, areas were assigned to the closest distance cohort. Populations residing more than 5 km from a reservoir perimeter (large or small) were considered to be free of risk from dam induced malaria transmission because the maximum mosquitoes’ flight range is considered to be  0.1 malaria cases per 1000 population), unstable (≤ 0.1 malaria cases per 1000 population) and no malaria (zero malaria incidence) based on the level of malaria incidence in each of the four years: 2000, 2005, 2010, and 2015. The number of dams in each of the three stability categories for each of the four years was determined, as well as the population at-risk of dam-related malaria (i.e.,  More

  • in

    Seasonal influence on the bathymetric distribution of an endangered fish within a marine protected area

    1.Lejeusne, C., Chevaldonné, P., Pergent-Martini, C., Boudouresque, C. F. & Pérez, T. Climate change effects on a miniature ocean: The highly diverse, highly impacted Mediterranean Sea. Trends Ecol. Evol. 25, 250–260 (2010).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    2.Fulton, E. A. et al. Modelling marine protected areas: Insights and hurdles. Philos. Trans. R. Soc. B Biol. Sci. 370, 20140278 (2015).Article 

    Google Scholar 
    3.Roff, J. & Zacharias, M. Marine Conservation Ecology (Earthscan, 2011).
    Google Scholar 
    4.Mitcheson, Y. S. D. et al. A global baseline for spawning aggregations of reef fishes. Conserv. Biol. 22, 1233–1244 (2008).Article 

    Google Scholar 
    5.Salinas-de-León, P., Rastoin, E. & Acuña-Marrero, D. First record of a spawning aggregation for the tropical eastern Pacific endemic grouper Mycteroperca olfax in the Galapagos Marine Reserve. J. Fish Biol. 87, 179–186 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    6.Bueno, L. S. et al. Evidence for spawning aggregations of the endangered Atlantic goliath grouper Epinephelus itajara in southern Brazil. J. Fish Biol. 89, 876–889 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    7.Kumar, V. et al. Biological clocks and regulation of seasonal reproduction and migration in birds. Physiol. Biochem. Zool. 83, 827–835 (2010).ADS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    8.van Haren, H. & Compton, T. J. Diel vertical migration in deep sea plankton is finely tuned to latitudinal and seasonal day length. PLoS ONE 8, e64435 (2013).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    9.Horký, P. & Slavík, O. Diel and seasonal rhythms of asp Leuciscus aspius (L.) in a riverine environment. Ethol. Ecol. Evol. 29, 449–459 (2017).Article 

    Google Scholar 
    10.Falcón, J., Besseau, L., Sauzet, S. & Boeuf, G. Melatonin effects on the hypothalamo–pituitary axis in fish. Trends Endocrinol. Metab. 18, 81–88 (2007).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    11.Oliveira, C. et al. Monthly day/night changes and seasonal daily rhythms of sexual steroids in Senegal sole (Solea senegalensis) under natural fluctuating or controlled environmental conditions. Comp. Biochem. Physiol. A: Mol. Integr. Physiol. 152, 168–175 (2009).ADS 
    Article 
    CAS 

    Google Scholar 
    12.Wuitchik, D. M. et al. Seasonal temperature, the lunar cycle and diurnal rhythms interact in a combinatorial manner to modulate genomic responses to the environment in a reef-building coral. Mol. Ecol. 28, 3629–3641 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    13.Sanchez-Cardenas, C. et al. Pituitary growth hormone network responses are sexually dimorphic and regulated by gonadal steroids in adulthood. PNAS 107, 21878–21883 (2010).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    14.Stock, C. A. et al. Seasonal sea surface temperature anomaly prediction for coastal ecosystems. Prog. Oceanogr. 137, 219–236 (2015).ADS 
    Article 

    Google Scholar 
    15.Bisagni, J. J. Salinity variability along the eastern continental shelf of Canada and the United States, 1973–2013. Cont. Shelf Res. 126, 89–109 (2016).ADS 
    Article 

    Google Scholar 
    16.Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M. & West, G. B. Toward a metabolic theory of ecology. Ecology 85, 1771–1789 (2004).Article 

    Google Scholar 
    17.Dorts, J. et al. Evidence that elevated water temperature affects the reproductive physiology of the European bullhead Cottus gobio. Fish Physiol. Biochem. 38, 389–399 (2012).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    18.Arfuso, F. et al. Water temperature influences growth and gonad differentiation in European sea bass (Dicentrarchus labrax, L. 1758). Theriogenology 88, 145–151 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    19.Aspillaga, E. et al. Thermal stratification drives movement of a coastal apex predator. Sci. Rep. 7, 526 (2017).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    20.Martin, T. L. & Huey, R. B. Why, “suboptimal” is optimal: Jensen’s inequality and ectotherm thermal preferences. Am. Nat. 171, E102–E118 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Freitas, C., Olsen, E. M., Moland, E., Ciannelli, L. & Knutsen, H. Behavioral responses of Atlantic cod to sea temperature changes. Ecol. Evol. 5, 2070–2083 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    22.Harmelin, J.-G. & Marinopoulos, J. Recensementde la Population de corbs (Sciaena umbra, Linneaus1758: Pisces) du Parc National de Port-Cros (Méditerrannée, France) par Inventaires Visuels 265–275 (1993).23.Coll, J., Linde, M., García-Rubies, A., Riera, F. & Grau, A. M. Spear fishing in the Balearic Islands (west central Mediterranean): Species affected and catch evolution during the period 1975–2001. Fish. Res. 70, 97–111 (2004).Article 

    Google Scholar 
    24.Lloret, J. et al. Spearfishing pressure on fish communities in rocky coastal habitats in a Mediterranean marine protected area. Fish. Res. 94, 84–91 (2008).Article 

    Google Scholar 
    25.Harmelin, J.-G. Statut du corb (Sciaena umbra) en méditerranée. In Les Espèces Marines à Protéger en Méditerranée (eds Boudouresque, C. F. et al.) 219–227 (GiS Posidonie Publ., 1991).
    Google Scholar 
    26.Mayol, J., Grau, A. M., Riera, F. & Oliver, J. Llista Ver-mella dels Peixos de les Balears (2000).
    Google Scholar 
    27.Chao, L. The IUCN Red List of Threatened Species 2020: e.T198707A130230194Sciaena umbra (2020). https://doi.org/10.2305/IUCN.UK.2020-2.RLTS.T198707A130230194.en.28.Forcada, A. et al. Effects of habitat on spillover from marine protected areas to artisanal fisheries. Mar. Ecol. Prog. Ser. 379, 197–211 (2009).ADS 
    Article 

    Google Scholar 
    29.Franco, A. D., Bussotti, S., Navone, A., Panzalis, P. & Guidetti, P. Evaluating effects of total and partial restrictions to fishing on Mediterranean rocky-reef fish assemblages. Mar. Ecol. Prog. Ser. 387, 275–285 (2009).ADS 
    Article 

    Google Scholar 
    30.Le Préfet de la region Provence-Alpes-Côte d’Azur, Préfet de la zone de défense et de sécurité Sud, Préfet des Bouches-du-Rhône. http://www.dirm.mediterranee.developpement-durable.gouv.fr/IMG/pdf/ap_corb_med_continentale_20_dec_2018-2.pdf (Accessed 20 December 2018).31.Harmelin-Vivien, M. et al. Effects of reserve protection level on the vulnerable fish species Sciaena umbra and implications for fishing management and policy. Glob. Ecol. Conserv. 3, 279–287 (2015).Article 

    Google Scholar 
    32.Chakroun-Marzouk, N. & Ktari, M.-H. Le Corb des côtes Tunisiennes, Sciaena umbra (Sciaenidae): Cycle Sexuel, Age et Croissance 15 (2003).33.Derbal, F. & Kara, M. H. Régime Alimentaire du corb Sciaena umbra (Sciaenidae) des côtes de l’est Algérien 9 (2007).34.Engin, S. & Seyhan, K. Age, growth, sexual maturity and food composition of Sciaena umbra in the south-eastern Black Sea, Turkey. J. Appl. Ichthyol. 25, 96–99 (2009).Article 

    Google Scholar 
    35.Botsford, L. W. et al. Connectivity, sustainability, and yield: Bridging the gap between conventional fisheries management and marine protected areas. Rev. Fish Biol. Fish. 19, 69–95 (2009).Article 

    Google Scholar 
    36.Alós, J. & Cabanellas-Reboredo, M. Experimental acoustic telemetry experiment reveals strong site fidelity during the sexual resting period of wild brown meagre, Sciaena umbra. J. Appl. Ichthyol. 28, 606–611 (2012).Article 

    Google Scholar 
    37.Jadot, C., Donnay, A., Acolas, M. L., Cornet, Y. & Bégout Anras, M. L. Activity patterns, home-range size, and habitat utilization of Sarpa salpa (Teleostei: Sparidae) in the Mediterranean Sea. ICES J. Mar. Sci. 63, 128–139 (2006).Article 

    Google Scholar 
    38.Jorgensen, S. J. et al. Limited movement in blue rockfish Sebastes mystinus: Internal structure of home range. Mar. Ecol. Prog. Ser. 327, 157–170 (2006).ADS 
    Article 

    Google Scholar 
    39.Kerwath, S. E., Götz, A., Attwood, C. G., Sauer, W. H. H. & Wilke, C. G. Area utilisation and activity patterns of roman Chrysoblephus laticeps (Sparidae) in a small marine protected area. Afr. J. Mar. Sci. 29, 259–270 (2007).Article 

    Google Scholar 
    40.Collins, A. B., Heupel, M. R. & Motta, P. J. Residence and movement patterns of cownose rays Rhinoptera bonasus within a south-west Florida estuary. J. Fish Biol. 71, 1159–1178 (2007).Article 

    Google Scholar 
    41.Abecasis, D. & Erzini, K. Site fidelity and movements of gilthead sea bream (Sparus aurata) in a coastal lagoon (Ria Formosa, Portugal). Estuar. Coast. Shelf Sci. 79, 758–763 (2008).ADS 
    Article 

    Google Scholar 
    42.Afonso, P. et al. A multi-scale study of red porgy movements and habitat use, and its application to the design of marine reserve networks. In Tagging and Tracking of Marine Animals with Electronic Devices (eds Nielsen, J. L. et al.) 423–443 (Springer, 2009).Chapter 

    Google Scholar 
    43.March, D., Palmer, M., Alós, J., Grau, A. & Cardona, F. Short-term residence, home range size and diel patterns of the painted comber Serranus scriba in a temperate marine reserve. Mar. Ecol. Prog. Ser. 400, 195–206 (2010).ADS 
    Article 

    Google Scholar 
    44.Zeller, D. C. Ultrasonic telemetry: Its application to coral reef fisheries research. Fish. Bull. 97, 1058–1065 (1999).
    Google Scholar 
    45.Heupel, M. R., Semmens, J. M. & Hobday, A. J. Automated acoustic tracking of aquatic animals: Scales, design and deployment of listening station arrays. Mar. Freshw. Res. 57, 1–13 (2006).Article 

    Google Scholar 
    46.Lowe, C. G., Topping, D. T., Cartamil, D. P. & Papastamatiou, Y. P. Movement patterns, home range, and habitat utilization of adult kelp bass Paralabrax clathratus in a temperate no-take marine reserve. Mar. Ecol. Prog. Ser. 256, 205–216 (2003).ADS 
    Article 

    Google Scholar 
    47.Kaunda-Arara, B. & Rose, G. A. Homing and site fidelity in the greasy grouper Epinephelus tauvina (Serranidae) within a marine protected area in coastal Kenya. Mar. Ecol. Prog. Ser. 277, 245–251 (2004).ADS 
    Article 

    Google Scholar 
    48.Parsons, D. & Egli, D. Fish movement in a temperate marine reserve: New insights through application of acoustic tracking. Mar. Technol. Soc. J. 39, 56–63 (2005).Article 

    Google Scholar 
    49.Topping, D. T., Lowe, C. G. & Caselle, J. E. Home range and habitat utilization of adult California sheephead, Semicossyphus pulcher (Labridae), in a temperate no-take marine reserve. Mar. Biol. 147, 301–311 (2005).Article 

    Google Scholar 
    50.Pastor, J. et al. Acoustic telemetry survey of the dusky grouper (Epinephelus marginatus) in the Marine Reserve of Cerbère-Banyuls: Informations on the territoriality of this emblematic species. C.R. Biol. 332, 732–740 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    51.D’Anna, G., Giacalone, V. M., Pipitone, C. & Badalamenti, F. Movement pattern of white seabream, Diplodus sargus (L., 1758) (Osteichthyes, Sparidae) acoustically tracked in an artificial reef area. Ital. J. Zool. 78, 255–263 (2011).Article 

    Google Scholar 
    52.La Mesa, G., Consalvo, I., Annunziatellis, A. & Canese, S. Movement patterns of the parrotfish Sparisoma cretense in a Mediterranean marine protected area. Mar. Environ. Res. 82, 59–68 (2012).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    53.Picciulin, M. et al. Passive acoustic monitoring of Sciaena umbra on rocky habitats in the Venetian littoral zone. Fish. Res. 145, 76–81 (2013).Article 

    Google Scholar 
    54.Lenfant, P., Louisy, P. & Licari, M.-L. Recensement des Mérous bruns (Epinephelus marginatus) de la Réserve Naturelle de Cerbère-Banyuls (France, Méditerranée) Effectué en Septembre 2001, aprés 17 Années de Protection 10 (2003).55.Koeck, B., Gudefin, A., Romans, P., Loubet, J. & Lenfant, P. Effects of intracoelomic tagging procedure on white seabream (Diplodus sargus) behavior and survival. J. Exp. Mar. Biol. Ecol. 440, 1–7 (2013).Article 

    Google Scholar 
    56.Garcia, J., Mourier, J. & Lenfant, P. Spatial behavior of two coral reef fishes within a Caribbean marine protected area. Mar. Environ. Res. 109, 41–51 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    57.Percie du Sert, N. et al. The ARRIVE guidelines 2.0: Updated guidelines for reporting animal research. PLoS Biol. 18(7), e3000410. https://doi.org/10.1371/journal.pbio.3000410 (2020).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    58.Grau, A., Linde, M. & Grau, A. M. Reproductive biology of the vulnerable species Sciaena umbra Linnaeus, 1758 (Pisces: Sciaenidae). Sci. Mar. 73, 67–81 (2009).Article 

    Google Scholar 
    59.McKinzie, M. K., Jarvis, E. T. & Lowe, C. G. Fine-scale horizontal and vertical movement of barred sand bass, Paralabrax nebulifer, during spawning and non-spawning seasons. Fish. Res. 150, 66–75 (2014).Article 

    Google Scholar 
    60.Kassambara, A. & Mundt, F. factoextra: Extract and Visualize the Results of Multivariate Data Analyses (2017).61.Zuur, A., Ieno, E. N., Walker, N., Saveliev, A. A. & Smith, G. M. Mixed Effects Models and Extensions in Ecology with R (Springer, 2009).MATH 
    Book 

    Google Scholar 
    62.Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D. & R Core Team. nlme: Linear and Nonlinear Mixed Effects Models (2018).63.Wood, S. N. Generalized Additive Models: An Introduction with R 2nd edn. (Chapman and Hall/CRC, 2017).MATH 
    Book 

    Google Scholar 
    64.R Core Team. R: A Language and Environment for Statistical Computing. (R Foundation for Statistical Computing, 2018).65.Wearmouth, V. J. & Sims, D. W. Chapter 2 sexual segregation in marine fish, reptiles, birds and mammals: Behaviour patterns, mechanisms and conservation implications. Adv. Mar. Biol. 54, 107–170 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    66.Haraldstad, Ø. & Jonsson, B. Age and sex segregation in habitat utilization by brown trout in a Norwegian Lake. Trans. Am. Fish. Soc. 112, 27–37 (1983).Article 

    Google Scholar 
    67.L’Abée-Lund, J. H., Langeland, A., Jonsson, B. & Ugedal, O. Spatial segregation by age and size in Arctic Charr: A trade-off between feeding possibility and risk of predation. J. Anim. Ecol. 62, 160–168 (1993).Article 

    Google Scholar 
    68.Oxenford, H. A. & Hunte, W. Feeding habits of the dolphinfish (Coryphaena hippurus) in the eastern Caribbean. Sci. Mar. 63, 303–315 (1999).Article 

    Google Scholar 
    69.Sarà, G. et al. Effect of boat noise on the behaviour of bluefin tuna Thunnus thynnus in the Mediterranean Sea. Mar. Ecol. Prog. Ser. 331, 243–253 (2007).ADS 
    Article 

    Google Scholar 
    70.Codarin, A., Wysocki, L. E., Ladich, F. & Picciulin, M. Effects of ambient and boat noise on hearing and communication in three fish species living in a marine protected area (Miramare, Italy). Mar. Pollut. Bull. 58, 1880–1887 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    71.Picciulin, M., Sebastianutto, L., Codarin, A., Calcagno, G. & Ferrero, E. A. Brown meagre vocalization rate increases during repetitive boat noise exposures: A possible case of vocal compensation. J. Acoust. Soc. Am. 132, 3118–3124 (2012).ADS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    72.McCormick, M. I., Allan, B. J. M., Harding, H. & Simpson, S. D. Boat noise impacts risk assessment in a coral reef fish but effects depend on engine type. Sci. Rep. 8, 3847 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    73.Robichaud, D. & Rose, G. A. Sex differences in cod residency on a spawning ground. Fish. Res. 60, 33–43 (2003).Article 

    Google Scholar 
    74.Fiorentino, F. et al. On a spawning aggregation of the brown meagre Sciaena umbra L. 1758 (Sciaenidae, Osteichthyes) in the Maltese waters (Sicilian Channel—Central Mediterranean). Rapp. Commun. Int. Mer Médit. 36, 266 (2001).
    Google Scholar 
    75.Furukawa, S. et al. Vertical movements of Pacific bluefin tuna (Thunnus orientalis) and dolphinfish (Coryphaena hippurus) relative to the thermocline in the northern East China Sea. Fish. Res. 149, 86–91 (2014).Article 

    Google Scholar 
    76.Claireaux, G., Webber, D., Kerr, S. & Boutilier, R. Physiology and behaviour of free-swimming Atlantic cod (Gadus morhua) facing fluctuating temperature conditions. J. Exp. Biol. 198, 49–60 (1995).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    77.Armstrong, J. B. et al. Diel horizontal migration in streams: Juvenile fish exploit spatial heterogeneity in thermal and trophic resources. Ecology 94, 2066–2075 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    78.Bobe, J. & Labbé, C. Egg and sperm quality in fish. Gen. Comp. Endocrinol. 165, 535–548 (2010).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    79.Pepin, P. Effect of temperature and size on development, mortality, and survival rates of the pelagic early life history stages of marine fish. Can. J. Fish. Aquat. Sci. 48, 503–518 (1991).Article 

    Google Scholar 
    80.Guevara-Fletcher, C., Alvarez, P., Sanchez, J. & Iglesias, J. Effect of temperature on the development and mortality of European hake (Merluccius merluccius L.) eggs from southern stock under laboratory conditions. J. Exp. Mar. Biol. Ecol. 476, 50–57 (2016).Article 

    Google Scholar 
    81.Dubrovský, M. et al. Multi-GCM projections of future drought and climate variability indicators for the Mediterranean region. Reg. Environ Change 14, 1907–1919 (2014).Article 

    Google Scholar 
    82.Pankhurst, N. W. & Munday, P. L. Effects of climate change on fish reproduction and early life history stages. Mar. Freshw. Res. 62, 1015–1026 (2011).CAS 
    Article 

    Google Scholar 
    83.McKenzie, D. J. et al. Conservation physiology of marine fishes: State of the art and prospects for policy. Conserv. Physiol. 4, 046 (2016).Article 

    Google Scholar 
    84.Fabi, G., Panfili, M. & Spagnolo, A. Note on feeding of Sciaena umbra L. (Asteichthyes:Sciaenidae) in the central Adriatic Sea. Rapp. Comm. Int. Mer Médit. 35, 426 (1998).
    Google Scholar 
    85.Fabi, G., Manoukian, S. & Spagnolo, A. Feeding behavior of three common fishes at an artificial reef in the northern Adriatic Sea. Bull. Mar. Sci. 78, 39–56 (2006).
    Google Scholar 
    86.Ramcharitar, J., Gannon, D. P. & Popper, A. N. Bioacoustics of fishes of the family Sciaenidae (croakers and drums). Trans. Am. Fish. Soc. 135, 1409–1431 (2006).Article 

    Google Scholar 
    87.Mesa, M. L., Colella, S., Giannetti, G. & Arneri, E. Age and growth of brown meagre Sciaena umbra (Sciaenidae) in the Adriatic Sea. Aquat. Living Resour. 21, 153–161 (2008).Article 

    Google Scholar 
    88.Picciulin, M. et al. Diagnostics of nocturnal calls of Sciaena umbra (L., fam. Sciaenidae) in a nearshore Mediterranean marine reserve. Bioacoustics 22, 109–120 (2013).Article 

    Google Scholar 
    89.Schmidt, M. B. & Gassner, H. Influence of scuba divers on the avoidance reaction of a dense vendace (Coregonus albula L.) population monitored by hydroacoustics. Fish. Res. 82, 131–139 (2006).Article 

    Google Scholar 
    90.Moffitt, E. A., Botsford, L. W., Kaplan, D. M. & O’Farrell, M. R. Marine reserve networks for species that move within a home range. Ecol. Appl. 19, 1835–1847 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar  More

  • in

    Plant-soil feedbacks help explain biodiversity-productivity relationships

    1.Tilman, D. et al. The influence of functional diversity and composition on ecosystem processes. Science 277, 1300–1302 (1997).CAS 
    Article 

    Google Scholar 
    2.Cardinale, B. J. et al. Impacts of plant diversity on biomass production increase through time because of species complementarity. Proc. Natl Acad. Sci. USA 104, 18123–18128 (2007).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    3.Van Ruijven, J. & Berendse, F. Diversity-productivity relationships: Initial effects, long-term patterns, and underlying mechanisms. Proc. Natl Acad. Sci. USA 102, 695–700 (2005).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    4.Jochum, M. et al. The results of biodiversity–ecosystem functioning experiments are realistic. Nat. Ecol. Evol. 4, 1485–1494 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    5.Jing, J., Bezemer, T. M. & van der Putten, W. H. Complementarity and selection effects in early and mid-successional plant communities are differentially affected by plant-soil feedback. J. Ecol. 103, 641–647 (2015).Article 

    Google Scholar 
    6.Tilman, D., Hill, J. & Lehman, C. Carbon-negative biofuels from low-input high-diversity grassland biomass. Science 314, 1598–1600 (2006).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    7.Mueller, K. E., Tilman, D., Fornara, D. A. & Hobbie, S. E. Root depth distribution and the diversity–productivity relationship in a long-term grassland experiment. Ecology 94, 787–793 (2013).Article 

    Google Scholar 
    8.Hector, A., Bazeley-White, E., Loreau, M., Otway, S. & Schmid, B. Overyielding in grassland communities: testing the sampling effect hypothesis with replicated biodiversity experiments. Ecol. Lett. 5, 502–511 (2002).Article 

    Google Scholar 
    9.Barry, K. E. et al. The future of complementarity: disentangling causes from consequences. Trends Ecol. Evol. 34, 167–180 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    10.Kulmatiski, A., Beard, K. H. & Heavilin, J. Plant-soil feedbacks provide an additional explanation for diversity-productivity relationships. Proc. R. Soc. B Biol. Sci. 279, 3020–3026 (2012).Article 

    Google Scholar 
    11.Loreau, M. & Hector, A. Partitioning selection and complementarity in biodiversity experiments. Nature 412, 72–76 (2001).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    12.Tedersoo, L., Bahram, M. & Zobel, M. How mycorrhizal associations drive plant population and community biology. Science 367, 6480 (2020).Article 
    CAS 

    Google Scholar 
    13.Maron, J. L., Marler, M., Klironomos, J. N. & Cleveland, C. C. Soil fungal pathogens and the relationship between plant diversity and productivity. Ecol. Lett. 14, 36–41 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    14.Wang, G. et al. Soil microbiome mediates positive plant diversity‐productivity relationships in late successional grassland species. Ecol. Lett. 22, 13273 (2019).Article 

    Google Scholar 
    15.Wright, A. J., Wardle, D. A., Callaway, R. & Gaxiola, A. The overlooked role of facilitation in biodiversity experiments. Trends Ecol. Evol. 32, 383–390 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    16.Bever, J. D., Platt, T. G. & Morton, E. R. Microbial population and community dynamics on plant roots and their feedbacks on plant communities. Annu. Rev. Microbiol. 66, 265–283 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    17.Bauer, J. T., Koziol, L. & Bever, J. D. Local adaptation of mycorrhizae communities changes plant community composition and increases aboveground productivity. Oecologia 192, 735–744 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    18.Bever, J. D. Feeback between plants and their soil communities in an old field community. Ecology 75, 1965–1977 (1994).Article 

    Google Scholar 
    19.Hendriks, M. et al. Independent variations of plant and soil mixtures reveal soil feedback effects on plant community overyielding. J. Ecol. 101, 287–297 (2013).Article 

    Google Scholar 
    20.Zuppinger-Dingley, D. L., Flynn, D. F. B., De Deyn, G. B., Petermann, J. S. & Schmid, B. Plant selection and soil legacy enhance long-term biodiversity effects. Ecology 97, 15–0599.1 (2015).
    Google Scholar 
    21.Mommer, L. et al. Lost in diversity: the interactions between soil-borne fungi, biodiversity and plant productivity. N. Phytol. 218, 542–553 (2018).Article 

    Google Scholar 
    22.Guerrero‐Ramírez, N. R., Reich, P. B., Wagg, C., Ciobanu, M. & Eisenhauer, N. Diversity‐dependent plant–soil feedbacks underlie long‐term plant diversity effects on primary productivity. Ecosphere 10, e02704 (2019).Article 

    Google Scholar 
    23.van Ruijven, J., Ampt, E., Francioli, D. & Mommer, L. Do soil-borne fungal pathogens mediate plant diversity–productivity relationships? Evidence and future opportunities. J. Ecol. 108, 1810–1821 (2020).Article 

    Google Scholar 
    24.Schnitzer, S. A. et al. Soil microbes drive the classic plant diversity–productivity pattern. Ecology 92, 296–303 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    25.Lekberg, Y. et al. Relative importance of competition and plant-soil feedback, their synergy, context dependency and implications for coexistence. Ecol. Lett. 21, 1268–1281 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    26.Cowles, J. Mechanisms of Coexistence: Implications for Biodiversity-Ecosystem Functioning Relationships in a Changing World. Dissertation, The University of Minnesota (2015).27.Forero, L. E., Grenzer, J., Heinze, J., Schittko, C. & Kulmatiski, A. Greenhouse- and field-measured plant-soil feedbacks are not correlated. Front. Environ. Sci. 7, 184 (2019).Article 

    Google Scholar 
    28.Kulmatiski, A. & Kardol, P. in Getting Plant—Soil Feedbacks out of the Greenhouse: Experimental and Conceptual Approaches 449–472 (Springer, 2008).29.Pernilla Brinkman, E., Van der Putten, W. H., Bakker, E. J. & Verhoeven, K. J. F. Plant-soil feedback: experimental approaches, statistical analyses and ecological interpretations. J. Ecol. 98, 1063–1073 (2010).Article 

    Google Scholar 
    30.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 
    31.Rinella, M. J. & Reinhart, K. O. Toward more robust plant-soil feedback research. Ecology 99, 550–556 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    32.Crawford, K. M. et al. When and where plant‐soil feedback may promote plant coexistence: a meta‐analysis. Ecol. Lett. 22, 13278 (2019).Article 

    Google Scholar 
    33.Clark, A. T. et al. How to estimate complementarity and selection effects from an incomplete sample of species. Methods Ecol. Evol. 10, 2141–2152 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    34.Anacker, B. L., Klironomos, J. N., Maherali, H., Reinhart, K. O. & Strauss, S. Y. Phylogenetic conservatism in plant-soil feedback and its implications for plant abundance. Ecol. Lett. 17, 1613–1621 (2014).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    35.Mehrabi, Z. & Tuck, S. L. Relatedness is a poor predictor of negative plant–soil feedbacks. N. Phytol. 205, 1071–1075 (2015).Article 

    Google Scholar 
    36.Kulmatiski, A., Beard, K. H., Stevens, J. R. & Cobbold, S. M. Plant-soil feedbacks: a meta-analytical review. Ecol. Lett. 11, 980–992 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    37.Beals, K. K. et al. Predicting plant-soil feedback in the field: meta-analysis reveals that competition and environmental stress differentially influence psf. Front. Ecol. Evol. 8, 191 (2020).Article 

    Google Scholar 
    38.Kos, M., Tuijl, M. A. B., de Roo, J., Mulder, P. P. J. & Bezemer, T. M. Species-specific plant-soil feedback effects on above-ground plant-insect interactions. J. Ecol. 103, 904–914 (2015).CAS 
    Article 

    Google Scholar 
    39.Bukowski, A. R. & Petermann, J. S. Intraspecific plant-soil feedback and intraspecific overyielding in Arabidopsis thaliana. Ecol. Evol. 4, 2533–2545 (2014).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    40.Tilman, D., Wedin, D. & Knops, J. Productivity and sustainability influenced by biodiversity in grassland ecosystems. Nature 379, 718–720 (1996).CAS 
    Article 

    Google Scholar 
    41.Fornara, D. A. & Tilman, D. Ecological mechanisms associated with the positive diversity–productivity relationship in an N-limited grassland. Ecology 90, 408–418 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.Laughlin, D. C. et al. The hierarchy of predictability in ecological restoration: are vegetation structure and functional diversity more predictable than community composition? J. Appl. Ecol. 54, 1058–1069 (2017).Article 

    Google Scholar 
    43.Metcalfe, H., Milne, A. E., Deledalle, F. & Storkey, J. Using functional traits to model annual plant community dynamics. Ecology 101, e03167 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    44.Moulin, T., Perasso, A., Calanca, P. & Gillet, F. DynaGraM: a process-based model to simulate multi-species plant community dynamics in managed grasslands. Ecol. Modell. 439, 109345 (2021).Article 

    Google Scholar 
    45.Putten, W. H., Bradford, M. A., Pernilla Brinkman, E., Voorde, T. F. J. & Veen, G. F. Where, when and how plant–soil feedback matters in a changing world. Funct. Ecol. 30, 1109–1121 (2016).Article 

    Google Scholar 
    46.Eisenhauer, N., Reich, P. B. & Scheu, S. Increasing plant diversity effects on productivity with time due to delayed soil biota effects on plants. Basic Appl. Ecol. 13, 571–578 (2012).Article 

    Google Scholar 
    47.Hawkes, C. V., Kivlin, S. N., Du, J. & Eviner, V. T. The temporal development and additivity of plant-soil feedback in perennial grasses. Plant Soil 369, 141–150 (2013).CAS 
    Article 

    Google Scholar 
    48.Latz, E., Eisenhauer, N., Rall, B. C., Scheu, S. & Jousset, A. Unravelling linkages between plant community composition and the pathogen-suppressive potential of soils. Sci. Rep. 6, 1–10 (2016).Article 
    CAS 

    Google Scholar 
    49.Chung, Y. A. & Rudgers, J. A. Plant–soil feedbacks promote negative frequency dependence in the coexistence of two aridland grasses. Proc. R. Soc. B Biol. Sci. 283 (2016).50.Mahaut, L., Fort, F., Violle, C. & Freschet, G. T. Multiple facets of diversity effects on plant productivity: species richness, functional diversity, species identity and intraspecific competition. Funct. Ecol. 34, 287–298 (2020).Article 

    Google Scholar 
    51.Barry, K. E. et al. Limited evidence for spatial resource partitioning across temperate grassland biodiversity experiments. Ecology 101, 2905 (2020).Article 

    Google Scholar 
    52.Hooper, D. U. et al. Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecol. Monogr. 75, 3–35 (2005).Article 

    Google Scholar 
    53.Pillai, P. & Gouhier, T. C. Not even wrong: the spurious measurement of biodiversity’s effects on ecosystem functioning. Ecology 100, e02645 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    54.Manning, P. et al. Transferring biodiversity-ecosystem function research to the management of ‘real-world’ ecosystems. Adv. Ecol. Res. 61, 323–356 (2019).Article 

    Google Scholar 
    55.Fargione, J. et al. From selection to complementarity: Shifts in the causes of biodiversity-productivity relationships in a long-term biodiversity experiment. Proc. R. Soc. B Biol. Sci. 274, 871–876 (2007).Article 

    Google Scholar 
    56.Helander, M. et al. Decreases mycorrhizal colonization and affects plant-soil feedback. Sci. Total Environ. 642, 285–291 (2018).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    57.Tilman, D. et al. Diversity and productivity in a long-term grassland experiment. Science 294, 843–845 (2001).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    58.Cadotte, M. W., Cavender-Bares, J., Tilman, D. & Oakley, T. H. Using phylogenetic, functional and trait diversity to understand patterns of plant community productivity. PLoS ONE 4, e5695 (2009).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    59.Kulmatiski, A., Heavilin, J. & Beard, K. H. Testing predictions of a three-species plant-soil feedback model. J. Ecol. 99, 542–550 (2011).
    Google Scholar 
    60.Kulmatiski, A., Beard, K. H., Grenzer, J., Forero, L. & Heavilin, J. Using plant-soil feedbacks to predict plant biomass in diverse communities. Ecology 97, 2064–2073 (2016).PubMed 
    Article 
    PubMed Central 

    Google Scholar  More

  • in

    Invasive Lactuca serriola seeds contain endophytic bacteria that contribute to drought tolerance

    1.Bulgarelli, D., Schlaeppi, K., Spaepen, S., van Themaat, E. V. L. & Schulze-Lefert, P. Structure and functions of the bacterial microbiota of plants. Annu. Rev. Plant Biol. 64, 807–838 (2013).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    2.Reinhold-Hurek, B. & Hurek, T. Living inside plants: bacterial endophytes. Curr. Opin. Plant Biol. 14, 435–443 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    3.Naveed, M., Mitter, B., Reichenauer, T. G., Wieczorek, K. & Sessitsch, A. Increased drought stress resilience of maize through endophytic colonization by Burkholderia phytofirmans PsJN and Enterobacter sp FD17. Environ. Exp. Bot. 97, 30–39 (2014).CAS 
    Article 

    Google Scholar 
    4.Santoyo, G., Moreno-Hagelsieb, G., del Carmen Orozco-Mosqueda, M. & Glick, B. R. Plant growth-promoting bacterial endophytes. Microbiol. Res. 183, 92–99 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    5.Ali, S., Charles, T. C. & Glick, B. R. Amelioration of high salinity stress damage by plant growth-promoting bacterial endophytes that contain ACC deaminase. Plant Physiol. Biochem. 80, 160–167 (2014).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    6.Weyens, N., van der Lelie, D., Taghavi, S. & Vangronsveld, J. Phytoremediation: plant–endophyte partnerships take the challenge. Curr. Opin. Biotechnol. 20, 248–254 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    7.Vandenkoornhuyse, P., Quaiser, A., Duhamel, M., Le Van, A. & Dufresne, A. The importance of the microbiome of the plant holobiont. New Phytol. 206, 1196–1206 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    8.Card, S. D. et al. Beneficial endophytic microorganisms of Brassica—A review. Biol. Control 90, 102–112 (2015).Article 

    Google Scholar 
    9.Shahzad, R., Khan, A. L., Bilal, S., Asaf, S. & Lee, I. J. What Is there in seeds? Vertically transmitted endophytic resources for sustainable improvement in plant growth. Front. Plant Sci. 9, 24. https://doi.org/10.3389/fpls.2018.00024 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    10.Truyens, S., Weyens, N., Cuypers, A. & Vangronsveld, J. Bacterial seed endophytes: genera, vertical transmission and interaction with plants. Environ. Microbiol. Rep. 7, 40–50 (2015).Article 

    Google Scholar 
    11.Catford, J. A., Jansson, R. & Nilsson, C. Reducing redundancy in invasion ecology by integrating hypotheses into a single theoretical framework. Divers. Distrib. 15, 22–40 (2009).Article 

    Google Scholar 
    12.van Kleunen, M., Dawson, W. & Maurel, N. Characteristics of successful alien plants. Mol. Ecol. 24, 1954–1968 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    13.Coats, V. C. & Rumpho, M. E. The rhizosphere microbiota of plant invaders: an overview of recent advances in the microbiomics of invasive plants. Front. Microbiol. 5, 368. https://doi.org/10.3389/fmicb.2014.00368 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    14.Richardson, D. M., Allsopp, N., D’antonio, C. M., Milton, S. J. & Rejmánek, M. Plant invasions—the role of mutualisms. Biol. Rev. 75, 65–93 (2000).CAS 
    PubMed 
    Article 

    Google Scholar 
    15.Pringle, A. et al. Mycorrhizal symbioses and plant invasions. Annu. Rev. Ecol. Evol. Syst. 40, 699–715 (2009).Article 

    Google Scholar 
    16.Sun, Z.-K. & He, W.-M. Evidence for enhanced mutualism hypothesis: Solidago canadensis plants from regular soils perform better. PLoS ONE 5, e15418. https://doi.org/10.1371/journal.pone.0015418 (2010).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    17.Kowalski, K. P. et al. Advancing the science of microbial symbiosis to support invasive species management: a case study on Phragmites in the Great Lakes. Front. Microbiol. 6, 95. https://doi.org/10.3389/fmicb.2015.00095 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    18.Dai, Z. C. et al. Different growth promoting effects of endophytic bacteria on invasive and native clonal plants. Front. Plant Sci. 7, 706. https://doi.org/10.3389/fpls.2016.00706 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    19.Rout, M. E. et al. Bacterial endophytes enhance competition by invasive plants. Am. J. Bot. 100, 1726–1737 (2013).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    20.Soares, M. A. et al. Functional role of bacteria from invasive Phragmites australis in promotion of host growth. Microb. Ecol. 72, 407–417 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Kim, Y.-H., Kil, J.-H., Hwang, S.-M. & Lee, C.-W. Spreading and distribution of Lactuca scariola, invasive alien plant, by habitat types in Korea. Weed Turfgrass Sci. 2, 138–151 (2013).Article 

    Google Scholar 
    22.Moon, S.-I. et al. Isolation and characterization of bio-active materials from prickly lettuce (Lactuca serriola). J. Life Sci. 19, 206–212 (2009).Article 

    Google Scholar 
    23.Lebeda, A. et al. Acquisition and ecological characterization of Lactuca serriola L germplasm collected in the Czech Republic, Germany, the Netherlands and United Kingdom. Genet. Resour. Crop Evol. 54, 555–562 (2007).Article 

    Google Scholar 
    24.Mallory-Smith, C. A., Thill, D. C. & Dial, M. J. Identification of sulfonylurea herbicide-resistant prickly lettuce (Lactuca serriola). Weed Technol. 4, 163–168 (1990).Article 

    Google Scholar 
    25.Glick, B. R. Plant growth-promoting bacteria: mechanisms and applications. Scientifica 2012, 963401. https://doi.org/10.6064/2012/963401 (2012).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    26.Costa, O. Y. A., Raaijmakers, J. M. & Kuramae, E. E. Microbial extracellular polymeric substances: ecological function and impact on soil aggregation. Front. Microbiol. 9, 1636. https://doi.org/10.3389/fmicb.2018.01636 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    27.Alami, Y., Achouak, W., Marol, C. & Heulin, T. Rhizosphere soil aggregation and plant growth promotion of sunflowers by an exopolysaccharide-producing Rhizobiums strain isolated from sunflower roots. Appl. Environ. Microbiol. 66, 3393–3398 (2000).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    28.Sandhya, V., Grover, M., Reddy, G. & Venkateswarlu, B. Alleviation of drought stress effects in sunflower seedlings by the exopolysaccharides producing Pseudomonas putida strain GAP-P45. Biol. Fertility Soils 46, 17–26 (2009).CAS 
    Article 

    Google Scholar 
    29.Vardharajula, S. Exopolysaccharide production by drought tolerant Bacillus spp and effect on soil aggregation under drought stress. J. Microbiol. Biotechnol. Food Sci. 9, 51–57 (2020).
    Google Scholar 
    30.Kumar, S., Stecher, G. & Tamura, K. MEGA7: molecular evolutionary genetics analysis version 70 for bigger datasets. Mol. Biol. Evol. 33, 1870–1874 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    31.Kang, S. H. et al. Two bacterial entophytes eliciting both plant growth promotion and plant defense on pepper (Capsicum annuum L). J. Microbiol. Biotechnol. 17, 96–103 (2007).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    32.Panwar, M., Tewari, R. & Nayyar, H. Native halo-tolerant plant growth promoting rhizobacteria Enterococcus and Pantoea sp. improve seed yield of Mungbean (Vigna radiata L) under soil salinity by reducing sodium uptake and stress injury. Physiol. Mol. Biol. Plants 22, 445–459 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    33.Selvakumar, G. et al. Characterization of a cold-tolerant plant growth-promoting bacterium Pantoea dispersa 1A isolated from a sub-alpine soil in the North Western Indian Himalayas. World J. Microbiol. Biotechnol. 24, 955–960 (2008).CAS 
    Article 

    Google Scholar 
    34.Egamberdieva, D. et al. High incidence of plant growth-stimulating bacteria associated with the rhizosphere of wheat grown on salinated soil in Uzbekistan. Environ. Microbiol. 10, 1–9 (2008).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    35.Pereira, S., Castro, P. & Research, P. Diversity and characterization of culturable bacterial endophytes from Zea mays and their potential as plant growth-promoting agents in metal-degraded soils. Environ. Sci. Pollut. Res. 21, 14110–14123 (2014).CAS 
    Article 

    Google Scholar 
    36.Sun, Z. et al. IAA producing Bacillus altitudinis alleviates iron stress in Triticum aestivum L seedling by both bioleaching of iron and up-regulation of genes encoding ferritins. Plant Soil 419, 1–11 (2017).CAS 
    Article 

    Google Scholar 
    37.Pierik, R., Tholen, D., Poorter, H., Visser, E. J. W. & Voesenek, L. A. C. J. The Janus face of ethylene: growth inhibition and stimulation. Trends Plant Sci. 11, 176–183 (2006).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    38.Glick, B. R. Bacteria with ACC deaminase can promote plant growth and help to feed the world. Microbiol. Res. 169, 30–39 (2014).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    39.Cardinale, M., Grube, M., Erlacher, A., Quehenberger, J. & Berg, G. Bacterial networks and co-occurrence relationships in the lettuce root microbiota. Environ. Microbiol. 17, 239–252 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    40.Yu, Y.-C., Yum, S.-J., Jeon, D.-Y. & Jeong, H.-G. Analysis of the microbiota on lettuce (Lactuca sativa L.) cultivated in South Korea to identify foodborne pathogens. J. Microbiol. Biotechnol. 28, 1318–1331 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    41.Brady, C. et al. Isolation of Enterobacter cowanii from Eucalyptus showing symptoms of bacterial blight and dieback in Uruguay. Lett. Appl. Microbiol. 49, 461–465 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.Chimwamurombe, P. M., Grönemeyer, J. L. & Reinhold-Hurek, B. Isolation and characterization of culturable seed-associated bacterial endophytes from gnotobiotically grown Marama bean seedlings. FEMS Microbiol. Ecol. 92, fiw083. https://doi.org/10.1093/femsec/fiw083 (2016).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    43.Gao, H. et al. Production exopolysaccharide from Kosakonia cowanii LT-1 through solid-state fermentation and its application as a plant growth promoter. Int. J. Biol. Macromol. 150, 955–964 (2020).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    44.Wang, L. et al. Development of sugarcane resource for efficient fermentation of exopolysaccharide by using a novel strain of Kosakonia cowanii LT-1. Bioresour. Technol. 280, 247–254 (2019).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    45.Borlee, B. R. et al. Pseudomonas aeruginosa uses a cyclic-di-GMP-regulated adhesin to reinforce the biofilm extracellular matrix. Mol. Microbiol. 75, 827–842 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    46.Huang, X.-F. et al. Mitsuaria sp. and Burkholderia sp. from Arabidopsis rhizosphere enhance drought tolerance in Arabidopsis thaliana and maize (Zea mays L.). Plant Soil 419, 523–539 (2017).CAS 
    Article 

    Google Scholar 
    47.Marulanda, A., Barea, J.-M. & Azcón, R. Stimulation of plant growth and drought tolerance by native microorganisms (AM fungi and bacteria) from dry environments: mechanisms related to bacterial effectiveness. J. Plant Growth Regul. 28, 115–124 (2009).CAS 
    Article 

    Google Scholar 
    48.Niu, X., Song, L., Xiao, Y. & Ge, W. Drought-tolerant plant growth-promoting rhizobacteria associated with foxtail millet in a semi-arid agroecosystem and their potential in alleviating drought stress. Front. Microbiol. 8, 2580. https://doi.org/10.3389/fmicb.2017.02580 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    49.Sandhya, V., Ali, S. Z., Grover, M., Reddy, G. & Venkateswarlu, B. Effect of plant growth promoting Pseudomonas spp. on compatible solutes, antioxidant status and plant growth of maize under drought stress. Plant Growth Regulat. 62, 21–30 (2010).CAS 
    Article 

    Google Scholar 
    50.Chen, C. et al. Pantoea alhagi, a novel endophytic bacterium with ability to improve growth and drought tolerance in wheat. Sci. Rep. 7, 41564. https://doi.org/10.1038/srep41564 (2017).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    51.Mönchgesang, S. et al. Natural variation of root exudates in Arabidopsis thaliana-linking metabolomic and genomic data. Sci. Rep. 6, 1–11 (2016).Article 
    CAS 

    Google Scholar 
    52.Zhang, N. et al. Effects of different plant root exudates and their organic acid components on chemotaxis, biofilm formation and colonization by beneficial rhizosphere-associated bacterial strains. Plant Soil 374, 689–700 (2014).CAS 
    Article 

    Google Scholar 
    53.Johnston-Monje, D. & Raizada, M. N. Conservation and diversity of seed associated endophytes in Zea across boundaries of evolution, ethnography and ecology. PLoS ONE 6, e20396. https://doi.org/10.1371/journal.pone.0020396 (2011).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    54.Coombs, J. T. & Franco, C. M. M. Isolation and identification of Actinobacteria from surface-sterilized wheat roots. Appl. Environ. Microbiol. 69, 5603–5608. https://doi.org/10.1128/aem.69.9.5603-5608.2003 (2003).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    55.Kimura, M. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J. Mol. Evol. 16, 111–120 (1980).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    56.Saitou, N. & Nei, M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol. Biol. Evol. 4, 406–425 (1987).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    57.Mehta, S. & Nautiyal, C. S. An efficient method for qualitative screening of phosphate-solubilizing bacteria. Curr. Microbiol. 43, 51–56 (2001).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    58.Milagres, A. M., Machuca, A. & Napoleao, D. Detection of siderophore production from several fungi and bacteria by a modification of chrome azurol S (CAS) agar plate assay. J. Microbiol. Methods 37, 1–6 (1999).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    59.Dworkin, M. & Foster, J. Experiments with some microorganisms which utilize ethane and hydrogen. J. Bacteriol. 75, 592–603 (1958).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    60.Singh, J. K., Adams, F. G. & Brown, M. H. Diversity and function of capsular polysaccharide in Acinetobacter baumannii. Front. Microbiol. 9, 3301 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    61.Polak-Berecka, M., Waśko, A., Skrzypek, H. & Kreft, A. Production of exopolysaccharides by a probiotic strain of Lactobacillus rhamnosus: biosynthesis and purification methods. Acta Aliment. 42, 220–228 (2013).CAS 
    Article 

    Google Scholar 
    62.Tschaplinski, T. J. et al. The nature of the progression of drought stress drives differential metabolomic responses in Populus deltoides. Ann. Bot. 124, 617–626 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    63.Michel, B. E. & Kaufmann, M. R. The osmotic potential of polyethylene glycol 6000. Plant Physiol. 51, 914–916 (1973).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    64.Hanna, A., Berg, M., Stout, V. & Razatos, A. Role of capsular colanic acid in adhesion of uropathogenic Escherichia coli. Appl. Environ. Microbiol. 69, 4474–4481 (2003).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    65.Liu, S.-B. et al. Structure and ecological roles of a novel exopolysaccharide from the Arctic sea ice bacterium Pseudoalteromonas sp. strain SM20310. Appl. Environ. Microbiol. 79, 224–230 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    66.Dubois, M., Gilles, K. A., Hamilton, J. K., Rebers, P. & Smith, F. Colorimetric method for determination of sugars and related substances. Analyt. Chem. 28, 350–356 (1956).CAS 
    Article 

    Google Scholar 
    67.Yahaghi, Z., Shirvani, M., Nourbakhsh, F. & Pueyo, J. J. Uptake and effects of lead and zinc on alfalfa (Medicago sativa L.) seed germination and seedling growth: Role of plant growth promoting bacteria. S. Afr. J. Bot. 124, 573–582 (2019).CAS 
    Article 

    Google Scholar 
    68.Zhang, Z. & Huang, R. Analysis of malondialdehyde, chlorophyll proline, soluble sugar, and glutathione content in Arabidopsis seedling. Bio-Protoc. 3, e817 (2013).
    Google Scholar 
    69.Türkan, I., Bor, M., Özdemir, F. & Koca, H. Differential responses of lipid peroxidation and antioxidants in the leaves of drought-tolerant P. acutifolius Gray and drought-sensitive P. vulgaris L. subjected to polyethylene glycol mediated water stress. Plant Sci. 168, 223–231 (2005).Article 
    CAS 

    Google Scholar  More

  • in

    SPX-related genes regulate phosphorus homeostasis in the marine phytoplankton, Phaeodactylum tricornutum

    SPX gene and CRISPR/Cas9 knockoutUsing the SPX domain as a query to search in the P. tricornutum genome, we found a total of six genes that harbor an SPX domain (Supplementary Table 1), including Vpt1 and Vtc4 that were recently described by Dell’Aquila et al.21. Pfam analysis of the six identified sequences in P. tricornutum resulted in the identification of the SPX domain in these proteins, which shared several conserved sites with land plants’ SPX domains (Fig. 1a, b). Phylogenetic analyses further verified the high similarity of these sequences with other known SPX domain proteins (Supplementary Fig. 1). One of these genes possesses a SPX domain as the sole functional domain (named SPX, and its encoding gene named SPX gene, from here on) while the other five (including Vpt1 and Vtc4) contain at least one other domain. From our recently published transcriptome dataset27, we found that SPX, Vpt1, and Vtc4 genes were differentially expressed (log2 Fold Change > 1 and adjusted p value  More

  • in

    Coral distribution and bleaching vulnerability areas in Southwestern Atlantic under ocean warming

    1.Hoegh-Guldberg, O. et al. Coral reefs under rapid climate change and ocean acidification. Science 318, 1737–1742 (2007).CAS 
    PubMed 
    Article 
    ADS 

    Google Scholar 
    2.Frölicher, T. L., Fischer, E. M. & Gruber, N. Marine heatwaves under global warming. Nature 560, 360–364 (2018).PubMed 
    Article 
    ADS 
    CAS 

    Google Scholar 
    3.Oliver, E. C. J. et al. Projected marine heatwaves in the 21st century and the potential for ecological impact. Front. Mar. Sci. 6, 1–12 (2019).Article 

    Google Scholar 
    4.LaJeunesse, T. C. et al. Systematic revision of symbiodiniaceae highlights the antiquity and diversity of coral endosymbionts. Curr. Biol. 28, 2570-2580.e6 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    5.Hoegh-Guldberg, O. Climate change, coral bleaching and the future of the world’ s coral reefs. Ove Hoegh-Guldberg (1998).6.Hughes, T. P. et al. Spatial and temporal patterns of mass bleaching of corals in the Anthropocene. Science (80-.) 359, 80–83 (2018).CAS 
    Article 
    ADS 

    Google Scholar 
    7.Van Hooidonk, R., Maynard, J. A., Manzello, D. & Planes, S. Opposite latitudinal gradients in projected ocean acidification and bleaching impacts on coral reefs. Glob. Chang. Biol. 20, 103–112 (2014).PubMed 
    Article 
    ADS 

    Google Scholar 
    8.Muller, E. M., Rogers, C. S., Spitzack, A. S. & Van Woesik, R. Bleaching increases likelihood of disease on Acropora palmata (Lamarck) in Hawksnest Bay, St John, US Virgin Islands. Coral Reefs 27, 191–195 (2008).Article 
    ADS 

    Google Scholar 
    9.Cróquer, A. & Weil, E. Changes in Caribbean coral disease prevalence after the 2005 bleaching event. Dis. Aquat. Organ. 87, 33–43 (2009).PubMed 
    Article 

    Google Scholar 
    10.Grottoli, A. G. et al. The cumulative impact of annual coral bleaching can turn some coral species winners into losers. Glob. Chang. Biol. 20, 3823–3833 (2014).PubMed 
    Article 
    ADS 

    Google Scholar 
    11.Schoepf, V. et al. Annual coral bleaching and the long-term recovery capacity of coral. Proc. R. Soc. B Biol. Sci. 282, 20151887 (2015).Article 
    CAS 

    Google Scholar 
    12.Neal, B. P. et al. Caribbean massive corals not recovering from repeated thermal stress events during 2005–2013. Ecol. Evol. 7, 1339–1353 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    13.Alvarez-Filip, L., Dulvy, N. K., Gill, J. A., Côté, I. M. & Watkinson, A. R. Flattening of Caribbean coral reefs: Region-wide declines in architectural complexity. Proc. R. Soc. B Biol. Sci. 276, 3019–3025 (2009).Article 

    Google Scholar 
    14.Pratchett, M. S., Hoey, A. S., Wilson, S. K., Messmer, V. & Graham, N. A. J. Changes in biodiversity and functioning of reef fish assemblages following coral bleaching and coral loss. Diversity 3, 424–452 (2011).Article 

    Google Scholar 
    15.Poloczanska, E. S. et al. Responses of marine organisms to climate change across oceans. Front. Mar. Sci. 3, 1–21 (2016).Article 

    Google Scholar 
    16.Pinsky, M. L., Selden, R. L. & Kitchel, Z. J. Climate-driven shifts in marine species ranges: Scaling from organisms to communities. Ann. Rev. Mar. Sci. 12, 153–179 (2020).PubMed 
    Article 

    Google Scholar 
    17.Kleypas, J. A., McManu, J. W. & Mene, L. A. B. Environmental limits to coral reef development: Where do we draw the line?. Am. Zool. 39, 146–159 (1999).Article 

    Google Scholar 
    18.Descombes, P. et al. Forecasted coral reef decline in marine biodiversity hotspots under climate change. Glob. Chang. Biol. 21, 2479–2487 (2015).PubMed 
    Article 
    ADS 

    Google Scholar 
    19.Vergés, A. et al. The tropicalization of temperate marine ecosystems: Climate-mediated changes in herbivory and community phase shifts. Proc. R. Soc. B Biol. Sci. 281, 20140846 (2014).Article 

    Google Scholar 
    20.Mies, M. et al. South Atlantic coral reefs are major global warming refugia and less susceptible to bleaching. Front. Mar. Sci. 7, 1–13 (2020).Article 
    ADS 

    Google Scholar 
    21.Perry, C. T. & Larcombe, P. Marginal and non-reef-building coral environments. Coral Reefs 22, 427–432 (2003).Article 

    Google Scholar 
    22.Loiola, M. et al. Structure of marginal coral reef assemblages under different turbidity regime. Mar. Environ. Res. 147, 138–148 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    23.Beger, M., Sommer, B., Harrison, P. L., Smith, S. D. A. & Pandolfi, J. M. Conserving potential coral reef refuges at high latitudes. Divers. Distrib. 20, 245–257 (2014).Article 

    Google Scholar 
    24.Glynn, P. W. Coral reef bleaching: Facts, hypotheses and implications. Glob. Chang. Biol. 2, 495–509 (1996).Article 
    ADS 

    Google Scholar 
    25.Semmler, R. F., Hoot, W. C. & Reaka, M. L. Are mesophotic coral ecosystems distinct communities and can they serve as refugia for shallow reefs?. Coral Reefs 36, 433–444 (2017).Article 
    ADS 

    Google Scholar 
    26.Rocha, L. A. et al. Mesophotic coral ecosystems are threatened and ecologically distinct from shallow water reefs. Science (8-.) 361, 281–284 (2018).CAS 
    Article 
    ADS 

    Google Scholar 
    27.Eckert, R. J., Studivan, M. S. & Voss, J. D. Populations of the coral species Montastraea cavernosa on the Belize Barrier Reef lack vertical connectivity. Sci. Rep. 9, 1–11 (2019).CAS 
    Article 

    Google Scholar 
    28.Serrano, X. M. et al. Geographic differences in vertical connectivity in the Caribbean coral Montastraea cavernosa despite high levels of horizontal connectivity at shallow depths. Mol. Ecol. 23, 4226–4240 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    29.Serrano, X. M. et al. Long distance dispersal and vertical gene flow in the Caribbean brooding coral Porites astreoides. Sci. Rep. 6, 1–12 (2016).Article 
    CAS 

    Google Scholar 
    30.Morais, J. & Santos, B. A. Limited potential of deep reefs to serve as refuges for tropical Southwestern Atlantic corals. Ecosphere 9, e02281 (2018).Article 

    Google Scholar 
    31.Torda, G. et al. Rapid adaptive responses to climate change in corals. Nat. Clim. Chang. https://doi.org/10.1038/nclimate3374 (2017).Article 

    Google Scholar 
    32.Danielle, C. et al. Dynamic symbioses reveal pathways to coral survival through prolonged heatwaves. Nat Commun. 11(1), https://doi.org/10.1038/s41467-020-19169-y. (2020)33.D’Angelo, C. & Wiedenmann, J. Impacts of nutrient enrichment on coral reefs: New perspectives and implications for coastal management and reef survival. Curr. Opin. Environ. Sustain. 7, 82–93 (2014).Article 

    Google Scholar 
    34.Donovan, M.K.et al. Local conditions magnify coral loss after marine heatwaves. Science 372(6545), 977–980. https://doi.org/10.1126/science.abd9464. (2021)CAS 
    PubMed 
    Article 

    Google Scholar 
    35.Hughes, T. et al. Climate change, Human impacts, and the resilience of coral reefs. Laser Induced Damage Opt. Mater. 2009 7504, 75041H (2003).36.Carilli, J. E., Norris, R. D., Black, B. A., Walsh, S. M. & McField, M. Local stressors reduce coral resilience to bleaching. PLoS ONE 4, 1–5 (2009).Article 
    CAS 

    Google Scholar 
    37.Donner, S. D., Heron, S. F. & Skirving, W. J. Future scenarios: A review of modelling efforts to predict the future of coral reefs in an era of climate change. 159–173. https://doi.org/10.1007/978-3-540-69775-6_10 (2018).38.McLeod, E. et al. Warming seas in the coral triangle: Coral reef vulnerability and management implications. Coast. Manag. 38, 518–539 (2010).Article 

    Google Scholar 
    39.Maynard, J. A. et al. Vulnerability to coral reefs. 1–8 (2019).40.Leão, Z. M. A. N., Kikuchi, R. K. P. & Testa, V. Corals and coral reefs of Brazil. Latin Am. Coral Reefs https://doi.org/10.1016/B978-044451388-5/50003-5 (2003).Article 

    Google Scholar 
    41.Leão, Z. M. A. N. & Kikuchi, R. K. P. A relic coral fauna threatened by global changes and human activities, Eastern Brazil. Mar. Pollut. Bull. 51, 599–611 (2005).PubMed 
    Article 
    CAS 

    Google Scholar 
    42.Francini-Filho, R. B. & De Moura, R. L. Dynamics of fish assemblages on coral reefs subjected to different management regimes in the Abrolhos Bank, eastern Brazil. Aquat. Conserv. Mar. Freshw. Ecosyst. 18, 1166–1179 (2008).Article 

    Google Scholar 
    43.Moura, R. L. et al. Spatial patterns of benthic megahabitats and conservation planning in the Abrolhos Bank. Cont. Shelf Res. 70, 109–117 (2013).Article 
    ADS 

    Google Scholar 
    44.Vergés, A. et al. Tropicalisation of temperate reefs: Implications for ecosystem functions and management actions. Funct. Ecol. 33, 1000–1013 (2019).Article 

    Google Scholar 
    45.Yamano, H., Sugihara, K. & Nomura, K. Rapid poleward range expansion of tropical reef corals in response to rising sea surface temperatures. Geophys. Res. Lett. 38, (2011).46.Precht, W. F. & Aronson, R. B. Climate flickers and range shifts of reef corals. Front. Ecol. Environ. 2, 307 (2004).Article 

    Google Scholar 
    47.Aued, A. W. et al. Large-scale patterns of benthic marine communities in the brazilian province. PLoS ONE 13, 1–15 (2018).Article 
    CAS 

    Google Scholar 
    48.Sully, S., Burkepile, D. E., Donovan, M. K., Hodgson, G. & van Woesik, R. A global analysis of coral bleaching over the past two decades. Nat. Commun. 10, 1–5 (2019).CAS 
    Article 

    Google Scholar 
    49.Phillips, N. A Companion to the e-Book “YaRrr!: The Pirate’s Guide to R”. (2017).50.R Core Team. R: A Language and Environment for Statistical Computing. (2020).51.Banha, T. N. S. et al. Low coral mortality during the most intense bleaching event ever recorded in subtropical Southwestern Atlantic reefs. Coral Reefs https://doi.org/10.1007/s00338-019-01856-y (2019).Article 

    Google Scholar 
    52.Oliveira, U. D. R., Gomes, P. B., Cordeiro, R. T. S., De Lima, G. V. & Pérez, C. D. Modeling impacts of climate change on the potential habitat of an endangered Brazilian endemic coral: Discussion about deep sea refugia. PLoS ONE 14, 1–24 (2019).
    Google Scholar 
    53.Cowen, R. K. & Sponaugle, S. Larval dispersal and marine population connectivity. Ann. Rev. Mar. Sci. 1, 443–466 (2009).PubMed 
    Article 

    Google Scholar 
    54.Price, N. N. et al. Global biogeography of coral recruitment: Tropical decline and subtropical increase. Mar. Ecol. Prog. Ser. 621, 1–17 (2019).Article 
    ADS 

    Google Scholar 
    55.Cacciapaglia, C. & van Woesik, R. Reef-coral refugia in a rapidly changing ocean. Glob. Chang. Biol. 21, 2272–2282 (2015).PubMed 
    Article 
    ADS 

    Google Scholar 
    56.Baird, A. H. et al. A decline in bleaching suggests that depth can provide a refuge from global warming in most coral taxa. Mar. Ecol. Prog. Ser. 603, 257–264 (2018).Article 
    ADS 

    Google Scholar 
    57.Bongaerts, P., Ridgway, T., Sampayo, E. M. & Hoegh-Guldberg, O. Assessing the ‘deep reef refugia’ hypothesis: Focus on Caribbean reefs. Coral Reefs 29, 1–19 (2010).Article 

    Google Scholar 
    58.Coles, S. L. et al. Evidence of acclimatization or adaptation in Hawaiian corals to higher ocean temperatures. PeerJ 2018, 1–24 (2018).
    Google Scholar 
    59.Bay, R. A., Rose, N. H., Logan, C. A. & Palumbi, S. R. Genomic models predict successful coral adaptation if future ocean warming rates are reduced. Sci. Adv. 3, 1–10 (2017).Article 

    Google Scholar 
    60.Wooldridge, S., Done, T., Berkelmans, R., Jones, R. & Marshall, P. Precursors for resilience in coral communities in a warming climate: A belief network approach. Mar. Ecol. Prog. Ser. 295, 157–169 (2005).Article 
    ADS 

    Google Scholar 
    61.Mazzei, E. F. et al. Newly discovered reefs in the southern Abrolhos Bank, Brazil: Anthropogenic impacts and urgent conservation needs. Mar. Pollut. Bull. 114, 123–133 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    62.Duarte, G. A. S. et al. Heat waves are a major threat to turbid coral reefs in Brazil. Front. Mar. Sci. 7, 179 (2020).Article 

    Google Scholar 
    63.Ferreira L.C. et al. Different responses of massive and branching corals to a major heatwave at the largest and richest reef complex in South Atlantic. Mar. Biol. 168(5), https://doi.org/10.1007/s00227-021-03863-6. (2021)64.Teixeira, C. D. et al. Sustained mass coral bleaching (2016–2017) in Brazilian turbid-zone reefs: taxonomic, cross-shelf and habitat-related trends. Coral Reefs https://doi.org/10.1007/s00338-019-01789-6 (2019).Article 

    Google Scholar 
    65.França, F. M. et al. Climatic and local stressor interactions threaten tropical forests and coral reefs. Philos. Trans. R. Soc. B Biol. Sci. 375, 20190116 (2020).Article 

    Google Scholar 
    66.Mumby, P. J. & Harborne, A. R. Marine reserves enhance the recovery of corals on Caribbean reefs. PLoS ONE 5, 1–7 (2010).Article 
    CAS 

    Google Scholar 
    67.Kwiatkowski, L. et al. Twenty-first century ocean warming, acidification, deoxygenation, and upper ocean nutrient decline from CMIP6 model projections. Biogeosci. Discuss. https://doi.org/10.5194/bg-2020-16 (2020).68.Jokiel, P. L. Evaluating the assumptions involved. ICES J. Mar. Sci. 73, 550–557 (2015).Article 

    Google Scholar 
    69.Matz, M. V., Treml, E. A. & Haller, B. C. Estimating the potential for coral adaptation to global warming across the Indo-West Pacific. Glob. Chang. Biol. 26, 3473–3481 (2020).PubMed 
    Article 
    ADS 

    Google Scholar 
    70.Tyberghein, L. et al. Bio-ORACLE: A global environmental dataset for marine species distribution modelling. Glob. Ecol. Biogeogr. 21, 272–281 (2012).Article 

    Google Scholar 
    71.Assis, J. et al. Bio-ORACLE v2.0: Extending marine data layers for bioclimatic modelling. Glob. Ecol. Biogeogr. 27, 277–284 (2018).Article 

    Google Scholar 
    72.Sbrocco, E. J. & Barber, P. H. MARSPEC: Ocean climate layers for marine spatial ecology. Ecology 94, 979–979 (2013).Article 

    Google Scholar 
    73.Hijmans, J. R. et al. Package ‘ raster ’ R topics documented (2016).74.Sappington, J. M., Longshore, K. M. & Thompson, D. B. Quantifying landscape ruggedness for animal habitat analysis: A case study using bighorn sheep in the Mojave desert. J. Wildl. Manag. 71, 1419–1426 (2007).Article 

    Google Scholar 
    75.IPCC. Climate Change 2014 Part A: Global and Sectoral Aspects. Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (2014).76.Fox, J. & Weisberg, S. Multivariate Linear Models in R. An R Companion to Appl. Regres. 1–31 (2011).77.Zuur, A. F., Ieno, E. N., Walker, N. J., Saveliev, A. A. & Smith, G. M. Required Pre-knowledge : A Linear Regression. Mixed Effects Models and Extensions in Ecology with R 1, (2009).78.Martins, T. G., Simpson, D., Lindgren, F. & Rue, H. Bayesian computing with INLA: New features. (2012).79.Rue, H., Martino, S. & Chopin, N. Approximate Bayesian inference for latent Gaussian models by using integrated nested Laplace approximations. J. R. Stat. Soc. Ser. B Stat. Methodol. 71, 319–392 (2009).MathSciNet 
    MATH 
    Article 

    Google Scholar 
    80.Lindgren, F., Rue, H. & Lindström, J. An explicit link between Gaussian fields and Gaussian MARKOV random fields: The stochastic partial differential equation approach. J. R. Stat. Soc. Ser. B Stat. Methodol. 73, 423–498 (2011).MathSciNet 
    MATH 
    Article 

    Google Scholar 
    81.Muñoz, F., Pennino, M. G., Conesa, D., López-Quílez, A. & Bellido, J. M. Estimation and prediction of the spatial occurrence of fish species using Bayesian latent Gaussian models. Stoch. Environ. Res. Risk Assess. 27, 1171–1180 (2013).Article 

    Google Scholar 
    82.Held, L., Schrödle, B. & Rue, H. Posterior and cross-validatory predictive checks: A comparison of MCMC and INLA. Stat. Model. Regres. Struct. Festschrift Honour Ludwig Fahrmeir 1–20 (2010). https://doi.org/10.1007/978-3-7908-2413-183.Watanabe, S. Asymptotic equivalence of Bayes cross validation and widely applicable information criterion in singular learning theory. J. Mach. Learn. Res. 11, 3571–3594 (2010).MathSciNet 
    MATH 

    Google Scholar 
    84.Roos, M. & Held, L. Sensitivity analysis in Bayesian generalized linear mixed models for binary data. Bayesian Anal. 6, 259–278 (2011).MathSciNet 
    MATH 
    Article 

    Google Scholar 
    85.Fonseca, V. P., Pennino, M. G., de Nóbrega, M. F., Oliveira, J. E. L. & de Figueiredo Mendes, L. Identifying fish diversity hot-spots in data-poor situations. Mar. Environ. Res. 129, 365–373 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    86.Pennino, M. G., Vilela, R., Bellido, J. M. & Velasco, F. Balancing resource protection and fishing activity: The case of the European hake in the northern Iberian Peninsula. Fish. Oceanogr. 28, 54–65 (2019).Article 

    Google Scholar 
    87.Martínez-Minaya, J. et al. A hierarchical Bayesian Beta regression approach to study the effects of geographical genetic structure and spatial autocorrelation on species distribution range shifts. Mol. Ecol. Resour. 19, 929–943 (2019).PubMed 
    Article 

    Google Scholar  More

  • in

    Divide-and-conquer: machine-learning integrates mammalian and viral traits with network features to predict virus-mammal associations

    Our framework to predict unknown associations between known viruses and potential mammalian hosts or susceptible species comprised three distinct perspectives: viral, mammalian and network. Each perspective produced predictions from a unique vantage point (that of each virus, each mammal, and the network connecting them respectively). Subsequently, their results were consolidated via majority voting. This approach suggested that 20,832 (median, 90% CI = [2,736, 97,062], hereafter values in square brackets represent 90% CI) unknown associations potentially exist between our mammals and their known viruses, (18,920 [2,440, 91,517] in wild or semi-domesticated mammals). Number of unknown associations predicted by each perspective individually were as follows: mammalian only = 41,537 [4,275, 23,8971], viral only = 21,352 [2,536, 95,630], and network only = 76,081 [27,738, 20,5814]. Our results indicated a ~4.29-fold increase ([~1.43, ~16.33]) in virus-mammal associations (~4.89 [~1.5, ~19.81] in wild and semi-domesticated mammals).Additionally, we trained an independent pipeline including only the 3534 supported by evidence extracted from meta-data accompanying nucleotide sequences, as indexed in EID2 (55.82% of all associations – see Methods section and Supplementary Results 8). Our sequence-evidence pipeline indicated that 15,721 (median, 90% CI = [1,603, 88,553]) unknown associations could potentially exist (13,930 [1,298, 83,043] in wild or semi-domesticated mammals).In the following subsections we first illustrate the mechanism of our framework via an example, then further explore the predictive power of our approach for viruses and mammals.ExampleOur multi-perspective framework generates predictions for each known or unknown virus-mammal association (2,722,656 possible associations between 1,896 viruses and 1,436 terrestrial mammals). We highlight this functionality using two examples (Fig. 1). West Nile virus (WNV) a flavivirus with wide host range, and the bat Rousettus leschenaultia (order: Chiroptera). We first consider each of our perspectives separately, and then showcase how these perspectives are consolidated to produce final predictions.Fig. 1: Example showcasing final and intermediate predictions of West Nile Virus (WNV), and Rousettus leschenaultii.Panel A Top 60 predicted mammalian species susceptible to WNV. Mammals were ordered by mean probability of predictions derived from mammalian (all models), viral (WNV models) and network perspectives, and top 60 were selected. Circles represent the following information in order: 1) whether the association is known (documented in our sources) or not (potential or undocumented). Hosts are omitted for known associations. 2) Mean probability of the three perspectives (per association). 3) Median mammalian perspective probabilities of predicted associations. These probabilities are obtained from 3000 models (50 replicate models for each mammal), trained with viral features – SMOTE class balancing. 4) Median viral perspective probabilities of predicted associations (50 WNV replicate models trained with mammalian features – SMOTE class balancing). 5) Median network perspective probabilities of predicted associations (100 replicate models, balanced under-sampling). 6) Taxonomic order of predicted susceptible species. Orders are shortened as follows: Artiodactyla (Art), Carnivora (Crn), Chiroptera (Chp), primates (Prm), Rodentia (Rod), and Others (Oth). Panel B Top 50 predicted viruses of R. leschenaultii. Viruses were ordered by mean probability of predictions derived from mammalian (R. leschenaultii models), viral (all models) and network perspectives. Circles as per Panel A. Baltimore represents Baltimore classification. Panel C Median probability of predicted WNV-mammal associations in each of the three perspectives per mammalian order. Points represent susceptible species predicted by voting (at least two of the three perspectives – n = 137). Median ensemble probability is computed in each perspective (50 replicate models for each virus/mammal, 100 replicate network models). Predictions derived from each perspective at 0.5 probability cut-off. Supplementary Data 1 presents full WNV results. Panel D Median probability of virus-R. leschenaultii associations in the three perspectives per Baltimore group. Points represent susceptible species predicted by voting (at least two of the three perspectives – n = 64), predictions are derived as per panel C. Supplementary Data 2 lists full results for R. leschenaultii. Supplementary Fig. 7 illustrate the results when research effort into viruses and mammals is included in mammalian and viral perspectives, respectively.Full size image1) The mammalian perspective: our mammalian perspective models, trained with features expressing viral traits (Table 1), suggested a median of 90 [17, 410] unknown associations between WNV and terrestrial mammals could form when predicting virus-mammal associations based on viral features alone – a ~2.61-fold increase [~1.3, ~8.32]. Similarly, our results indicated that 64 [4, 331] new associations could form between our selected mammal (R. leschenaultia) and our viruses – a ~4.37-fold increase [~1.21, ~18.42] (Supplementary Results 4).Table 1 Viral traits & features used to build our mammalian models.Full size table(2) The viral perspective: our viral models, trained with features expressing mammalian traits (Table 2), indicated a median of 48 [0, 214] new hosts of WNV (~1.86- fold increase [~1, 4.82]). Results for our example mammal (R. leschenaultia) suggested 18 [3, 76], existing viruses could be found in this host (~1.95-fold increase [~1.16, ~5.00]) – Supplementary Results 5).Table 2 mammalian traits & features used to build our viral models.Full size table(3) The network perspective: Our network models indicated a median of 721 [448, 1,317] (~13.88 [9, 24.52] fold increase) unknown associations between WNV and terrestrial mammals, and that 246 [91, 336] existing viruses could be found in our selected host (R. leschenaultia), equivalent to a ~13.95 [~5.79, ~18.68] fold increase (Supplementary Results 6).Considering that each of the above perspectives approached the problem of predicting virus-mammal associations from a different angle, the agreement between these perspectives varied. In the case of WNV: mammalian and viral perspectives achieved 92.3% agreement [72.6%–98.5%]; mammals and network perspectives had 55.3% agreement [33.4%–69.5%]; and viruses and network had 52.9% agreement [19.8%–68.7%]. In the case of R. leschenaultia these numbers were as follows: 96.15% [82.44%, 99.58%], 87.24% [76.37%, 95.04%], and 87.61% [75.90%, 95.25%], respectively. The agreements between our perspectives across the 2,722,656 possible associations were as follows: 98.04% [90.36%, 99.73%] between mammalian and viral perspectives, 96.71% [88.62%, 98.92%] between mammalian and network perspectives, and 97.11% [91.57%, 98.95%] between viral and network perspectives.After voting, our framework suggested that a median of 117 [15, 509] new or undetected associations could be missing between WNV and terrestrial mammals (~3.45-fold increase [~1.3, ~12.2]). Similarly, our results indicated that R. leschenaultia could be susceptible to an additional 45 [5, 235] viruses that were not captured in our input (~1.37-fold increase [~1.26, ~13.37]). Figure 1 illustrates top predicted and detected associations for WNV (Supplementary Data 1) and R. leschenaultia (Supplementary Data 2). Supplementary Results 1 illustrate results with research effort into viruses, and mammals included as a predictor in our mammalian and viral perspective models, respectively. Predictions with and without research effort incorporated into models trained in these perspectives broadly agreed.Relative importance of viral featuresOur multi-perspective approach trained a suite of models for each mammalian species with two or more known viruses (n = 699, response variable = 1 if the virus is known to associate with the focal mammalian species, 0 otherwise). This enabled us to assess the relative importance (influence) of viral traits (Table 1) to each of our mammalian models. This in turn showcased variations of how these viral traits contribute to the models at the level of individual species (e.g. humans), and at an aggregated level (e.g. by order or domestication status). The results, highlighted in Fig. 2A, indicate that mean phylogenetic (median = 95.4% [75.6%, 100%]) and mean ecological (90.90% [43.50%, 100%]) distances between potential and known hosts of each virus were the top predictors of associations between the focal host and each of the input viruses. Maximum phylogenetic breadth was also important (74.7 0%, [16.60%, 100%]).Fig. 2: Results (viruses).Panel A Variable importance (relative contribution) of viral traits to mammalian perspective models. Variable importance is calculated for each constituent ensemble (n = 699) of our mammalian perspective (median of a suite of 50 replicate models, trained with viral features, with SMOTE sampling), and then aggregated (mean) per each reported group (columns). Panel B – Number of known and new mammalian species associated with each virus. Rabies lyssavirus was excluded from panel B to allow for better visualisation. Top 40 (by number of new hosts) are labelled. Species in bold have over 150 predicted hosts (Supplementary Data 3 lists details of these viruses including CI). Panel C Predicted number of viruses per species of wild and semi-domesticated mammals (group by mammalian order). Following orders (clockwise) are presented: Artiodactyla, Carnivora, Chiroptera, Perissodactyla, Primates, and Rodentia. Source of the silhouette graphics is PhyloPic.org. (Supplementary Data 4 lists aggregated results per mammalian order). Circles represent each mammalian species (with predicted viruses > 0), coloured by number of known viruses previously not associated with this species. Boxplots indicate median (centre), the 25th and 75th percentiles (bounds of box) and inter quantile range (whiskers) and are aggregated at the order level. Large red circles with error bars (90% CI) illustrate the median number of known viruses per species in each order. Number of species presented (n) is as follows: All = 1293 (Artiodactyla = 104, Carnivora = 177, Chiroptera = 548, Perissodactyla = 11, Primates = 171, and Rodentia = 282); Group I = 666 (94, 109, 156, 10, 160, 137); Group II = 371 (32, 120, 111, 1, 54, 53); Group III = 410 (87,62,123,9,51,78); Group IV = 739 (98, 102, 221, 9, 148, 161); Group V = 1129 (87, 173, 528, 8, 107, 226); Group VI = 358 (55, 64, 30, 6, 139, 64); and Group VII = 110 (3,2,53,1,43,8). Supplementary Fig. 8 presents results derived with research effort into mammalian hosts and viruses included in the constituent models trained in the viral and mammalian perspectives, respectively.Full size imageMammalian host rangeOur results suggested that the average mammalian host range of our viruses is 14.33 [4.78, 54.53] (average fold increase of ~3.18 [~1.23, ~9.86] in number of hosts detected per virus). Overall, RNA viruses had the average host range of 21.65 [7.01, 82.96] hosts (~4.00- fold increase [~1.34, ~14.15]). DNA viruses, on the other hand, had 7.85 [2.81, 29.47] hosts on average (~2.43 [~1.14, ~6.89] fold increase). Table 3 lists the results of our framework at Baltimore group level and selected family and transmission routes of our viruses. Figure 2 illustrates predicted mammalian host range of our viruses (Fig. 2B, Supplementary Data 3), and the increase in predicted number of viruses per species in species-rich mammalian orders of interest (Fig. 2C, Supplementary Data 4).Table 3 Predicted range of susceptible mammalian species of viruses per Baltimore group, family (top 15 families, ranked by fold increase) and transmission route.Full size tableRelative importance of mammalian featuresWe trained a suite of models for each virus species with two or more known mammalian hosts (n = 556, response variable = 1 if the mammal is known to associate with the focal virus species, 0 otherwise). This allowed us to calculate relative importance of mammalian traits (Table 2) to our viral models. We were also able to capture variations in how these features contribute to our viral models at various levels (e.g. Baltimore classification, or transmission route) as highlighted in Fig. 3A. Our results indicated that distances to known hosts of viruses were the top predictor of associations between the focal virus and our terrestrial mammals. The breakdown was: 1) mean phylogenetic distance – all viruses = 98.75% [93.01%, 100%], DNA = 99.48% [96.03%, 100%], RNA = [91.93%, 100%]; 2) mean ecological distance all viruses = 94.39% [71.86%, 100%], DNA = 96.36% [80.99%, 100%], RNA = [69.48%, 100%]. In addition, life-history traits significantly improved our models, in particular: longevity (all viruses = 60.9% [12.12%, 98.88%], DNA = 68.03% [11.22%, 99.69%], RNA = [13.55%, 96.37%]); body mass (all viruses = 62.92% [5.4%, 97.65%], DNA = 72.75% [18.49%, 100%], RNA = 57.45% [4.32%, 95.5%]); and reproductive traits (all viruses = 53.37% [5.67%, 95.99%]%, DNA = 59.46% [8.27%, 99.32%], RNA = 50.17% [4.85%, 92.17%]).Fig. 3: Results (Mammals).Panel A Variable importance (relative contribution) of mammalian traits to viral perspective models. Variable importance is calculated for each constituent model (n = 556) of our viral perspective (trained with mammalian features), and then aggregated (median) per each reported group (columns). Panel B Number of known and new viruses associated with each mammal. Labelled mammals are as follows: top 4 (by number of new viruses) for each of Artiodactyla, Carnivora, Chiroptera, Primates, Rodentia, and other orders. Species in bold have 100 or more predicted viruses (Supplementary Data 5). Panel C Top 18 genera (by number of predicted wild or semi-domesticated mammalian host species) in selected orders (Other indicated results for all orders not included in the first five circles). Each order figure comprises the following circles (from outside to inside): 1) Number of hosts predicted to have an association with viruses within the viral genus. 2) Number of hosts detected to have association. 3) Number of hosts predicted to harbour viral zoonoses (i.e. known or predicted to share at least one virus species with humans). 4) Number of hosts predicted to share viruses with domesticated mammals of economic significance (domesticated mammals in orders: Artiodactyla, Carnivora, Lagomorpha and Perissodactyla). 5) Baltimore classification of the selected genera (Supplementary Data 6). Supplementary Fig. 9 presents results derived with research effort into mammalian hosts and viruses included in the constituent models trained in the viral and mammalian perspectives, respectively.Full size imageWild and semi-domesticated susceptible mammalian hosts of virusesour framework indicated ~4.28 -fold increase [~1.2, ~14.64] of the number of virus species in wild and or semi-domesticated mammalian hosts (16.86 [4.95, 68.5] viruses on average per mammalian species). These results indicated an average of 13.45 [1.73, 65.04] unobserved virus species for each wild or semi-domesticated mammalian host (known viruses that are yet to be associated with these mammals). Our framework highlighted differences in the number of viruses predicted per order (Table 4). Figure 3 illustrates the predicted number of viruses in wild or semi-domesticated mammal by mammalian host range (Fig. 3B, Supplementary Data 5), and the top 18 virus genera (per number of host-virus associations) in selected orders (Fig. 3C, Supplementary Data 6). Supplementary Results 1 lists the results with the inclusion of research effort into mammalian species in our viral perspective models.Table 4 Predicted number of viruses per top 15 orders by fold increase in number of viruses predicted in wild or semi-domesticated mammalian hosts (per species).Full size tableNetwork perspective – Potential motifs
    We quantified the topology of the network linking virus and mammal species by means of counts of potential motifs21. Figure 4 illustrates how potential motifs are captured in our network. Briefly, for each virus-mammal association for which we want to make predictions (n = 2,722,656, of which 6,331 are supported by our evidence, see methods section), we “force insert” this focal association into our network (Fig. 4A, B) and enumerate all instances of 3 (n = 2), 4 (n = 6), and 5-node (n = 20) potential motifs in which this association might feature if it actually existed21 (Fig. 4C visualises these different motifs). Following this process, a features-set is generated comprising the counts potential motifs for all included associations. Figure 4D illustrates the count of motifs (logged) grouped by mammalian order and virus Baltimore classification.Fig. 4: The network perspective – potential motifs (subgraphs) in our virus-host bipartite network.A The concept of potential motif. The association TBEV-P. leo is a forced insertion into the network prior to calculating motifs for the association. B Motifs space: networks represent 2 steps and 3 steps ego networks (union) of host (here P. leo) and virus (TBEV). 1, 2 and 3 step ego networks comprise the counting space for TBEV-P. leo potential motifs. Dark grey nodes represent viruses, light grey nodes represent hosts. Size of nodes is adjusted to represent overall number of hosts or viruses with known associations to the node. Red edges represent nodes reachable from the mammal (P. leo) in 1 or 2 steps (links). Blue edges represent nodes reachable from the virus (TBEV) with 1 or 2 steps (links). Humans and rabies virus were excluded from these networks. C 3, 4 and 5-node potential motifs in our virus-host bipartite network. Circles represent viruses and squares represent mammals. Red circles represent the focal virus (v), and blue squares represent the focal mammal (m) of the association v-m for which the motifs are being counted (dashed yellow line). This association has two states: either already known (documented in EID2), or unknown. Grey lines illustrate existing associations in our network. D Motifs counts. Heatmap illustrating distribution of motif-features (counts of potential motifs per each focal association) in our bipartite network, grouped by mammalian order and Baltimore classification. The counts are logged to allow for better visualisation. E Variable importance (relative contribution) of motif-features (variables) to our network perspective models (SVM-RW). Motifs (subgraphs) are coloured by the number of nodes (K = 3, 4, 5). Boxplots indicate median (centre), the 25th and 75th percentiles (bounds of box) and inter quantile range (whiskers). Points represent variable importance in individual runs (n = 100). Research effort into both viruses and mammals is included as independent variables in our network models (coloured in yellow).Full size imageRelative importance of network (motif) featuresFigure 4E illustrates that M4.1 was the most important feature in our network models: median = 100% [90.19%, 100%]. Followed by: M5.1 = 97.84% [89.19%, 99.93%], M5.7 = 98.8 97.22% [87.7%, 98.77%] and M4.6 = 96.75% [86.13%, 100%]. Research effort of viruses and mammals had relative importance = 90.26% [82.94%, 95.36%], 88.42% [78.38%, 94.87%] respectively. Overall, 5-node motif-features had median relative influence = 75.06% [1.21%, 98.14%]; whereas 3 and 4-node motif-features had relative influence = 71.69% [55.76%, 85.34%], and 61.06% [27.14%, 100%], respectively. Supplementary Fig. 29 illustrate the partial dependence of network perspective models on each of our network features.ValidationWe validated our framework in three ways: 1) against a held-out test set; 2) by systematically removing selected known viral-mammalian associations and attempting to predict them; and 3) against external data source, comprising viral-mammalian associations extracted using an exhaustive literature search targeting wild mammals and their viruses4,30.Our held-out test set comprised 15% of all data (randomly selected, n = 407,265; 954 known virus-mammal associations, see methods below). We removed this set from our network, computed network features (motifs), and trained constituent models in each perspective with the remainder data. We then estimated our framework performance metrics against the held-out test set. Our framework achieved overall AUC = 0.938 [0.862–0.959], F1-Score = 0.284 [0.464–0.124], and TSS = 0.876 [0.724–0.918], when trained without including research effort in its mammalian and viral perspectives. When research effort was included in these perspectives, performance metrics were as follows: AUC = 0.920 [0.823, 0.944], F1-Score = 0.272 [0.526, 0.093], and TSS = 0.840 [0.646, 0.888].The performance of our voting approach was better than any individual perspective, or combination of perspectives (Supplementary Tables 8–11). The most significant improvement was in F1-score, where individual perspectives scores were as follows: network = 0.104 [0.210–0.051], mammalian = 0.115 [0.009–0.064] (0.131 [0.284–0.035] with research effort), and viral = 0.181 [0.374–0.074] (0.196 [0.373–0.067]).Additionally, we conducted a systematic test to predict removed virus-mammal associations. In this test, we systematically removed one known virus-mammal association at a time from our framework, recalculated all inputs (including from network) and attempted to predict these removed associations. Our framework succeeded in predicting 90% of removed associations (90.70% for associations removed for viruses, 89.92% for associations removed from mammals, Supplementary Results 3).Finally, our framework predicted 84.02% [77.69%, 89.60%] of the externally obtained viral-mammalian associations (with detection quality  > 0) where both host and virus were included in our pipeline, and 77.82% [68.46%, 86.51%] (any detection quality). When including research effort in our mammalian and viral perspectives, these results were: 84.47% [78.15%, 89.60%], and 78.41% [68.83%, 86.37%], respectively. More