More stories

  • in

    Coastal algal blooms have intensified over the past 20 years

    RESEARCH BRIEFINGS
    01 March 2023

    Global spatial and temporal patterns of coastal phytoplankton blooms were characterized using daily satellite imaging between 2003 and 2020. These blooms were identified on the coast of 126 of the 153 ocean-bordering countries examined. The extent and frequency of blooms have increased globally over the past two decades. More

  • in

    Sub-continental-scale carbon stocks of individual trees in African drylands

    OverviewThis study establishes a framework for mapping carbon stocks at the level of individual trees at a sub-continental scale in semi-arid sub-Saharan Africa north of the Equator. We used satellite imagery from the early dry season (Extended Data Fig. 1). The deep learning method developed by a previous study1 allowed us to map billions of discrete tree crowns at the 50-cm scale from West Africa to the Red Sea. Then we used allometry to convert tree crown area into tree wood, foliage and root carbon for the 0–1,000 mm year−1 precipitation zone in which our allometry was collected (Extended Data Fig. 2). We introduce a viewer that enables the billions of trees to be viewed at different scales, with information on location, metadata of the Maxar satellite image used, tree crown area and the estimated wood, foliage and root carbon content based on our allometry (Fig. 4). We also make available our output data for the 1,000 mm year−1 precipitation zone southward to 9.5° N latitude with information on location, precipitation, metadata of the Maxar satellite image used, tree crown area, tree wood carbon, tree root carbon and tree leaf carbon.Satellite imageryWe used 326,523 Maxar multispectral images from the QuickBird-2, GeoEye-1, WorldView-2 and WorldView-3 satellites collected from 2002 to 2020 from November to March from 9.5° N to 24° N latitude within Universal Transverse Mercator (UTM) zones 28–37 for Africa (Extended Data Table 1a). These images were obtained by NASA through the NextView License from the National Geospatial-Intelligence Agency. Data were assembled over several years with a focus on later years to achieve a relatively recent and complete wall-to-wall coverage.When using satellite data from different satellites over several years, with varying sun–target–satellite angles, with varying radiometric calibration of satellite spectral bands and different atmospheric compositions through which the surface is imaged, there are two possibilities for using hundreds of thousands of satellite images together quantitatively. One approach, used extensively in NASA’s, NOAA’s and the European Space Agency’s Earth-viewing satellite programmes, is to quantitatively inter-calibrate radiometrically the satellite channels through time; correct these data for time-dependent atmospheric effects such as aerosols, clouds, haze, smoke, dust and other atmospheric constituent effects and then normalize the viewing perspective to the same sun–target–satellite angle38. Another approach is to use the satellite data as collected; assemble training data of trees viewed from different satellites under different sun–target–satellite angles, different times, different atmospheric conditions and use machine learning with high-performance computing to perform the tree mapping at the 50-cm scale. The key to successful machine learning is to account for all the sources of variation within the domain of study in the training data to ensure accurate identification of trees under all circumstances. We included trees viewed substantially off-nadir, trees collected under different aerosol optical thicknesses, trees collected under cirrus cloud conditions, trees viewed in the forward and backward scan directions, trees on sandy soils, trees on clay soils, trees on burn scars, trees in laterite areas and trees in riverine settings. Our training data were collected by one team member and are a carefully selected manual delineation of 89,899 individual trees under a range of atmospheric conditions, viewing perspectives and ecological settings.All multispectral and panchromatic bands associated with our Maxar images were orthorectified to a common mapping basis. We next pan-sharpened all multispectral bands to the 0.5-m scale with the associated panchromatic band. The absolute locational uncertainty of pixels at the 0.5-m scale from orbit is approximately ±11 m, considering the root-mean-square location errors among the QuickBird-2, GeoEye-1, WorldView-2 and WorldView-3 satellites (Extended Data Table 1). We formed the normalized difference vegetation index (NDVI)39 from every image in the traditional way from the pan-sharpened red and near-infrared bands. We also associated the panchromatic band with the NDVI band and ensured that the panchromatic and NDVI bands were highly co-registered. The NDVI was used to distinguish tree crowns from non-vegetated background because the images were taken from a period when only woody plants were photosynthetically active in this area36. Our training data were labelled on images from the early dry season when only trees have green leaves. Because most semi-arid savannah trees continue to photosynthesize in the early dry season after herbaceous vegetation senesces, green leaf tree crowns are easily mapped because of their higher NDVI values than their senescent herbaceous vegetation surroundings. We substantiate this by analysis of 308 individual trees using NDVI time series with 4-m PlanetScope imagery that emphasized the importance of satellite data from the November, December and January early dry-season months (Extended Data Fig. 1).We next formed our data into mosaics by applying a set of decision rules, resulting in a collection of 16 × 16-km tiles within each UTM zone from 9.5° N to 24° N latitude for Africa. The initial round of scoring considered percentage cloud cover, sun elevation angle and sensor off-nadir angle: preference was given to imagery that had lower cloud cover, then higher sun elevation angle and finally view angles closest to nadir. In the second round of scoring, selections were assigned priority to favour early dry-season months and off-nadir view angles: preference was given to imagery from November, December and January with off-nadir angle less than ±15°; second to imagery from November to January with off-nadir angle between ±15° and ±30°; third to imagery from February or March with off-nadir angle less than ±15°; and finally to imagery from February or March with off-nadir angle between ±15° and ±30°. Image mosaics were necessary to eliminate multiple counting of trees. We formed mosaics using 94,502 images for tree segmentation, with 94% of these being from November, December and January. Ninety percent of our selected mosaic imagery was within ±15° of nadir, 87% were acquired between 2010 and 2020 and 94% were from the early dry season (Extended Data Fig. 7). A summary of month, year, solar elevation and off-nadir angle by UTM zone can be found in Supplemental Information Fig. 1.Possible obscuration of the surface by clouds totalled 4.1% of our input mosaic data area and aerosol optical depth >0.6 at 470-nm (ref. 40) areas totalled 3.4% of our input data. However, we mapped 691,477,772 trees in our possible cloud-cover-affected and aerosol-affected areas, indicating that cloud and aerosol effects were lower than these numbers. In addition, 0.9% of our input data did not process. We include a data layer in our viewer for these three conditions.Mapping tree crowns with deep learningWe used convolutional neural network models developed by a previous study1. The models were trained with manually delineated and annotated 89,899 individual trees along a north–south gradient from 0 to 1,000 mm year−1 rainfall1. Only features that showed a distinct crown area and associated shadow were included, which excluded small bushes, grass tussocks, rocks and other features that might have green leaves or cast a shadow from our classification. All training data and model training was done in UTM zones 28 and 29. Because tree floristic diversity in the 0–1,000 mm year−1 zone of our study is highly similar from the Atlantic Ocean to the Red Sea across Africa41,42,43, we added no further training data as our study moved further eastward. We used state-of-the-art deep learning to segment trees crowns at the 50-cm scale1. We used two different models based on a U-Net architecture, one for lower-rainfall desert regions with 150 mm year−1. Details about the network architecture, training process and hyperparameter choices can be found in ref. 1. Previous evaluation showed that early dry-season images performed better than late dry-season images, which was a limitation of our previous study. We reduced this error by using early dry-season images with only 6% of our area being covered by images from February and March. The models were also designed to separate clumped trees by highlighting spaces between different crowns during the learning process, similar to a strategy for separating touching cells in microscopic imagery22.AllometryVery-high-resolution satellite images and deep learning have achieved mapping of individual trees over large areas1. Each tree is georeferenced in the satellite data and defined by crown area. The challenge was to develop allometric equations for foliage, wood and root dry masses or carbon based on crown area regardless of species. This was met by reanalysing existing Sahelian and Sudanian woody plant data from destructive sampling. Overall, the seasonal maximum foliage, wood and root dry masses were measured on 900, 698 and 26 trees or shrubs from 27, 26 and 5 species, respectively, for which crown area was also measured. Several allometric regression models tested for foliage, wood or root masses are power functions and independent of species. All the regression outputs were inter-compared for fit indicators, by systematic estimates of prediction uncertainty and by root-to-wood ratios and foliage-to-wood ratios over the range of crown areas. This resulted in a set of ordinary least squares log–log equations with crown area as the independent variable. The Sahelian and Sudanian allometry equations were also compared with published allometry equations for tropical trees, primarily from more humid tropics, which are generally based on stem diameter, tree height and wood density. Our allometric predictions are within the range of other allometry predictions, reinforcing the confidence in their use beyond the Sahelian and Sudanian domains into sub-humid savannahs for discrete trees19.On the basis of ref. 19, we predicted the wood (w), foliage (f) and root (r) dry mass as functions of the crown area (A) of a single tree as:$$begin{array}{c}{text{mass}}_{{rm{w}}}(A)=3.9448times {A}^{1.1068},({N}_{{rm{w}}}=698)\ {text{mass}}_{{rm{f}}}(A)=0.2693times {A}^{0.9441},({N}_{{rm{f}}}=900)\ {text{mass}}_{{rm{r}}}(A)=0.8339times {A}^{1.1730},({N}_{{rm{r}}}=26)end{array}$$The tree mass components of wood, leaves and roots were combined to predict the total mass(A) in kg of a tree from its crown area A in m2:$$text{mass}left(Aright)={text{mass}}_{{rm{w}}}left(Aright)+{text{mass}}_{{rm{f}}}left(Aright)+{text{mass}}_{{rm{r}}}left(Aright)$$As in ref. 1, a crown area of size A  > 200 m2 was split into ({rm{lfloor }}A/100{rm{rfloor }}) areas of size 100 m2 and one area with the remaining m2 if necessary. We converted dry mass to carbon by multiplying with a factor of 0.47 (ref. 44).Uncertainty analysisWe evaluated the uncertainty of our tree crown area mapping and carbon estimation in two ways. First, we quantified our tree crown mapping omission and commission errors by inspecting randomly selected areas from UTM zones 28–37, validating that our neural network generalized over UTM zones consistently (Extended Data Fig. 8).Second, we quantified the relative error of our tree crown area estimation. We consider the uncertainty Δx of a quantity x and the corresponding relative uncertainty δx defined by the absolute and relative error, respectively45. To assess the relative error in crown area estimation resulting from errors by the neural network, we considered external validation data from ref. 1, which were not used in the model-building process. We considered expert-labelled tree crowns as well as the predicted tree crowns from 78 plots of 256 × 256 pixels. The hand-labelled set contained 5,925 trees and the system delineated 5,915 trees. The total hand-labelled tree crown area was 118,327 m2 and the neural network predicted 121,898 m2. This gave a relative error in crown area mapping of δarea = 3.3%. We matched expert-labelled and predicted tree crowns and computed the root-mean-square error (RMSE) per tree, taking overlapping areas and missed trees into account (see Extended Data Fig. 8). We estimated the allometric uncertainty (δallometric) using the data from ref. 19 (see below). The two relative errors δarea and δallometric were combined to an overall uncertainty estimate for the carbon prediction of ±19.8% (see below).Omission and commission errorsWe evaluated our tree crown mapping accuracy by analysis of 1,028 randomly selected 512 × 256-pixel areas over the 9.5° N to 24° N latitude within UTM zones 28–37. Because the drier 60% of our study area only contains 1% of the 9,947,310,221 trees we mapped in the 0–1,000 mm year−1 rainfall zone, we applied an 80% bias for selecting evaluation areas above the 200 mm year−1 precipitation line46, as >98% of tree identifications were above the 200 mm year−1 precipitation isoline. Identified tree polygons were further categorized into tree crown area classes from 0–15 m2, 15–50 m2, 50–200 m2 and >200 m2, with a total of 50,570 trees evaluated. Although a previous study reported greatest uncertainty in both the smallest and largest area classes1, our more expansive work found the greatest uncertainty in our smallest tree class. We excluded from evaluation any tiles that had annual precipitation46 >1,000 mm year−1 and all areas that were devoid of vegetation, leaving us with 850 areas.Seven members of our team evaluated the accuracy in terms of commission and omission by tree crown area classes for the 850 areas. Input data provided for every area were the NDVI layer, the panchromatic shadow layer and the neural net mapping results in each of the four crown area classes. Ancillary data available to evaluators included the centre coordinates for comparison with Google Earth data, the Funk et al.46 rainfall, the acquisition date of the area evaluated and the viewing perspective.We identified areas wrongly classified as tree crowns (commission errors), missed trees (omission errors) and crown areas corresponding to clumped trees (Extended Data Fig. 8). Clumped trees were most common for >200 m2 tree crown area. They were rare in the 3–15 m2 and 15–50 m2 tree classes, which comprise 88% of our tree crowns. In the 850 patches, the number of trees ranged from one tree to 326 trees, with a total of 50,570 trees evaluated and 3,765 errors identified. Overall, the commission and omission error rates were 4.9% and 2.7%, respectively, a net uncertainty of 2.2%.Allometric uncertainty estimationThe prediction of tree carbon from the crown area for a single tree based on crown area alone is inherently uncertain47,48. As the allometric equations are based on three different datasets, we compute their uncertainties independently, combine them and put them in relation to the total carbon measured in the three datasets.The allometric equations were established using an optimal least-squares fit of an affine linear model predicting the logarithmic carbon from the logarithmic tree crown area19. To estimate the uncertainty of the allometric equations, we repeated the fitting using random subsampling. The datasets were randomly split into training data (80%) for fitting the allometric equations and validation data (20%) for assessing the uncertainty. For example, from the root measurements, (({A}_{1},{y}_{1}),ldots ,({A}_{{N}_{{rm{r}}}},,{y}_{{N}_{{rm{r}}}})), we compute ({mu }_{{rm{r}}}=frac{1}{{N}_{{rm{r}}}}mathop{sum }limits_{i=1}^{{N}_{{rm{r}}}}{y}_{i}) and ({hat{mu }}_{{rm{r}}}=frac{1}{{N}_{{rm{r}}}}mathop{sum }limits_{i=1}^{{N}_{{rm{r}}}}{text{mass}}_{{rm{r}}}({A}_{i})). The corresponding error is ({varDelta }_{{rm{r}}}=|{mu }_{{rm{r}}}-{hat{mu }}_{{rm{r}}}|).Because the total carbon for a tree with a certain crown area is the sum of the three carbon components, we add the absolute uncertainties assuming independence45.$${varDelta }_{{rm{a}}{rm{l}}{rm{l}}{rm{o}}{rm{m}}{rm{e}}{rm{t}}{rm{r}}{rm{i}}{rm{c}}}simeq sqrt{{varDelta }_{{rm{f}}}^{2}+{varDelta }_{{rm{w}}}^{2}+{varDelta }_{{rm{r}}}^{2}}$$and compute the relative uncertainty as ({delta }_{text{allometric}}=frac{{varDelta }_{text{allometric}}}{{mu }_{text{mass}}}), in which the average mass μmass is given by the sum of the averages for wood (μw), leaves (μf) and root (μr). This process was repeated ten times, resulting in a mean relative uncertainty of$${bar{delta }}_{{rm{allometric}}}=19.5 % .$$Total carbon uncertaintyWe combine the uncertainties from the neural net mapping and our allometric equations, which can be viewed as considering (1 + A)·(1 + B) with A and B being random variables with standard deviations δarea and δallometric. Neglecting higher-order and interaction terms, we combine the two sources of uncertainty to (delta simeq sqrt{{delta }_{{rm{area}}}^{2}+{bar{delta }}_{{rm{allometric}}}^{2}}), resulting in an uncertainty in total tree carbon for our study of ±19.8%. See also Extended Data Fig. 9 for the RMSEs of our predicted crown areas calculated on external validation data from ref. 1, binned on the basis of the 50th quantiles of the hand-labelled crown areas and converted also into carbon. Extended Data Fig. 10 is a flow diagram summarizing our methods.Our viewerVisualizing our large tree-mapping dataset in an interactive format was essential for quality-control purposes, exploration of the data and hypothesis creation. Creating a web-based viewer serves the purpose of being the initial point of interaction with our dataset for fellow researchers, local stakeholders or the general public. The visualization of more than 10 billion trees in a web browser required maintaining performance, interactivity and individual metadata for each polygon. Users should be able to zoom in to any area within the dataset to view individual tree polygons and query their statistics while at the same time accurately depicting the overall trends of the dataset at lower zoom levels. The visualization also needed to clearly denote where data were missing or possibly affected by clouds or aerosols. Finally, the extent and origin of the source imagery, its acquisition date and a preview of the imagery needed to be available. To accomplish these goals, a vector-tile-based approach was taken, with the data visualized in a Mapbox GL JS map within a React web application. To create vector tiles covering the entire study area, we developed a data-processing pipeline using high-performance computing resources to transform the data into compatible formats, as well as to package, optimize and combine the vector tiles themselves.We used two tracks to store and visualize the results of this study on the web: vector polygon data and generalized rasters representing tree crown density. At the native spatial resolution of 50 cm, the map shows the full-resolution tree polygon dataset. At lower-spatial-resolution zoom levels, rasterized representations of tree density are shown. Visualizing generalized rasters in place of vector polygons improves performance substantially. As users zoom in to higher spatial resolutions, the raster layer fades away and is replaced by the full-resolution polygon layer. Once zoomed far enough to resolve individual polygons, users can click to select a polygon to show a map overlay containing various properties of the tree, as well as the date on which the source imagery was acquired and a link to preview the source imagery.Rainfall dataWe used the rainfall data of Funk et al. to estimate annual rainfall at 5.6-m grids46. We averaged the available data from 1982 to 2017 and extracted the mean annual rainfall for each mapped tree and bilinearly interpolated it to 100 × 100-m resolution. The rainfall data were also used to classify the study area into mean annual precipitation zones: hyper-arid from 0–150 mm year−1, arid from 150–300 mm year−1, semi-arid from 300–600 mm year−1 and sub-humid from 600–1,000 mm year−1 zones. The rainfall data are found at https://data.chc.ucsb.edu/products/CHIRPS-2.0/africa_monthly/ (ref. 46). More

  • in

    Open-source software for geospatial analysis

    Satellite imagery provides insight into where and how Earth’s surface changes, particularly in remote areas where in situ measurements are generally lacking. With the large volumes of data produced by satellites, we need streamlined computational pipelines for optimized processing capabilities. Although a multitude of platforms exists to process satellite data, these often have expensive license requirements that price out much of the geospatial community. Moreover, many of these platforms are propriety, but transparency is key when developing geospatial processing workflows. Open-source programming is critical to the creation of efficient imagery processing pipelines. More

  • in

    Carbon stocks of billions of individual African dryland trees estimated

    Tucker, C. et al. Nature 615, 80–86 (2023).Article 

    Google Scholar 
    Bayala, J. et al. Agric. Ecosyst. Environ. 205, 25–35 (2015).Article 

    Google Scholar 
    Keesstra, S. D. et al. Soil 2, 111–128 (2016).Article 

    Google Scholar 
    Dewi, S. et al. Int. J. Biodivers. Sci. Ecosyst. Serv. Mgmt 13, 312–329 (2017).Article 

    Google Scholar 
    Ahlström, A. et al. Science 348, 895–899 (2015).Article 
    PubMed 

    Google Scholar 
    Poulter, B. et al. Nature 509, 600–603 (2014).Article 
    PubMed 

    Google Scholar 
    Prăvălie, R. et al. Environ. Res. 201, 111580 (2021).Article 
    PubMed 

    Google Scholar 
    Reij, C. P. & Smaling, E. M. A. Land Use Policy 25, 410–420 (2008).Article 

    Google Scholar 
    Zomer, R. J., Bossio, D. A., Trabucco, A., van Noordwijk, M. & Xu, J. Circ. Agric. Syst. 2, 3 (2022).Article 

    Google Scholar 
    Chomba, S., Sinclair, F., Savadogo, P., Bourne, M. & Lohbeck, M. Front. For. Glob. Change 3, 571679 (2020).Article 

    Google Scholar 
    Dakpogan, A., Bayala, J., Ouattara, I. & Ellington, E. in United for Lands: From National Coalitions to a Pipeline of Bankable Projects for the Great Green Wall 54–56 (United Nations, 2022).
    Google Scholar 
    Garrity, D. P. & Bayala, J. in Sustainable Development Through Trees on Farms: Agroforestry in its Fifth Decade (ed. van Noordwijk, M.) 153–175 (World Agroforestry, 2019).
    Google Scholar 
    Schnell, S., Kleinn, C. & Ståhl, G. Environ. Monit. Assess. 187, 600 (2015).Article 
    PubMed 

    Google Scholar  More

  • in

    Individual personality predicts social network assemblages in a colonial bird

    Réale, D. et al. Personality and the emergence of the pace-of-life syndrome concept at the population level. Philos. Trans. R. Soc. Lond. B 365, 4051–4063 (2010).Article 

    Google Scholar 
    Gosling, S. D. From mice to men: What can we learn about personality from animal research?. Psychol. Bull. 127, 45 (2001).Article 
    CAS 

    Google Scholar 
    Dingemanse, N. J., Class, B. & Holtmann, B. Nonrandom mating for behavior in the wild?. Trends Ecol. Evol. 36, 177–179 (2021).Article 

    Google Scholar 
    Croft, D. P. et al. Behavioural trait assortment in a social network: Patterns and implications. Behav. Ecol. Sociobiol. 63, 1495–1503 (2009).Article 

    Google Scholar 
    Morton, F. B., Weiss, A., Buchanan-Smith, H. M. & Lee, P. C. Capuchin monkeys with similar personalities have higher-quality relationships independent of age, sex, kinship and rank. Anim. Behav. 105, 163–171 (2015).Article 

    Google Scholar 
    Su, X. et al. Agonistic behaviour and energy metabolism of bold and shy swimming crabs Portunus trituberculatus. J. Exp. Biol. https://doi.org/10.1242/jeb.188706 (2019).Article 

    Google Scholar 
    Jolles, J. W., King, A. J. & Killen, S. S. The role of individual heterogeneity in collective animal behaviour. Trends Ecol. Evol. 35, 278–291 (2020).Article 

    Google Scholar 
    Bell, A. M., Hankison, S. J. & Laskowski, K. L. The repeatability of behaviour: A meta-analysis. Anim. Behav. 77, 771–783 (2009).Article 

    Google Scholar 
    Frost, A. J., Winrow-Giffen, A., Ashley, P. J. & Sneddon, L. U. Plasticity in animal personality traits: Does prior experience alter the degree of boldness?. P. Roy. Soc. B-Biol. Sci. 274, 333–339 (2007).
    Google Scholar 
    Krause, J., James, R. & Croft, D. P. Personality in the context of social networks. Philos. Trans. R. Soc. Lond. B 365, 4099 (2010).Article 
    CAS 

    Google Scholar 
    David, M., Auclair, Y. & Cézilly, F. Personality predicts social dominance in female zebra finches, Taeniopygia guttata, in a feeding context. Anim. Behav. 81, 219–224 (2011).Article 

    Google Scholar 
    Favati, A., Leimar, O. & Løvlie, H. Personality predicts social dominance in male domestic fowl. PLoS ONE 9, e103535 (2014).Article 
    ADS 

    Google Scholar 
    McGhee, K. E. & Travis, J. Repeatable behavioural type and stable dominance rank in the Bluefin killifish. Anim. Behav. 79, 497–507 (2010).Article 

    Google Scholar 
    Krause, J., Croft, D. P. & James, R. Social network theory in the behavioural sciences: Potential applications. Behav. Ecol. Sociobiol. 62, 15–27 (2007).Article 
    CAS 

    Google Scholar 
    Flack, J. C., Girvan, M., de Waal, F. & Krakauer, D. C. Policing stabilizes construction of social niches in primates. Nature 439, 426–429 (2006).Article 
    ADS 
    CAS 

    Google Scholar 
    Croft, D. P., James, R. & Krause, J. Exploring Animal Social Networks (Princeton University Press, 2008).Book 

    Google Scholar 
    Patriquin, K. J., Leonard, M. L., Broders, H. G. & Garroway, C. J. Do social networks of female northern long-eared bats vary with reproductive period and age?. Behav. Ecol. Sociobiol. 64, 899–913 (2010).Article 

    Google Scholar 
    Gomes, A. C. R., Beltrão, P., Boogert, N. J. & Cardoso, G. C. Familiarity, dominance, sex and season shape common waxbill social networks. Behav. Ecol. 33, 526–540 (2022).Article 

    Google Scholar 
    Croft, D. P., Krause, J. & James, R. Social networks in the guppy (Poecilia reticulata). P. Roy. Soc. B-Biol. Sci. 271, S516–S519 (2004).Article 

    Google Scholar 
    Pike, T. W., Samanta, M., Lindström, J. & Royle, N. J. Behavioural phenotype affects social interactions in an animal network. P. Roy. Soc. B-Biol. Sci. 275, 2515–2520 (2008).
    Google Scholar 
    Aplin, L. M. et al. Individual personalities predict social behaviour in wild networks of great tits (Parus major). Ecol. Lett. 16, 1365–1372 (2013).Article 
    CAS 

    Google Scholar 
    Massen, J. J. & Koski, S. E. Chimps of a feather sit together: Chimpanzee friendships are based on homophily in personality. Evol. Hum. Behav. 35, 1–8 (2014).Article 

    Google Scholar 
    Rault, J.-L. Friends with benefits: Social support and its relevance for farm animal welfare. Appl. Anim. Behav. Sci. 136, 1–14 (2012).Article 

    Google Scholar 
    Schneider, G. & Krueger, K. Third-party interventions keep social partners from exchanging affiliative interactions with others. Anim. Behav. 83, 377–387 (2012).Article 

    Google Scholar 
    Fraser, O. N. & Bugnyar, T. Do ravens show consolation? Responses to distressed others. PLoS ONE 5, e10605 (2010).Article 
    ADS 

    Google Scholar 
    Rose, P. & Croft, D. The potential of social network analysis as a tool for the management of zoo animals. Anim. Welf. 24, 123–138 (2015).Article 

    Google Scholar 
    Clark, F. E. Space to choose: network analysis of social preferences in a captive chimpanzee community, and implications for management. Am. J. Primatol. 73, 748–757 (2011).Article 

    Google Scholar 
    Corner, L., Pfeiffer, D. & Morris, R. Social-network analysis of Mycobacterium bovis transmission among captive brushtail possums (Trichosurus vulpecula). Prev. Vet. Med. 59, 147–167 (2003).Article 
    CAS 

    Google Scholar 
    Hansen, H., McDonald, D. B., Groves, P., Maier, J. A. & Ben-David, M. Social networks and the formation and maintenance of river otter groups. Ethology 115, 384–396 (2009).Article 

    Google Scholar 
    Radosevich, L. M., Jaffe, K. E. & Minier, D. E. The utility of social network analysis for informing zoo management: Changing network dynamics of a group of captive hamadryas baboons (Papio hamadryas) following an introduction of two young males. Zoo Biol. 40, 503–516 (2021).Article 

    Google Scholar 
    Pacheco, X. P. & Madden, J. R. Does the social network structure of wild animal populations differ from that of animals in captivity?. Behav. Processes 190, 104446 (2021).Article 

    Google Scholar 
    Watters, J. V. & Powell, D. M. Measuring animal personality for use in population management in zoos: Suggested methods and rationale. Zoo Biol. 31, 1–12 (2012).Article 

    Google Scholar 
    Koski, S. E. Social personality traits in chimpanzees: temporal stability and structure of behaviourally assessed personality traits in three captive populations. Behav. Ecol. Sociobiol. 65, 2161–2174 (2011).Article 

    Google Scholar 
    Račevska, E. & Hill, C. M. Personality and social dynamics of zoo-housed western lowland gorillas (Gorilla gorilla gorilla). J. Zoo Aqua. Res. 5, 116–122 (2017).
    Google Scholar 
    Stoinski, T. S., Jaicks, H. F. & Drayton, L. A. Visitor effects on the behavior of captive western lowland gorillas: The importance of individual differences in examining welfare. Zoo Biol. 31, 586–599 (2012).Article 

    Google Scholar 
    Wielebnowski, N. C. Behavioral differences as predictors of breeding status in captive cheetahs. Zoo Biol. 18, 335–349 (1999).Article 

    Google Scholar 
    Barrett, L. P. et al. Personality assessment of headstart Texas horned lizards (Phrynosoma cornutum) in human care prior to release. Appl. Anim. Behav. Sci. 254, 105690 (2022).Article 

    Google Scholar 
    Rose, P. E., Brereton, J. E. & Croft, D. P. Measuring welfare in captive flamingos: Activity patterns and exhibit usage in zoo-housed birds. Appl. Anim. Behav. Sci. 205, 115–125 (2018).Article 

    Google Scholar 
    Rose, P. E. & Croft, D. P. Social bonds in a flock bird: Species differences and seasonality in social structure in captive flamingo flocks over a 12-month period. Appl. Anim. Behav. Sci. 193, 87–97 (2017).Article 

    Google Scholar 
    Rose, P. E. & Croft, D. P. Quantifying the social structure of a large captive flock of greater flamingos (Phoenicopterus roseus): Potential implications for management in captivity. Behav. Processes 150, 66–74 (2018).Article 

    Google Scholar 
    Rose, P. E., Croft, D. P. & Lee, R. A review of captive flamingo (Phoenicopteridae) welfare: A synthesis of current knowledge and future directions. Intern. Zoo Yearb. 48, 139–155 (2014).Article 

    Google Scholar 
    Rose, P. E. & Croft, D. P. Evaluating the social networks of four flocks of captive flamingos over a five-year period: Temporal, environmental, group and health influences on assortment. Behav. Processes 175, 104118 (2020).Article 

    Google Scholar 
    Munson, A. A., Jones, C., Schraft, H. & Sih, A. You’re just my type: Mate choice and behavioral types. Trends Ecol. Evol. 35, 823–833 (2020).Article 

    Google Scholar 
    Schuett, W., Tregenza, T. & Dall, S. R. Sexual selection and animal personality. Biol. Rev. 85, 217–246 (2010).Article 

    Google Scholar 
    Jackson, W. M. Why do winners keep winning?. Behav. Ecol. Sociobiol. 28, 271–276 (1991).Article 

    Google Scholar 
    Dammhahn, M. & Almeling, L. Is risk taking during foraging a personality trait? A field test for cross-context consistency in boldness. Anim. Behav. 84, 1131–1139 (2012).Article 

    Google Scholar 
    Van Oers, K., Drent, P. J., De Goede, P. & Van Noordwijk, A. J. Realized heritability and repeatability of risk-taking behaviour in relation to avian personalities. P. Roy. Soc. B-Biol. Sci. 271, 65–73 (2004).Article 

    Google Scholar 
    Hinton, M. G. et al. Patterns of aggression among captive American flamingos (Phoenicopterus ruber). Zoo Biol. 32, 445–453 (2013).Article 

    Google Scholar 
    Royer, E. A. & Anderson, M. J. Evidence of a dominance hierarchy in captive Caribbean flamingos and its relation to pair bonding and physiological measures of health. Behav. Processes 105, 60–70 (2014).Article 

    Google Scholar 
    Carere, C., Drent, P. J., Privitera, L., Koolhaas, J. M. & Groothuis, T. G. Personalities in great tits, Parus major: Stability and consistency. Anim. Behav. 70, 795–805 (2005).Article 

    Google Scholar 
    Jouventin, P., Lequette, B. & Dobson, F. S. Age-related mate choice in the wandering albatross. Anim. Behav. 57, 1099–1106 (1999).Article 
    CAS 

    Google Scholar 
    Black, J. M. Partnerships in Birds: The Study of Monogamy (Oxford University Press, USA, 1996).
    Google Scholar 
    Estevez, I., Andersen, I.-L. & Nævdal, E. Group size, density and social dynamics in farm animals. Appl. Anim. Behav. Sci. 103, 185–204 (2007).Article 

    Google Scholar 
    Pickering, S. The comparative breeding biology of flamingos Phoenicopteridae at the Wildfowl and Wetlands Trust Centre, Slimbridge. Intern. Zoo Yearbook 31, 139–146 (1992).Article 

    Google Scholar 
    Whitehead, H. Analyzing Animal Societies: Quantitative Methods for Vertebrate Social Analysis (University of Chicago Press, 2008).Book 

    Google Scholar 
    Wilson, A. D., Krause, S., Dingemanse, N. J. & Krause, J. Network position: A key component in the characterization of social personality types. Behav. Ecol. Sociobiol. 67, 163–173 (2013).Article 

    Google Scholar 
    Renner, M. J. & Kelly, A. L. Behavioral decisions for managing social distance and aggression in captive polar bears (Ursus maritimus). J. Appl. Anim. Welf. Sci. 9, 233–239 (2006).Article 
    CAS 

    Google Scholar 
    Stevens, E. F. & Pickett, C. Managing the social environments of flamingos for reproductive success. Zoo Biol. 13, 501–507 (1994).Article 

    Google Scholar 
    Franks, D. W., Ruxton, G. D. & James, R. Sampling animal association networks with the gambit of the group. Behav. Ecol. Sociobiol. 64, 493–503 (2010).Article 

    Google Scholar 
    Haddadi, H. et al. Determining association networks in social animals: Choosing spatial–temporal criteria and sampling rates. Behav. Ecol. Sociobiol. 65, 1659–1668 (2011).Article 

    Google Scholar 
    Whitehead, H. & Dufault, S. Techniques for analyzing vertebrate social structure using identified individuals. Adv. Stud. Behav. 28, 33–74 (1999).Article 

    Google Scholar 
    Borgatti, S.P., M., E., G., & C., F.L. UCINET for windows: software for social network analysis. Analytic Technologies: Harvard, MA (2002).Borgatti, S. P. NetDraw: graph visualization software (Analytic Technologies, 2002).
    Google Scholar 
    Bejder, L., Fletcher, D. & Bräger, S. A method for testing association patterns of social animals. Anim. Behav. 56, 719–725 (1998).Article 
    CAS 

    Google Scholar 
    Farine, D. R. & Whitehead, H. Constructing, conducting and interpreting animal social network analysis. J. Anim. Ecol. 84, 1144–1163 (2015).Article 

    Google Scholar 
    Perdue, B. M., Gaalema, D. E., Martin, A. L., Dampier, S. M. & Maple, T. L. Factors affecting aggression in a captive flock of Chilean flamingos (Phoenicopterus chilensis). Zoo Biol. 30, 59–64 (2011).
    Google Scholar 
    IBMCorp. IBM SPSS Statistics for Windows. IBM Corp: Armonk, NY (2012).Clarke, K.R. & Gorley, R.N. PRIMER v6: User Manual/Tutorial. PRIMER-E, Plymouth. (2006).Kassambara, A. & Mundt, F. factoextra: Extract and Visualize the Results of Multivariate Data Analyses. (2020).RCoreTeam. R: A language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing. (2021).Budaev, S. V. Using principal components and factor analysis in animal behaviour research: Caveats and guidelines. Ethology 116, 472–480 (2010).Article 

    Google Scholar 
    Whitehead, H. SOCPROG programs: Analysing animal social structures. Behav. Ecol. Sociobiol. 63, 765–778 (2009).Article 

    Google Scholar 
    Whitehead, H. SOCPROG: Programs for analyzing social structure: Whitehead Lab (2019).Hanneman, R.A. & Riddle, M., Chapter 18: Some Statistical Tools. In: Introduction to Social Network Methods. (University of California, Riverside 2005). http://faculty.ucr.edu/~hanneman/.(2005) More

  • in

    Phototrophy by antenna-containing rhodopsin pumps in aquatic environments

    Balashov, S. P. et al. Xanthorhodopsin: a proton pump with a light-harvesting carotenoid antenna. Science 309, 2061–2064 (2005).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Imasheva, E. S., Balashov, S. P., Choi, A. R., Jung, K.-H. & Lanyi, J. K. Reconstitution of Gloeobacter violaceus rhodopsin with a light-harvesting carotenoid antenna. Biochemistry 48, 10948–10955 (2009).Article 
    CAS 
    PubMed 

    Google Scholar 
    Fuhrman, J. A., Schwalbach, M. S. & Stingl, U. Proteorhodopsins: an array of physiological roles? Nat. Rev. Microbiol. 6, 488–494 (2008).Article 
    CAS 
    PubMed 

    Google Scholar 
    Vollmers, J. et al. Poles apart: Arctic and Antarctic Octadecabacter strains share high genome plasticity and a new type of xanthorhodopsin. PLoS ONE 8, e63422 (2013).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bertsova, Y. V., Arutyunyan, A. M. & Bogachev, A. V. Na+-translocating rhodopsin from Dokdonia sp. PRO95 does not contain carotenoid antenna. Biochem. Mosc. 81, 414–419 (2016).Article 
    CAS 

    Google Scholar 
    Misra, R., Eliash, T., Sudo, Y. & Sheves, M. Retinal–salinixanthin interactions in a thermophilic rhodopsin. J. Phys. Chem. B 123, 10–20 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Béjà, O. et al. Bacterial rhodopsin: evidence for a new type of phototrophy in the sea. Science 289, 1902–1906 (2000).Article 
    ADS 
    PubMed 

    Google Scholar 
    Béjà, O., Spudich, E. N., Spudich, J. L., Leclerc, M. & DeLong, E. F. Proteorhodopsin phototrophy in the ocean. Nature 411, 786–789 (2001).Article 
    ADS 
    PubMed 

    Google Scholar 
    Atamna-Ismaeel, N. et al. Widespread distribution of proteorhodopsins in freshwater and brackish ecosystems. ISME J. 2, 656–662 (2008).Article 
    CAS 
    PubMed 

    Google Scholar 
    Frigaard, N.-U., Martinez, A., Mincer, T. J. & DeLong, E. F. Proteorhodopsin lateral gene transfer between marine planktonic Bacteria and Archaea. Nature 439, 847–850 (2006).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Finkel, O. M., Béjà, O. & Belkin, S. Global abundance of microbial rhodopsins. ISME J. 7, 448–451 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Gómez-Consarnau, L. et al. Microbial rhodopsins are major contributors to the solar energy captured in the sea. Sci. Adv. 5, eaaw8855 (2019).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    DeLong, E. F. & Béjà, O. The light-driven proton pump proteorhodopsin enhances bacterial survival during tough times. PLoS Biol. 8, e1000359 (2010).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Munson-McGee, J. H. et al. Decoupling of respiration rates and abundance in marine prokaryoplankton. Nature 612, 764–770 (2022).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wang, W.-W., Sineshchekov, O. A., Spudich, E. N. & Spudich, J. L. Spectroscopic and photochemical characterization of a deep ocean proteorhodopsin. J. Biol. Chem. 278, 33985–33991 (2003).Article 
    CAS 
    PubMed 

    Google Scholar 
    Man, D. Diversification and spectral tuning in marine proteorhodopsins. EMBO J. 22, 1725–1731 (2003).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lanyi, J. K. & Balashov, S. P. in Halophiles and Hypersaline Environments (eds. Ventosa, A., Oren, A. & Ma, Y.) 319–340 (Springer, 2011).Balashov, S. P. et al. Reconstitution of Gloeobacter rhodopsin with echinenone: role of the 4-keto group. Biochemistry 49, 9792–9799 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kopejtka, K. et al. A bacterium from a mountain lake harvests light using both proton-pumping xanthorhodopsins and bacteriochlorophyll-based photosystems. Proc. Natl Acad. Sci. USA 119, e2211018119 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Pushkarev, A. & Béjà, O. Functional metagenomic screen reveals new and diverse microbial rhodopsins. ISME J. 10, 2331–2335 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pushkarev, A. et al. A distinct abundant group of microbial rhodopsins discovered using functional metagenomics. Nature 558, 595–599 (2018).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Chazan, A. et al. Diverse heliorhodopsins detected via functional metagenomics in freshwater Actinobacteria, Chloroflexi and Archaea. Environ. Microbiol. 24, 110–121 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Inoue, K. et al. A light-driven sodium ion pump in marine bacteria. Nat. Commun. 4, 1678 (2013).Article 
    ADS 
    PubMed 

    Google Scholar 
    Bhosale, P. & Bernstein, P. S. Microbial xanthophylls. Appl. Microbiol. Biotechnol. 68, 445–455 (2005).Article 
    CAS 
    PubMed 

    Google Scholar 
    Demmig-Adams, B., Polutchko, S. K. & Adams, W. W. Structure–function–environment relationship of the isomers zeaxanthin and lutein. Photochem 2, 308–325 (2022).Article 

    Google Scholar 
    Barreiro C. & Barredo J. L. Microbial Carotenoids: Methods and Protocols (Humana Press, 2018).Ram, S., Mitra, M., Shah, F., Tirkey, S. R. & Mishra, S. Bacteria as an alternate biofactory for carotenoid production: a review of its applications, opportunities and challenges. J. Funct. Foods 67, 103867 (2020).Article 
    CAS 

    Google Scholar 
    Shibata, M. et al. Oligomeric states of microbial rhodopsins determined by high-speed atomic force microscopy and circular dichroic spectroscopy. Sci. Rep. 8, 8262 (2018).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Luecke, H. et al. Crystallographic structure of xanthorhodopsin, the light-driven proton pump with a dual chromophore. Proc. Natl Acad. Sci. USA 105, 16561–16565 (2008).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chuon, K. et al. Assembly of natively synthesized dual chromophores into functional actinorhodopsin. Front. Microbiol. 12, 652328 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yoshizawa, S., Kawanabe, A., Ito, H., Kandori, H. & Kogure, K. Diversity and functional analysis of proteorhodopsin in marine Flavobacteria. Environ. Microbiol. 14, 1240–1248 (2012).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ahmed, F. et al. Profiling of carotenoids and antioxidant capacity of microalgae from subtropical coastal and brackish waters. Food Chem. 165, 300–306 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Shihoya, W. et al. Crystal structure of heliorhodopsin. Nature 574, 132–136 (2019).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Kishi, K. E. et al. Structural basis for channel conduction in the pump-like channelrhodopsin ChRmine. Cell 185, 672–689.e23 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Balashov, S. P., Imasheva, E. S., Wang, J. M. & Lanyi, J. K. Excitation energy-transfer and the relative orientation of retinal and carotenoid in xanthorhodopsin. Biophys. J. 95, 2402–2414 (2008).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lakowicz, J. R. (ed.) in Principles of Fluorescence Spectroscopy 27–61 (Springer, 2006).Dana, J. et al. Testing the fate of nascent holes in CdSe nanocrystals with sub-10 fs pump–probe spectroscopy. Nanoscale 13, 1982–1987 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Polívka, T. et al. Femtosecond carotenoid to retinal energy transfer in xanthorhodopsin. Biophys. J. 96, 2268–2277 (2009).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Iyer, E. S. S., Gdor, I., Eliash, T., Sheves, M. & Ruhman, S. Efficient femtosecond energy transfer from carotenoid to retinal in Gloeobacter rhodopsin–salinixanthin complex. J. Phys. Chem. B 119, 2345–2349 (2015).Article 
    CAS 
    PubMed 

    Google Scholar 
    Doi, S., Tsukamoto, T., Yoshizawa, S. & Sudo, Y. An inhibitory role of Arg-84 in anion channelrhodopsin-2 expressed in Escherichia coli. Sci. Rep. 7, 41879 (2017).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Nagiri, C. et al. Crystal structure of human endothelin ETB receptor in complex with peptide inverse agonist IRL2500. Commun. Biol. 2, 236 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yamashita, K., Hirata, K. & Yamamoto, M. KAMO: towards automated data processing for microcrystals. Acta Crystallogr. D Struct. Biol. 74, 441–449 (2018).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kabsch, W. XDS. Acta Crystallogr. D Biol. Crystallogr. 66, 125–132 (2010).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Afonine, P. V. et al. Towards automated crystallographic structure refinement with phenix.refine.Acta Crystallogr. D Biol. Crystallogr. 68, 352–367 (2012).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zivanov, J. et al. New tools for automated high-resolution cryo-EM structure determination in RELION-3.eLife 7, e42166 (2018).Punjani, A., Rubinstein, J. L., Fleet, D. J. & Brubaker, M. A. cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290–296 (2017).Punjani, A., Zhang, H. & Fleet, D. J. Non-uniform refinement: adaptive regularization improves single-particle cryo-EM reconstruction. Nat. Methods 17, 1214–1221 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Rosenthal, P. B. & Henderson, R. Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy. J. Mol. Biol. 333, 721–745 (2003).Article 
    CAS 
    PubMed 

    Google Scholar 
    Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60, 2126–2132 (2004).Article 
    PubMed 

    Google Scholar 
    Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D Biol. Crystallogr. 66, 213–221 (2010).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yamashita, K., Palmer, C. M., Burnley, T. & Murshudov, G. N. Cryo-EM single-particle structure refinement and map calculation using Servalcat. Acta Crystallogr. D Struct. Biol. 77, 1282–1291 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Edgar, R. C. Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26, 2460–2461 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Inoue, K. et al. Exploration of natural red-shifted rhodopsins using a machine learning-based Bayesian experimental design. Commun. Biol. 4, 362 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Salazar, G. et al. Gene expression changes and community turnover differentially shape the global ocean metatranscriptome. Cell 179, 1068–1083.e21 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chen, I.-M. A. et al. The IMG/M data management and analysis system v.6.0: new tools and advanced capabilities. Nucleic Acids Res. 49, D751–D763 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Nayfach, S. et al. A genomic catalog of Earth’s microbiomes. Nat. Biotechnol. 39, 499–509 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Sunagawa, S. et al. Metagenomic species profiling using universal phylogenetic marker genes. Nat. Methods 10, 1196–1199 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wickham, H. in ggplot2 (eds Gentleman, R., Hornik, K. & Parmigiani, G.) 189–201 (Springer, 2016).Katoh, K., Misawa, K., Kuma, K. & Miyata, T. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 30, 3059–3066 (2002).Article 
    CAS 
    PubMed 
    PubMed Central 

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

    Google Scholar 
    Nguyen, L.-T., Schmidt, H. A., von Haeseler, A. & Minh, B. Q. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 32, 268–274 (2015).Article 
    CAS 
    PubMed 

    Google Scholar 
    Hoang, D. T., Chernomor, O., von Haeseler, A., Minh, B. Q. & Vinh, L. S. UFBoot2: improving the ultrafast bootstrap approximation. Mol. Biol. Evol. 35, 518–522 (2018).Article 
    CAS 
    PubMed 

    Google Scholar  More

  • in

    Effect of different plant communities on NO2 in an urban road greenbelt in Nanjing, China

    Cui, Y. Z. et al. Rapid growth in nitrogen dioxide pollution over Western China, 2005–2013. Atmos. Chem. Phys. 16, 6207–6221. https://doi.org/10.5194/acp-16-6207-2016 (2016).Article 
    ADS 
    CAS 

    Google Scholar 
    Gu, J. B. et al. Ground-Level NO2 concentrations over China inferred from the Satellite OMI and CMAQ model simulations. Remote Sens. 9, 519. https://doi.org/10.3390/rs9060519 (2017).Article 
    ADS 

    Google Scholar 
    Cui, Y. Z. et al. Spatio-Temporal heterogeneous impacts of the drivers of NO2 pollution in Chinese cities: Based on satellite observation data. Remote Sens. 14, 3487. https://doi.org/10.3390/rs14143487 (2022).Article 
    ADS 

    Google Scholar 
    Huang, Z. Y., Xu, X. K., Ma, M. G. & Shen, J. W. Assessment of NO2 population exposure from 2005 to 2020 in China. Environ. Sci. Pollut. Res. https://doi.org/10.1007/s11356-022-21420-6 (2022).Article 

    Google Scholar 
    Zheng, Z. H., Yang, Z. W., Wu, Z. F. & Marinello, F. Spatial variation of NO2 and its impact factors in China: An application of sentinel-5P products. Remote Sens. 11, 1939. https://doi.org/10.3390/rs11161939 (2019).Article 
    ADS 

    Google Scholar 
    Bignal, K. L., Ashmore, M. R., Headley, A. D., Stewart, K. & Weigert, K. Ecological impacts of air pollution from road transport on local vegetation. Appl. Geochem. 22, 1265–1271. https://doi.org/10.1016/j.apgeochem.2007.03.017 (2007).Article 
    ADS 
    CAS 

    Google Scholar 
    Zhu, Y. J. et al. Spatiotemporally mapping of the relationship between NO2 pollution and urbanization for a megacity in Southwest China during 2005–2016. Chemosphere 220, 155–162. https://doi.org/10.1016/j.chemosphere.2018.12.095 (2019).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Stieb, D. M. et al. A national study of the association between traffic-related air pollution and adverse pregnancy outcomes in Canada, 1999–2008. Environ. Res. 148, 513–526. https://doi.org/10.1016/j.envres.2016.04.025 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Hu, Y. et al. Associations between total mortality and personal exposure to outdoor-originated NO2 in 271 Chinese cities. Atmos. Environ. https://doi.org/10.1016/j.atmosenv.2020.118170 (2021).Article 

    Google Scholar 
    Han, K. M. Temporal analysis of OMI-Observed tropospheric NO2 columns over east Asia during 2006–2015. Atmosphere 10, 658 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    EEA. Air quality in Europe—2016 report. European Environment Agency EEA Report No 28/2016. Retrieved 2 Dec 2016 from: http://www.eea.europa.eu/publications/air-quality-in-europe-2016Ahmad, A. et al. A comparative study on capability of different tree species in accumulating heavy metals from soil and ambient air. Chemosphere 172, 459–467. https://doi.org/10.1016/j.chemosphere.2017.01.045 (2017).Article 
    ADS 
    CAS 

    Google Scholar 
    Erin, R. D., Bryan, K. P., Amy, X. L. & Ronald, C. C. Laboratory measurements of stomatal NO2 deposition to native California trees and the role of forests in the NOx cycle. Atmos. Chem. Phys. 22, 14023–14041. https://doi.org/10.5194/acp-20-14023-2020 (2020).Article 
    CAS 

    Google Scholar 
    Takahashi, M. et al. Differential assimilation of nitrogen dioxide by 70 taxa of roadside trees at an urban pollution level. Chemosphere 61, 633–639. https://doi.org/10.1016/j.chemosphere.2005.03.033 (2005).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Guo, L. L., Li, B. F. & Chen, H. A. A review of urban Micro-climate research on block scale in China. Urban Dev. Stud. 24, 75–81. https://doi.org/10.3969/j.issn.10063862.2017.01.010 (2017).Article 

    Google Scholar 
    Jung, S. & Yoon, S. Analysis of the effects of floor area ratio change in urban street canyons on microclimate and particulate matter. Energies 14, 714. https://doi.org/10.3390/en14030714 (2021).Article 
    CAS 

    Google Scholar 
    Yin, S. et al. Quantifying air pollution attenuation within urban parks: An experimental approach in Shanghai, China. Environ. Pollut. 159, 2155–2163. https://doi.org/10.1016/j.envpol.2011.03.009 (2011).Article 
    CAS 
    PubMed 

    Google Scholar 
    Lin, C., Feng, X. F. & Heal, M. R. Temporal persistence of intra-urban spatial contrasts in ambient NO2, O3 and Ox in Edinburgh, UK. Atmos. Pollut. Res. 7, 734–741. https://doi.org/10.1016/j.apr.2016.03.008 (2016).Article 

    Google Scholar 
    Brantley, H. L., Hagler, G. S. W., Deshmukh, P. J. & Baldauf, R. W. Field assessment of the effects of roadside vegetation on near-road black carbon and particulate matter. Sci. Total Environ. 468, 120–129. https://doi.org/10.1016/j.scitotenv.2013.08.001 (2014).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Irga, P. J., Burchett, M. D. & Torpy, F. R. Does urban forestry have a quantitative effect on ambient air quality in an urban environment?. Atmos. Environ. 120, 173–181. https://doi.org/10.1016/j.atmosenv.2015.08.050 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    Tong, Z. M., Baldauf, R. W., Isakov, V., Deshmunk, P. & Zhang, K. M. Roadside vegetation barrier design to mitigate near-road air pollution impacts. Sci. Total Environ. 541, 920–927. https://doi.org/10.1016/j.scitotenv.2015.09.067 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Setälä, H., Viippola, V., Rantalainen, A. L., Pennanen, A. & Yli-Pelkonen, V. Does urban vegetation mitigate air pollution in northern conditions?. Environ. Pollut. 183, 104–112. https://doi.org/10.1016/j.envpol.2012.11.010 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Xing, Y. & Brimblecombe, P. Role of vegetation in deposition and dispersion of air pollution in urban parks. Atmos. Environ. 201, 73–83. https://doi.org/10.1016/j.atmosenv.2018.12.027 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    Xu, C., Wang, Y. P. & Li, L. L. Study on spatiotemporal distribution of the tropospheric NO2 column concentration in China and its relationship to energy consumption based on the time-series data from 2005 to 2013. Energy Sources Part A 42, 2130–2144. https://doi.org/10.1080/15567036.2019.1607931 (2020).Article 
    CAS 

    Google Scholar 
    Xu, J. H., Lindqvist, H., Liu, Q. F., Wang, K. & Wang, L. Estimating the spatial and temporal variability of the ground-level NO2 concentration in China during 2005–2019 based on satellite remote sensing. Atmos. Pollut. Res. 12, 57–67. https://doi.org/10.1016/j.apr.2020.10.008 (2021).Article 
    CAS 

    Google Scholar 
    Daniel, L. G. et al. TROPOMI NO2 in the United States: A detailed look at the annual averages, weekly cycles, effects of temperature, and correlation with surface NO2 concentrations. Earths Feature 9, 4. https://doi.org/10.1029/2020EF001665 (2021).Article 
    CAS 

    Google Scholar 
    Mavroidis, I. & Chaloulakou, A. Long-term trends of primary and secondary NO2 production in the Athens area. Variation of the NO2/NOx ratio. Atmos. Environ. 45, 6872–6879. https://doi.org/10.1016/j.atmosenv.2010.11.006 (2011).Article 
    ADS 
    CAS 

    Google Scholar 
    Van der, A. R. J. et al. Detection of the trend and seasonal variation in tropospheric NO2 over China. J. Geophys. Res. Atmos. https://doi.org/10.1029/2005JD006594 (2006).Article 

    Google Scholar 
    Salama, D. S. et al. Satellite observations for monitoring atmospheric NO2 in correlation with the existing pollution sources under arid environment. Model. Earth Syst. Environ. 8, 4103–4121. https://doi.org/10.1007/s40808-022-01352-3 (2022).Article 
    PubMed 

    Google Scholar 
    Ahmad, S. S. & Aziz, N. Spatial and temporal analysis of ground level ozone and nitrogen dioxide concentration across the twin cities of Pakistan. Environ. Monit. Assess. 185, 3133–3147. https://doi.org/10.1007/s10661-012-2778-7 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Khaled, G., Abdulaziz, A., Watheq, A. & Mumin, A. Analysis of NOx, NO and NO2 ambient levels in Dhahran, Saudi Arabia. Urban Clim. 21, 232–242. https://doi.org/10.2495/AIR170081 (2017).Article 

    Google Scholar 
    Casquero-Vera, J. A. et al. Impact of primary NO2 emissions at different urban sites exceeding the European NO2 standard limit. Sci. Total Environ. 646, 1117–1125 (2019).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Desyana, R. D., Sulistyantara, B., Nasrullah, N. & Fatimah, I. S. Study of the effectiveness of several tree canopy types on roadside green belt in influencing the distribution of NO2 gas emitted from transportation. EES https://doi.org/10.1088/1755-1315/58/1/012045 (2017).Article 

    Google Scholar 
    Rotach, M. W. Profiles of turbulence statistics in and above an urban street canyon. Atmos. Environ. 29, 1473–1486. https://doi.org/10.1016/1352-2310(95)00084-C (1995).Article 
    ADS 
    CAS 

    Google Scholar 
    Luo, M. Study on Air Pollutants Removal Effects of Green Space with Different Community Structures (Huazhong Agricultural University, 2013).
    Google Scholar 
    Rao, M., George, L. A., Rosenstiel, T. N., Shandas, V. & Dinno, A. Assessing the relationship among urban trees, nitrogen dioxide, and respiratory health. Environ. Pollut. 194, 96–104. https://doi.org/10.1016/j.envpol.2014.07.011 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Yli-Pelkonen, V., Viippola, V., Kotze, D. J. & Setala, H. Greenbelts do not reduce NO2 concentrations in near-road environments. Urban Clim. 21, 306–317. https://doi.org/10.1016/j.uclim.2017.08.005 (2017).Article 

    Google Scholar 
    Fantozzi, F., Monaci, F., Blanusa, T. & Bargagli, R. Spatio-temporal variations of ozone and nitrogen dioxide concentrations under urban trees and in a nearby open area. Urban Clim. 12, 119–127. https://doi.org/10.1016/j.uclim.2015.02.001 (2015).Article 

    Google Scholar 
    Nie, L., Deng, Z. H. & Chen, Q. B. SO2 and NOx purify-cation ability of forest in Kunming City. J. West China For. Sci. 44, 116–120 (2015).
    Google Scholar 
    Baldauf, R. Roadside vegetation design characteristics that can improve local, near-road air quality. Transp. Res. Part D 52, 354–361. https://doi.org/10.1016/j.trd.2017.03.013 (2017).Article 

    Google Scholar 
    Lai, D. Y., Liu, Y. Q., Liao, M. C. & Yu, B. Q. Effects of different tree layouts on outdoor thermal comfort of green space in summer Shanghai. Urban Clim. 47, 101398 (2023).Article 

    Google Scholar 
    Lai, D., Liu, W., Gan, T., Liu, K. & Chen, Q. A review of mitigating strategies to improve the thermal environment and thermal comfort in urban outdoor spaces. Sci. Total Environ. 661, 337–353 (2019).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar  More

  • in

    Shifts from cooperative to individual-based predation defense determine microbial predator-prey dynamics

    In co-culture with the bacterivorous flagellate Poteriospumella lacustris, the prey bacterium Pseudomonas putida exhibited a characteristic succession of predation defenses. The initial and the final defense differed substantially from one another with regard to their mechanism and their population-level benefits to the bacteria.Our results strongly indicate that the initial bacterial defense falls into the category of chemical defense, and is regulated by phenotypic plasticity. This would require P. putida to be able to sense predator density and to regulate the excretion of inhibitory substances accordingly. Because a considerable proportion of the P. putida genome is known to be involved in regulation and signal transduction allowing for very flexible responses to environmental triggers [41] both conditions are likely to be met. The filtrate exposure tests (Fig. 3) provide specific evidence for the ability of P. putida KT2440 to up- and downregulate the excretion of compounds inhibiting flagellate growth in response to grazing pressure. Previous research [25] corroborated the ability of P. putida to escape grazing from bacterivorous flagellates through induced responses like aggregation or biofilm formation.To provide a possible characterization for the apparent bacterial toxin, the whole-genome sequences of P. putida KT2440 obtained here were aligned against the antiSMASH [42] database. The output suggests the existence of non-ribosomal peptide synthetase clusters mediating the production of pyoverdines, a particular class of siderophores. The latter are molecules released by bacteria into the environment, which enhance the uptake of essential metals like, e.g., iron under deficient conditions. Specific pyoverdines associated with P. putida KT2440 have previously been identified [43]. Recent findings have shown that the benefits from siderophore production are not limited to competitive advantages gained from enhanced resource exploitation [44]. Pyoverdines were also demonstrated to determine the virulence of Pseudomonads via the damage of mitochondria in colonized hosts [45]. Moreover, pyoverdines were shown to be involved in the inducible defense of P. putida against predatory myxobacteria [46]. Such multiple functions have been reported for a number of bacterial metabolites, especially in Pseudomonads [47], and the particular combination of pyoverdin effects would explain the observed simultaneous flagellate inhibition and promoted bacterial growth.In contrast to the initial chemical defense of P. putida, the subsequent filamentation clearly provides an example of rapid evolution. Although the responsible mutation(s) could only be pinpointed in a few isolates so far (Table S1), there is no doubt about the genetic manifestation and heritability of the filamentous phenotype due to its demonstrated non-reversible nature.Only recently, similar observations were made by long-term co-cultivation of Pseudomonas fluorescence with the amoeboid predator Neaglena grubei [48]. In that system, protective adaptations like enhanced biofilm formation and altered motility were traced down to mutations in two particular genes (wspF, amrZ).From the perspective of the bacterial population, filamentation appears to be a much less efficient defense mechanism than toxin production. This is clearly reflected by the ratio of prey to predator biomass, which differed by two orders of magnitude between the initial and final defense (Table 6). It raises the question of why bacteria would abandon a highly effective form of defense in favor of a much less effective one. As demonstrated experimentally, adaptation of predators to the toxin can be excluded as a cause (Fig. 4). Moreover, it was not instantly evident how the small-sized flagellate was ultimately able to persist in large numbers given a very high proportion of completely inedible prey individuals (Fig. 1D and Fig. S2).Table 6 Average abundance of predator and prey during the temporary steady state following the initial bacterial defense (day 13–16) and during the final steady state (beyond day 30).Full size tableTo develop a comprehensive understanding of the system addressing the questions raised above, we set up a semi-continuous differential equation model to simulate the dynamics of predator and prey phenotypes. The model considers seven state variables (carbon, densities of four bacterial phenotypes, flagellate density, and toxin concentration) whose dynamics are controlled by nine processes (Table 3, Fig. 2). In addition to microbial growth and grazing, the model implements a phenotypically plastic predation defense (toxin production) as well as a genetic defense (filamentation) which arises via mutation. The particular assumptions implemented in the model are as follows:Dual effect of bacterial metabolitesIn line with the above discussion on siderophore-like compounds, secondary metabolites excreted by P. putida were assumed to exhibit a dual function, both inhibiting the growth of flagellates and allowing for a more efficient exploitation of the resources by bacteria. The inhibition of predators was demonstrated directly (Figs. 3 and 4) while enhanced resource exploitation was inferred from bacterial abundances in co-cultures exceeding the carrying capacity observed in predator-free controls (Fig. 1A, day 11–18).Metabolite production is costlyThe production of bacterial metabolites was assumed to be associated with a slight fitness cost [49] since resources are diverted from reproduction, thus resulting in a lowered growth rate of toxin-producing bacteria. The assumed fitness cost of 11% (parameter cBx in Table 5) allowed for the best agreement between simulated and observed data and is in agreement with data on the cost of pyoverdine production by P. aeruginosa [50]. The cost only manifests when toxin production is upregulated.Predator recognition and quorum sensing interactIn the model, the production of bacterial metabolites is upregulated when the two conditions of high flagellate abundance and high bacterial abundance coincide. That is, the expression of the toxin-based bacterial defense is assumed to be jointly controlled by predator recognition and quorum sensing (QS). Examples for such joint control of bacterial defenses have been reported previously [8, 26, 51]. The involvement of QS in chemical defense strategies is particularly likely as effective toxin concentrations can only be reached when producers are highly abundant. While multiple QS systems have been described for other Pseudomonads, only a single system has been identified in P. putida KT2440 so far [52, 53].Mutation rates are conditional on stressThe emergence of mutations resulting in the filamentation of P. putida was assumed to be conditional on a high ambient concentration of bacterial metabolites. The latter was considered as a proxy for bacterial stress which can affect mutagenesis either directly or indirectly by a variety of mechanisms [54,55,56]. Without this assumption, the almost synchronous appearance of filaments in all replicates at a late point in time would be very difficult to explain. Specifically, if mutation frequencies were high, filaments would become the predominant phenotype early (Fig. S3) which contradicts observations. On the other hand, if frequencies were low but unconditional, the timing of filament appearance should vary between replicates, which is in contrast to observations either (Fig. 1B).Filamentation is associated with a fitness costMeasurements of growth rate constants revealed a significant fitness disadvantage of filamentous isolates in comparison to single-celled, undefended isolates (p  More