More stories

  • in

    Quantifying finer-scale behaviours using self-organising maps (SOMs) to link accelerometery signatures with behavioural patterns in free-roaming terrestrial animals

    Data collectionData were collected in the Sunshine Coast region in Queensland, Australia (− 26.65° S, 153.07° E), from February to April 2019. All methods were carried out in accordance with relevant guidelines and regulations. All experimental protocols and methods were approved and carried out in compliance with the ARRIVE guidelines under the approval of the University of the Sunshine Coast (USC) Animal Ethics permit (ANA/16/109T); Human Ethics permit (A181114) and in conjunction with the Sunshine Coast Council (SCC) Local Law permit (OM18/19).
    Animals used in the trialsWe recruited 10 domestic cats through an approved media release (males n = 6; females n = 4; weight 2.8–8.4 kg; age 1.5–12 years; body length 38–53 cm; foreleg length 16–19 cm). As per the Sunshine Coast Council local law requirements, all cats had to be neutered, registered and microchipped to participate in the study.
    EquipmentWe fitted each cat with a retail harness, to which we attached a tri-axial accelerometer (AX3; Axivity, Newcastle University, UK; 23 × 32.5 × 8.9 mm; 11 g) using cable ties (Fig. 1a). The accelerometer was initialised using the Open Movement Graphical User Interaction application (OMGUI; V1.0.0.37). Because a trade-off exists between data resolution and battery life, we logged data at 50 Hz and with a dynamic range of ± 8 g, with a 13-bit resolution, similar to a previous study23. When combined with the in-built memory storage capacity of 512 MB, and battery limitations, this configuration resulted in a maximum of 8–14 days of data collection. The quartz Real Time Clock and calendar provided a timestamp with a frequency of 32.768 kHz and a precision of ± 50 ppm, with manufacturer specifications indicating a drift of 0.18 s per hour. To overcome this drift over the eight days, we calibrated devices by video recording the signals of five claps/taps on the device, at the start and end of each individual data collection period, and also at random times during the day.Figure 1(a) The anatomical position of the accelerometer (AX3) on the sternum of the cat. (b) The activity of swatting stimulated by the use of a feather. (c) The axis orientation of the accelerometer planes, which are represented in the accelerometer trace data in the MATLAB interface. Fore-aft (surge), lateral (sway) and dorso-ventral (heave) movement is reflected in the X, Y and Z signals.Full size imageWe positioned the accelerometer on the scapular brace-strap of the harness, inverted such that the accelerometer was on the sternum of the cat (Fig. 1a–c). Field trials over four months on four cats in the study determined that this position, in comparison with mounting on the dorsal cranial median plane, did not interfere with the animals’ balance; it also removed all of the abnormal movement behaviours and unnecessary discomfort to the cat2. The positioning of the logging device on the frontal anterior, median plane, resulted in the primary axis for fore-aft (surge), lateral (sway) and dorso-ventral (heave) movement to be reflected in the X, Y and Z signals, respectively (Fig. 1c).The accelerometer harness was used in conjunction with the CatBib for the relevant treatment periods. The total combined mass of the harness, accelerometer and Catbib was to 34.1 g, with a minimum cat mass of 2.8 kg, suggesting the equipment did not weigh above 1.2% of total body weight in any cat studied. The CatBib is a prey protector device, manufactured from a lightweight, washable neoprene material, that is attached to a cat’s safety collar (Fig. 1b). The dimensions of the bib are 17.5 mm × 17.5 mm × 6.5 mm, with a total mass of 23.1 g and it is purple in colour. All cats adjusted to the harness and CatBib within the first hour of deployment and no subsequent adjustments were required. All cats had unrestricted access to roam freely outside during the eight days of field trials.To capture training data, each cat was filmed with a GoPro + 3 Hero device (H.264—1920 × 1080; f/2.8; 60 fps), undertaking natural or stimulated active behaviours through play (Fig. 1b). These activities or behaviours were manually documented to track the activity, date and the timestamps. We conducted two treatments over the eight days: in the first, cats were fitted with CatBib, whereas in the other, bibs were not worn. Each treatment was conducted for four consecutive days, and the sequence of treatments for each cat was randomised. The accelerometer device on the harness was left on the cats for the entire field trial and recorded continuously for the eight days (~ 192 h per cat; total = 2304 h).Data analysisEach accelerometer trace file was exported as a raw binary file through OMIGUI and imported into a custom-built MATLAB GUI. To build our training dataset, the video file timestamp information, determined using Mediainfo (version 18.08, 2018), was used to define the start time for a subset of the accelerometer trace, and the video length to define the end point (Supp. Fig. 1). Offsets between the accelerometer trace and video files were determined using the closest calibrated tap signal trace for each day. We were able to watch each video file in synchrony with the accelerometer trace, and manually annotate each movement/activity from the video files to the accelerometer subset (Clemente et al.)24 (Supp. 1.1. Matlab interface instructions; Supp. Fig. 1).We grouped activities according to behaviour into three classes: Sedentary, Eating and Locomotive and Hunting. We further subdivided each group into behaviours. Sedentary included lying, sitting, grooming and watching; Eating and Locomotive included—eating/drinking, walking, trotting; and for Hunting—galloping, jumping, pouncing, swatting, biting/holding (Supp. Table 1).The accelerometer trace was then further divided into rolling epochs of 50 samples in length, using 1 s duration at 50 Hz to ensure intensive acceleratory bursts of short duration such as jumping and pouncing are captured. The behaviour with the maximum frequency within each epoch was assigned as that epoch’s label. Raw accelerometer data in each epoch was assigned as that epoch’s label. Raw accelerometer data in each epoch was summarized using 26 of the most effective variables for procedure accuracy identified by Tatler et al.25. We included: axial acceleration (X, Y, Z),mean acceleration (X, Y, Z); minimum acceleration.(X, Y, Z); maximum acceleration (X, Y, Z); standard deviation of acceleration (X, Y, Z); Signal Magnitude Area, minimum Overall Dynamic Body Acceleration (ODBA); maximum ODBA, minimum Vectorial Dynamic Body Acceleration VDBA; maximum VDBA, sum ODBA; sum VDBA; correlation (XY, YZ, XZ); skewness (X, Y, Z); and kurtosis (X, Y, Z)25 (See Supp. Table 2 for a detailed description of each variable). Finally, we coded the two treatments: BibON and BibOFF and included this information in the training data set.Classification modellingTo determine whether we could predict cat hunting behaviours, we analysed the training data sets using a Kohonen super Self Organising Map (SOM) in the R package ‘Kohonen’ version 2.0.1926,27.Machine learning procedures such as random forest and support vector machines each provide computationally powerful methods of data classification, however each method is not equal in how it visualises its output. SOMS have been used in behavioural studies10,13,14,15 for their ability to efficiently create easily interpreted maps and identify patterns of behaviour. Self-organizing maps differ from other artificial neural networks as they apply competitive learning as opposed to error-correction learning. In this study, a self-organising map algorithm was chosen for its efficiency in visualising multi-dimensional and complex data onto an easily interpreted two dimensional map output. SOMs also have the ability to visualise which variables are most influential with the use of component planes (Fig. 3b–e) and unlike other procedures mentioned, SOMs use cluster analysis which in this study aids in identifying similar behaviours and visualising them closer together (in clusters) on the map output.To prepare data for the SOM function a random sample of the classifiers for the trained data were extracted, along with their associated behaviour, and combined into a list with 2 elements (measurements and activity). This list was then input into the function supersom.R function, with the grid argument defined using the somgrid.R function [e.g. supersom(TrainingData, grid = somgrid(7, 7, “hexagonal”))]. The 7 × 7 grid function was chosen based on a sensitivity analysis exploring all combinations of grids between 4 to 9 units in length (n = 36, Supp. Fig. 2). The 7 × 7 grid represented the grid which produced the highest accuracy and map symmetry26,28,29. We further tested the effect of the number of times the complete data set is presented to the network by varying the rlen argument in the supersom.R function. We found no obvious increase in overall accuracy with increased iterations, and therefore used the default length of 100 times (Supp. Figure 3). Each supersom procedure created was then tested using the predict.R function, with the newdata argument directed to a testing data set, which was a similar 2 element list containing all samples not included in the training data set. The result of this test was then assembled into a confusion matrix using the table.R function with predictions compared with the known behaviours in the test data set [e.g. table(predictions = ssom.pred$predictions$activity, activity = testData$activity) ]. A confusion matrix is a table where each row represents the instances in a predicted class, while each column represents the instances in the observed class, allowing mislabelled epochs to be easily identified. The confusion matrix was then finally used to compute four specific accuracy metrics—sensitivity (or recall), precision, specificity, as well as overall accuracy.To identify relationships between the size of training dataset, we trained a randomised subset of the BibOFF training data, to predict the remaining BibOFF data from all cats. We tested 35 different subset sample sizes from 100 to 100,000, replicating each sample size ten times (with replacement) to determine variation at each sample size.We then tested the extent to which accelerometer traces are modified by the presence of the CatBib. This modification was indicated by a change in overall prediction accuracy of the SOM between BibOFF and BibON treatments. To do this, we trained the SOM using a subset of the trained data for BibOFF and tested it against annotated classified BibON samples. In order to statistically compare results from bootstrap resampling, we took the median among bootstrap samples as the estimate of performance and quantified uncertainty using the corresponding 2.5th and 97.5th percentiles to represent credible 95% confidence intervals (CIs). We chose the median as a measure of central tendency, because resampling distributions were truncated at 1, so were skewed. If CIs for any pair of estimates (medians) do not overlap, then this is evidence of a significant difference between the estimates. If, however, one estimated median fell within the confidence interval for another estimate, then this was used as evidence of a lack of significant difference. For all other outcomes, differences are equivocal, and we interpreted them tentatively on the basis of the relative overlap in CIs.Finally we compared the output of the SOM with the output from a decision tree classification method using a random forest (RF) approach from the randomForest.R package30. We chose random forest as a comparison as this method has previously been shown to perform better than other similar methods (e.g. k-nearest neighbour, support vector machine, and naïve Bayes) when classifying behavioural data on free moving animals25,31. We trained both the SOM and RF procedures using the same 20,000 randomly selected epochs, and compared the overall accuracy for predicting the behaviour for the remaining ~ 192,000 epochs. The SOM was built using a 7 × 7 grid patterns, with the rlen argument set to 100. The RF was built with the number of trees set to 100 and the number of variables randomly sampled as candidates at each split set to 4. More

  • in

    Water quality drives the regional patterns of an algal metacommunity in interconnected lakes

    1.Leibold, M. A. et al. The metacommunity concept: a framework for multi-scale community ecology. Ecol. Lett. 7, 601–613. https://doi.org/10.1111/j.1461-0248.2004.00608.x (2004).Article 

    Google Scholar 
    2.McGill, B. J., Enquist, B. J., Weiher, E. & Westoby, M. Rebuilding community ecology from functional traits. Trends Ecol. Evol. 21, 178–185. https://doi.org/10.1016/j.tree.2006.02.002 (2006).Article 
    PubMed 

    Google Scholar 
    3.Kraft, N. et al. Community assembly, coexistence, and the environmental filtering metaphor. Funct. Ecol. https://doi.org/10.1111/1365-2435.12345 (2014).Article 

    Google Scholar 
    4.de la Sancha, N. U., Higgins, C. L., Presley, S. J. & Strauss, R. E. Metacommunity structure in a highly fragmented forest: has deforestation in the Atlantic Forest altered historic biogeographic patterns?. Divers. Distrib. 20, 1058–1070. https://doi.org/10.1111/ddi.12210 (2014).Article 

    Google Scholar 
    5.Leibold, M. & Mikkelson, G. Coherence, species turnover, and boundary clumping: Elements of meta-community structure. Oikos 97, 237–250. https://doi.org/10.1034/j.1600-0706.2002.970210.x (2002).Article 

    Google Scholar 
    6.Presley, S., Higgins, C. & Willig, M. A comprehensive framework for the evaluation of metacommunity structure. Oikos 119, 908–917. https://doi.org/10.1111/j.1600-0706.2010.18544.x (2010).Article 

    Google Scholar 
    7.Dallas, T. & Drake, J. M. Relative importance of environmental, geographic, and spatial variables on zooplankton metacommunities. Ecosphere 5, 1–13. https://doi.org/10.1890/ES14-00071.1 (2014).Article 

    Google Scholar 
    8.Heino, J., Mykrä, H. & Muotka, T. Temporal variability of nestedness and idiosyncratic species in stream insect assemblages. Divers. Distrib. 15, 198–206. https://doi.org/10.1111/j.1472-4642.2008.00513.x (2009).Article 

    Google Scholar 
    9.Henriques-Silva, R., Lindo, Z. & Peres-Neto, P. R. A community of metacommunities: exploring patterns in species distributions across large geographical areas. Ecology 94, 627–639. https://doi.org/10.1890/12-0683.1 (2013).Article 
    PubMed 

    Google Scholar 
    10.Dallas, T. & Drake, J. M. Relative importance of environmental, geographic, and spatial variables on zooplankton metacommunities. Ecosphere 5, art104. https://doi.org/10.1890/ES14-00071.1 (2014).Article 

    Google Scholar 
    11.Erős, T. et al. Quantifying temporal variability in the metacommunity structure of stream fishes: The influence of non-native species and environmental drivers. Hydrobiologia 722, 31–43. https://doi.org/10.1007/s10750-013-1673-8 (2014).Article 

    Google Scholar 
    12.Fernandes, I. M., Henriques-Silva, R., Penha, J., Zuanon, J. & Peres-Neto, P. R. Spatiotemporal dynamics in a seasonal metacommunity structure is predictable: The case of floodplain-fish communities. Ecography 37, 464–475. https://doi.org/10.1111/j.1600-0587.2013.00527.x (2014).Article 

    Google Scholar 
    13.Tonkin, J. D. et al. The role of dispersal in river network metacommunities: Patterns, processes, and pathways. Freshw. Biol. 63, 141–163. https://doi.org/10.1111/fwb.13037 (2018).Article 

    Google Scholar 
    14.Kim, S., Chung, S., Park, H., Cho, Y. & Lee, H. Analysis of environmental factors associated with cyanobacterial dominance after river weir installation. Water https://doi.org/10.3390/w11061163 (2019).Article 

    Google Scholar 
    15.Deng, J. et al. Effects of nutrients, temperature and their interactions on spring phytoplankton community succession in Lake Taihu, China. PLoS ONE 9, e113960–e113960. https://doi.org/10.1371/journal.pone.0113960 (2014).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    16.Yang, J., Jiang, H., Liu, W. & Wang, B. Benthic algal community structures and their response to geographic distance and environmental variables in the Qinghai-Tibetan lakes with different salinity. Front. Microbiol. 9, 578–578. https://doi.org/10.3389/fmicb.2018.00578 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    17.Zhou, J. et al. Microbial community structure and associations during a marine dinoflagellate bloom. Front. Microbiol. 9, 1201. https://doi.org/10.3389/fmicb.2018.01201 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    18.RDevelopmentCoreTeam. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2013).
    Google Scholar 
    19.Baird, R. B. Standard Methods for the Examination of Water and Wastewater 23rd edn. (Water Environment Federation, American Public Health Association, 2017).
    Google Scholar 
    20.Liaw, A. & Wiener, M. Classification and regression by randomForest. R News 2, 18–22 (2002).
    Google Scholar 
    21.Cajo, J. F. T. B. Canonical correspondence analysis: a new eigenvector technique for multivariate direct gradient analysis. Ecology 67, 1167–1179. https://doi.org/10.2307/1938672 (1986).Article 

    Google Scholar 
    22.Tuomisto, H. A diversity of beta diversities: straightening up a concept gone awry. Part 2. Quantifying beta diversity and related phenomena. Ecography 33, 23–45. https://doi.org/10.1111/j.1600-0587.2009.06148.x (2010).Article 

    Google Scholar 
    23.Clements, F. E. Nature and structure of the climax. J. Ecol. 24, 252–284. https://doi.org/10.2307/2256278 (1936).Article 

    Google Scholar 
    24.Kurthen, A. L. et al. Metacommunity structures of macroinvertebrates and diatoms in high mountain streams, Yunnan, China. Front. Ecol. Evol. https://doi.org/10.3389/fevo.2020.571887 (2020).Article 

    Google Scholar 
    25.Vannote, R. L., Minshall, G. W., Cummins, K. W., Sedell, J. R. & Cushing, C. E. The river continuum concept. Can. J. Fish. Aquat. Sci. 37, 130–137. https://doi.org/10.1139/f80-017 (1980).Article 

    Google Scholar 
    26.López-González, C., Presley, S. J., Lozano, A., Stevens, R. D. & Higgins, C. L. Metacommunity analysis of Mexican bats: environmentally mediated structure in an area of high geographic and environmental complexity. J. Biogeogr. 39, 177–192. https://doi.org/10.1111/j.1365-2699.2011.02590.x (2012).Article 

    Google Scholar 
    27.Heino, J., Soininen, J., Alahuhta, J., Lappalainen, J. & Virtanen, R. Metacommunity ecology meets biogeography: effects of geographical region, spatial dynamics and environmental filtering on community structure in aquatic organisms. Oecologia 183, 121–137. https://doi.org/10.1007/s00442-016-3750-y (2017).ADS 
    Article 
    PubMed 

    Google Scholar 
    28.Heino, J. & Alahuhta, J. Elements of regional beetle faunas: faunal variation and compositional breakpoints along climate, land cover and geographical gradients. J. Anim. Ecol. 84, 427–441. https://doi.org/10.1111/1365-2656.12287 (2015).Article 
    PubMed 

    Google Scholar 
    29.Mallin, M. A., McIver, M. R., Ensign, S. H. & Cahoon, L. B. Photosynthetic and heterotrophic impacts of nutrient loading to blackwater streams. Ecol. Appl. 14, 823–838. https://doi.org/10.1890/02-5217 (2004).Article 

    Google Scholar 
    30.B-Béres, V. et al. Autumn drought drives functional diversity of benthic diatom assemblages of continental intermittent streams. Adv. Water Resour. 126, 129–136. https://doi.org/10.1016/j.advwatres.2019.02.010 (2019).ADS 
    Article 

    Google Scholar 
    31.Kagalou, I., Petridis, D. & Tsimarakis, G. Seasonal variation of water quality parameters and plankton in a shallow Greek lake. J. Freshw. Ecol. 18, 199–206. https://doi.org/10.1080/02705060.2003.9664485 (2003).CAS 
    Article 

    Google Scholar 
    32.Padisák, J., Crossetti, L. O. & Naselli-Flores, L. Use and misuse in the application of the phytoplankton functional classification: a critical review with updates. Hydrobiologia 621, 1–19. https://doi.org/10.1007/s10750-008-9645-0 (2009).Article 

    Google Scholar 
    33.Schabhüttl, S. et al. Temperature and species richness effects in phytoplankton communities. Oecologia 171, 527–536. https://doi.org/10.1007/s00442-012-2419-4 (2013).ADS 
    Article 
    PubMed 

    Google Scholar 
    34.Chen, S. et al. Geographical patterns of algal communities associated with different urban lakes in China. Int. J. Environ. Res. Public Health 17, 1009. https://doi.org/10.3390/ijerph17031009 (2020).Article 
    PubMed Central 

    Google Scholar 
    35.Hwang, S.-J., Kim, H.-S., Shin, J.-K., Oh, J.-M. & Kong, D.-S. Grazing effects of a freshwater bivalve (Corbicula leana Prime) and large zooplankton on phytoplankton communities in two Korean lakes. Hydrobiologia 515, 161–179. https://doi.org/10.1023/B:HYDR.0000027327.06471.1e (2004).Article 

    Google Scholar 
    36.Moss, B. et al. How important is climate? Effects of warming, nutrient addition and fish on phytoplankton in shallow lake microcosms. J. Appl. Ecol. 40, 782–792. https://doi.org/10.1046/j.1365-2664.2003.00839.x (2003).Article 

    Google Scholar 
    37.Chen, S. et al. Local habitat heterogeneity determines the differences in benthic diatom metacommunities between different urban river types. Sci. Total Environ. 669, 711–720. https://doi.org/10.1016/j.scitotenv.2019.03.030 (2019).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar  More

  • in

    Spatio-temporal distribution patterns of Plutella xylostella (Lepidoptera: Plutellidae) in a fine-scale agricultural landscape based on geostatistical analysis

    1.Zalucki, M. P. et al. Estimating the economic cost of one of the world’s major insect pests, Plutella xylostella: Just how long is a piece of string?. J. Econ. Entomol. 105, 1115–1129 (2012).PubMed 
    Article 

    Google Scholar 
    2.Li, Z. Y., Feng, X., Liu, S. S., You, M. S. & Furlong, M. J. Biology, ecology, and management of the diamondback moth in China. Annu. Rev. Entomol. 61(1), 277–296 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    3.Talekar, N. S. & Shelton, A. M. Biology, ecology, and management of the diamondback moth. Annu. Rev. Entomol. 38(1), 275–301 (1993).Article 

    Google Scholar 
    4.Zhu, L. et al. Population dynamics of diamondback moth, Plutella xylostella (L.) in northern China: The effect of migration, cropping patterns and climate. Pest Manag. Sci. 74(8), 1845–1853 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    5.Furlong, M. J., Wright, D. J. & Dosdall, L. M. Diamondback moth ecology and management: Problems, progress, and prospects. Annu. Rev. Entomol. 58, 517–541 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    6.Sayyed, A. H., Saeed, S., Noorulane, M. & Crickmore, N. Genetic, biochemical, and physiological characterization of spinosad resistance in Plutella xylostella (Lepidoptera: Plutellidae). J. Econ. Entomol. 101(5), 1658–1666 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    7.Machekano, H., Mvumi, B. M. & Nyamukondiwa, C. Loss of coevolved basal and plastic responses to temperature may underlie trophic level host-parasitoid interactions under global change. Biol. Control 118, 44–54 (2018).Article 

    Google Scholar 
    8.Chapman, J. W. et al. High-altitude migration of the diamondback moth Plutella xylostella to the U.K.: A study using radar, aerial netting, and ground trapping. Ecol. Entomol. 27(6), 641–650 (2002).Article 

    Google Scholar 
    9.Mazzi, D. & Dorn, S. Movement of insect pests in agricultural landscapes. Ann. Appl. Biol. 160(2), 97–113 (2012).Article 

    Google Scholar 
    10.Wei, S. J. et al. Genetic structure and demographic history reveal migration of the diamondback moth Plutella xylostella (Lepidoptera: Plutellidae) from the southern to Northern Regions of China. PLoS ONE 8(4), e59654 (2013).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    11.Fu, X., Xing, Z., Liu, Z., Ali, A. & Wu, K. Migration of diamondback moth, Plutella xylostella, across the Bohai Sea in northern China. Crop Prot. 64, 143–149 (2014).Article 

    Google Scholar 
    12.Li, Z. et al. Population dynamics and management of diamondback moth (Plutella xylostella) in China: The relative contributions of climate, natural enemies and cropping patterns. Bull. Entomol. Res. 106(2), 197–214 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    13.Machekano, H. et al. Disentangling factors limiting diamondback moth, Plutella xylostella (L.), spatio-temporal population abundance: A tool for pest forecasting. J. Appl. Entomol. 143, 670–682 (2019).CAS 
    Article 

    Google Scholar 
    14.Eziah, V. Y., Rose, H. A., Wilkes, M., Clift, A. D. & Mansfiled, S. Population dynamics of the diamondback moth Plutella xylostella L. (Lepidoptera: Yponomeutidae) in the Sydney region of Australia. Int. J. Biol. Chem. Sci. 4(4), 1062–1082 (2011).
    Google Scholar 
    15.Alam, T., Raju, S. V. S., Raghuraman, M. & Kumar, K. R. Population dynamics of diamondback moth, Plutella xylostella (L.) on cauliflower Brassica oleracea L. var. Botrytis in relation to weather factors of eastern uttar pradesh region. J. Exp. Zool. India 19(1), 289–292 (2016).
    Google Scholar 
    16.Karimzadeh, J., Bonsall, M. B. & Wright, D. J. Bottom-up and top-down effects in a tritrophic system: The population dynamics of Plutella xylostella (L.)-Cotesia plutellae (Kurdjumov) on different host plants. Ecol. Entomol. 29(3), 285–293 (2004).Article 

    Google Scholar 
    17.Soufbaf, M., Fathipour, Y., Karimzadeh, J. & Zalucki, M. P. Effects of plant availability on population size and dynamics of an insect community: Diamondback moth and two of its parasitoids. Bull. Entomol. Res. 104(4), 418–431 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    18.Li, Z. Y. et al. Population dynamics and ‘outbreaks’ of diamondback moth, Plutella xylostella, in Guangdong province, China: Climate or the failure of management?. J. Econ. Entomol. 105(3), 739–752 (2012).PubMed 
    Article 

    Google Scholar 
    19.Sutcliffe, L. M. E., Batáry, P., Becker, T., Orci, K. M. & Leuschner, C. Both local and landscape factors determine plant and Orthoptera diversity in the semi-natural grasslands of Transylvania, Romania. Biodivers. Conserv. 24(2), 229–245 (2015).Article 

    Google Scholar 
    20.Carrière, Y. et al. Effects of local and landscape factors on population dynamics of a cotton pest. PLoS ONE 7(6), e39862 (2012).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    21.Moroń, D., Skórka, P., Lenda, M., Celary, W. & Tryjanowski, P. Railway lines affect spatial turnover of pollinator communities in an agricultural landscape. Divers. Distrib. 23(9), 1090–1097 (2017).Article 

    Google Scholar 
    22.Skellern, M. P., Welham, S. J., Watts, N. P. & Cook, S. M. Meteorological and landscape influences on pollen beetle immigration into oilseed rape crops. Agric. Ecosyst. Environ. 241, 150–159 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    23.Meisner, M. H., Zaviezo, T. & Rosenheim, J. A. Landscape crop composition effects on cotton yield, Lygus hesperus densities and pesticide use. Pest Manag. Sci. 73(1), 232–239 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    24.Furlong, M. J. et al. Ecology of diamondback moth in Australian canola: Landscape perspectives and the implications for management. Aust. J. Exp. Agric. 48(12), 1494–1505 (2008).Article 

    Google Scholar 
    25.Rogers, C. D., Guimaraes, R. M. L., Evans, K. A. & Rogers, S. A. Spatial and temporal analysis of wheat bulb fly (Delia coarctata, Fallén) oviposition: Consequences for pest population monitoring. J. Pest Sci. 88, 75–86 (2014).Article 

    Google Scholar 
    26.Silva, G. A. et al. Control failure likelihood and spatial dependence of insecticide resistance in the tomato pinworm, Tuta absoluta. Pest Manag. Sci. 67, 913–920 (2011).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    27.Rossi, R. E., Mulla, D. J., Journel, A. G. & Franz, E. H. Geostatistical tools for modeling and interpreting ecological spatial dependence. Ecol. Monogr. 62(2), 277–314 (1992).Article 

    Google Scholar 
    28.Leibhold, A. M., Rossi, R. E. & Kemp, W. P. Geostatistics and geographic information systems in applied insect ecology. Annu. Rev. Entomol. 38(1), 303–327 (1993).Article 

    Google Scholar 
    29.Veran, S. et al. Modeling spatiotemporal dynamics of outbreaking species: Influence of environment and migration in a locust. Ecology 96(3), 737–748 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    30.Martins, J. C. et al. Assessing the spatial distribution of Tuta absoluta (lepidoptera: gelechiidae) eggs in open-field tomato cultivation through geostatistical analysis. Pest Manag. Sci. 74(1), 30–36 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    31.Cocco, A., Serra, G., Lentini, A., Deliperi, S. & Delrio, G. Spatial distribution and sequential sampling plans for Tuta absoluta (Lepidoptera: Gelechiidae) in greenhouse tomato crops. Pest Manag. Sci. 71(9), 1311–1323 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    32.Sciarretta, A., Zinni, A., Mazzocchetti, A. & Trematerra, P. Spatial analysis of Lobesia botrana (lepidoptera: tortricidae) male population in a mediterranean agricultural landscape in Central Italy. Environ. Entomol. 37(2), 382 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    33.Sciarretta, A. & Trematerra, P. Spatio-temporal distribution of Ceratitis capitata population in a heterogeneous landscape in Central Italy. J. Appl. Entomol. 135(4), 241–251 (2011).Article 

    Google Scholar 
    34.Fuzhou. https://baike.baidu.com/item/%E7%A6%8F%E5%B7%9E/165311?fr=Aladdin (2021).35.Fujian Meteorological Service Center. http://fj.cma.gov.cn/#qxfw (2021).36.Farias, P. R. S., Roberto, S. R., Lopes, J. R. S. & Perecin, D. Geostatistical characterization of the spatial distribution of Xylella fastidiosa sharpshooter vectors on citrus. Neotrop. Entmol. 33, 13–20 (2002).Article 

    Google Scholar 
    37.Cambardella, C. A. et al. Field-scale variability of soil proprieties in central Iowa soils. Soil Sci. Soc. Am. J. 58, 1240–1248 (1994).Article 

    Google Scholar 
    38.Zhou, C. B., Lin, Z. F., Xie, S. H. & Ji, X. C. Population dynamics of Plutella xylostella and its influence factors in Hainan. Plant Prot 36(5), 124–128 (2010) (in Chinese, English abstract).
    Google Scholar 
    39.Golizadeh, A. L. I., Kamali, K., Fathipour, Y. & Abbasipour, H. Temperature-dependent development of diamondback moth, Plutella xylostella (Lepidoptera: Plutellidae) on two Brassicaceous host plants. Insect Sci. 14(4), 309–316 (2007).Article 

    Google Scholar 
    40.Bhagat, P., Yadu, Y. K. & Sharma, G. L. Seasonal incidence and effect of abiotic factors on population dynamics of diamondback moth (Plutella xylostella L.) on cabbage (Brassica oleracea var. Capitata L.) crop. J. Enotomol. Zool. Stud. 6(2), 2001–2003 (2018).
    Google Scholar 
    41.Wang, E. G. & Zheng, Y. L. Seasonal abundance of diamondback moth, Plutella xylostella, adult in Linhai, Zhejiang. Chin. Bull. Entomol. 44(2), 271–274 (2007) (in Chinese, English abstract).ADS 

    Google Scholar 
    42.Lin, X. J., Xie, W. L., Liu, J. B. & Zeng, L. Investigation of the occurrence of Plutella xylostella in Guangzhou. Guangdong Agric. Sci. 36(16), 91–97 (2013) (in Chinese, English abstract).
    Google Scholar 
    43.Harcourt, D. G. Major mortality factors in the population dynamics of the diamondback moth, Plutella maculipennis (Curt.) (Lepidoptera: Plutellidae). Mean. Can. Entomol. 32, 55–66 (1963).Article 

    Google Scholar 
    44.Rahman, M. M., Zalucki, M. P. & Furlong, M. J. Diamondback moth egg susceptibility to rainfall: Effects of host plant and oviposition behavior. Entomol. Exp. Appl. https://doi.org/10.1111/eea.12816 (2019).Article 

    Google Scholar 
    45.Kobori, Y. & Amano, H. Effect of rainfall on a population of the diamondback moth, Plutella xylostella (Lepidoptera: Plutellidae). Appl. Entomol. Zool. 38(2), 249–253 (2003).Article 

    Google Scholar 
    46.Ayalew, G., Sciarretta, A., Baumgärtner, J., Ogol, C. & Löhr, B. Spatial distribution of diamondback moth, Plutella xylostella L. (Lepidoptera: Plutellidae), at the field and the regional level in Ethiopia. Int. J. Pest Manag. 54(1), 31–38 (2008).Article 

    Google Scholar 
    47.Mo, J., Greg, B., Mike, K. & Rick, R. Local dispersal of the diamondback moth (Plutella xylostella (L.)) (Lepidoptera: Plutellidae). Environ. Entomol. 32(1), 71–79 (2003).Article 

    Google Scholar 
    48.Xiong, L. G. et al. Biological characteristic of overwintering in the diamondback moth, Plutella xylostella. Plant Prot. 36, 90–93 (2010) (in Chinese, English abstract).
    Google Scholar 
    49.Campos, W. G., Schoereder, J. H. & Sperber, C. F. Does the age of the host plant modulate migratory activity of Plutella xylostella?. Entomol. Sci. 7(4), 323–329 (2004).Article 

    Google Scholar 
    50.Zhao, Z. H., Hui, C., He, D. H. & Ge, F. Effects of position within wheat field and adjacent habitats on the density and diversity of cereal aphids and their natural enemies. Biocontrol 58, 765–776 (2013).CAS 
    Article 

    Google Scholar 
    51.Sciarretta, A. & Trematerra, P. Geostatistical tools for the study of insect spatial distribution: Practical implications in the integrated management of orchard and vineyard pests. Plant Prot. Sci. 50(2), 97–110 (2014).Article 

    Google Scholar 
    52.Saeed, R., Sayyed, A. H., Shad, S. A. & Zaka, S. M. Effect of different host plants on the fitness of diamond-back moth, Plutella xylostella (Lepidoptera: Plutellidae). Crop Prot. 29(2), 178–182 (2010).Article 

    Google Scholar 
    53.Chen, L. L. et al. Cover crops enhance natural enemies while help suppressing pests in a tea plantation. Ann.. Entomol. Soc. Am. 112(4), 348–355 (2019).Article 

    Google Scholar  More

  • in

    Beyond coronavirus: the virus discoveries transforming biology

    Download PDF

    Mya Breitbart has hunted novel viruses in African termite mounds, Antarctic seals and water from the Red Sea. But to hit pay dirt, she has only to step into her back garden in Florida. Hanging around her swimming pool are spiny-backed orbweavers (Gasteracantha cancriformis) — striking spiders with bulbous white bodies, black speckles and six scarlet spikes that make them look like a piece of medieval weaponry. Even more striking for Breitbart, a viral ecologist at the University of South Florida in St Petersburg, was what was inside. When she and her colleagues collected a few spiders and ground them up, they found two viruses previously unknown to science1.Although we humans have been focused on one particularly nasty virus since early 2020, there are legions of other viruses out there waiting to be discovered. Scientists estimate that there are about 1031 individual viral particles inhabiting the oceans alone at any given time — 10 billion times the estimated number of stars in the known Universe.It’s becoming clear that ecosystems and organisms rely on viruses. Tiny but mighty, they have fuelled evolution for millions of years by shuttling genes between hosts. In the oceans, they slice open microorganisms, spilling their contents into the sea and flooding the food web with nutrients. “Without viruses,” says Curtis Suttle, a virologist at the University of British Columbia in Vancouver, Canada, “we would not be alive.”There are just 9,110 named species listed by the International Committee on Taxonomy of Viruses (ICTV), but that’s obviously a pitiful fraction of the total. In part, that’s because officially classifying a virus used to require scientists to culture a virus in its host or host cells — a time-consuming if not impossible process. It’s also because the search has been biased towards viruses that cause diseases in humans or organisms we care about, such as farm animals and crop plants. Yet, as the COVID-19 pandemic has reminded us, it’s important to understand viruses that might jump from one host to another, threatening us, our animals or our crops.
    The new scope of virus taxonomy: partitioning the virosphere into 15 hierarchical ranks
    Over the past ten years, the number of known and named viruses has exploded, owing to advances in the technology for finding them, plus a recent change to the rules for identifying new species, to allow naming without having to culture virus and host. One of the most influential techniques is metagenomics, which allows researchers to sample the genomes in an environment without having to culture individual viruses. Newer technologies, such as single-virus sequencing, are adding even more viruses to the list, including some that are surprisingly common yet remained hidden until now. It’s an exciting time to be doing this kind of research, says Breitbart. “I think, in many ways, now is the time of the virome.”In 2020 alone, the ICTV added 1,044 species to its official list, and thousands more await description and naming. This proliferation of genomes prompted virologists to rethink the way they classify viruses and helped to clarify their evolution. There is strong evidence that viruses emerged multiple times, rather than sprouting from a single origin.Even so, the true range of the viral world remains mostly uncharted, says Jens Kuhn, a virologist at the US National Institute of Allergy and Infectious Diseases facility at Fort Detrick, Maryland. “We really have absolutely no idea what’s out there.”Here, there and everywhereAll viruses have two things in common: each encases its genome in a protein-based shell, and each relies on its host — be it a person, spider or plant — to reproduce itself. But beyond that general pattern lie endless variations.There are minuscule circoviruses with only two or three genes, and massive mimiviruses that are bigger than some bacteria and carry hundreds of genes. There are lunar-lander-looking phage that infect bacteria and, of course, the killer spiky balls the world is now painfully familiar with. There are viruses that store their genes as DNA, and others that use RNA; there’s even a phage that uses an alternative genetic alphabet, replacing the chemical base A in the standard ACGT system with a different molecule, designated Z.

    Studies of the spiny-backed orbweaver found two viruses previously unknown to science.Credit: Scott Leslie/Minden Pictures/Alamy

    Viruses are so ubiquitous that they can turn up even when scientists aren’t looking for them. Frederik Schulz did not intend to study viruses as he pored over genome sequences from waste water. As a graduate student at the University of Vienna, in 2015 he was using metagenomics to hunt for bacteria. This involves isolating DNA from a whole mix of organisms, chopping it into bits and sequencing all of them. A computer program then assembles the bits into individual genomes; it’s like solving hundreds of jigsaw puzzles whose pieces have been jumbled up.Among the bacterial genomes, Schulz couldn’t help but notice a whopper of a virus genome — obvious because it carried genes for a viral shell — with a remarkable 1.57 million base pairs2. It turned out to be a giant virus, part of a group whose members are large in terms of both genome size and absolute size (typically, 200 nanometres or more across). These viruses infect amoebae, algae and other protists, putting them in a position to influence ecosystems both aquatic and terrestrial.
    Profile of a killer: the complex biology powering the coronavirus pandemic
    Schulz, now a microbiologist at the US Department of Energy Joint Genome Institute in Berkeley, California, decided to search for related viruses in metagenome data sets. In 2020, in a single paper3, he and his colleagues described more than 2,000 genomes from the group that contains giant viruses; before that, just 205 such genomes had been deposited in public databases.Virologists have also looked inwards to find new species. Viral bioinformatician Luis Camarillo-Guerrero worked with colleagues at the Wellcome Sanger Institute in Hinxton, UK, to analyse metagenomes from the human gut, and built a database containing more than 140,000 kinds of phage. More than half of these were new to science. Their study4, published in February, matched others’ findings that one of the most common viruses to infect the bacteria in our guts is a group known as crAssphage (named after the cross-assembly software that picked it up in 2014). Despite its abundance, not much is known about how it contributes to our microbiome, says Camarillo-Guerrero, who now works at DNA-sequencing company Illumina in Cambridge, UK.Metagenomics has turned up a wealth of viruses, but it ignores many, too. RNA viruses aren’t sequenced in typical metagenomes, so microbiologist Colin Hill at University College Cork, Ireland, and his colleagues looked for them in databases of RNAs, called metatranscriptomes. Scientists normally use these data to understand the genes in a population that are actively being turned into messenger RNA in to make proteins, but RNA virus genomes can show up, too. Using computational techniques to pull sequences out of the data, the team found 1,015 viral genomes in metatrancriptomes from sludge and water samples5. Again, they’d massively increased the number of known viruses with a single paper.

    The giant tupanvirus,found in amoebae, is more than 1,000 nanometres long and has the largest set of protein-coding genes of any known virus.Credit: J. Abrahão et al./Nature Commun.

    Although it’s possible for these techniques to accidentally assemble genomes that aren’t real, researchers have quality-control techniques to guard against this. But there are other blind spots. For instance, viral species whose members are very diverse are fiendishly difficult to find because it’s hard for computer programs to piece together the disparate sequences.The alternative is to sequence viral genomes one at a time, as microbiologist Manuel Martinez-Garcia does at the University of Alicante, Spain. He decided to try trickling seawater through a sorting machine to isolate single viruses, amplified their DNA, and got down to sequencing.On his first attempt, he found 44 genomes. One turned out to represent some of the most abundant viruses in the ocean6. This virus is so diverse — its genetic jigsaw pieces so varied from one virus particle to the next — that its genome had never popped up in metagenomics studies. The team calls it 37-F6, for its location on the original laboratory dish, but Martinez-Garcia jokes that, given its ability to hide in plain sight, it should have been named 007, after fictional superspy James Bond.Virus family treesThe James Bond of ocean viruses lacks an official Latin species name, and so do most of the thousands of viral genomes discovered by metagenomics over the past decade. Those sequences presented the ICTV with a dilemma: is a genome enough to name a virus? Until 2016, proposing a new virus or taxonomic group to the ICTV required scientists to have that virus and its host in culture, with rare exceptions. But that year, after a contentious but cordial debate, virologists agreed that a genome was sufficient7.Proposals for new viruses and groups poured in (see ‘Adding to the family’). But the evolutionary relationships between these viruses were often unclear. Virologists usually categorize viruses on the basis of their shapes (long and thin, say, or a head with a tail) or their genomes (DNA or RNA, single- or double-stranded), but this says surprisingly little about shared ancestry. For example, viruses with double-stranded DNA genomes seem to have arisen on at least four separate occasions.

    Source: ICTV

    The original ICTV viral classification, which is entirely separate from the tree of cellular life, included only the lower rungs of the evolutionary hierarchy, from species and genus up to the order level — a tier equivalent to primates or trees with cones in the classification of multicellular life. There were no higher levels. And many viral families floated alone, with no links to other kinds of virus. So in 2018, the ICTV added higher-order levels: classes, phyla and kingdoms8.At the very top, it invented ‘realms’, intended as counterparts to the ‘domains’ of cellular life — Bacteria, Archaea and Eukaryota — but using a different word to differentiate between the two trees. (Several years ago, some scientists suggested that certain viruses might fit into the cell-based evolutionary tree, but that idea has not gained widespread favour.)The ICTV outlined the branches of the tree, and grouped RNA-based viruses into a realm called Riboviria. SARS-CoV-2 and other coronaviruses, which have single-stranded RNA genomes, are part of this realm. But then it was up to the broader community of virologists to propose further taxonomic groups. As it happened, Eugene Koonin, an evolutionary biologist at the National Center for Biotechnology Information in Bethesda, Maryland, had assembled a team to analyse all the viral genomes, as well as the latest research on viral proteins, to create a first-draft taxonomy9.They reorganized Riboviria and proposed three more realms (see ‘Virus realms’). There was some quibbling over the details, Koonin says, but the taxonomy was ratified without much trouble by ICTV members in 2020. Two further realms got the green light in 2021, but the original four realms will probably remain the largest, he says. Eventually, Koonin speculates, the realms might number up to 25.

    Source: ICTV (talk.ictvonline.org/taxonomy); ICTV Coronaviridae Study Group. Nature Microbiol. 5, 536–544 (2020)

    That number supports many scientists’ suspicion that there’s no one common ancestor for virus-kind. “There is no single root for all viruses,” says Koonin. “It simply does not exist.” That means that viruses probably arose several times in the history of life on Earth — and there’s no reason to think such emergence can’t happen again. “The de novo origin of new viruses, it’s still ongoing,” says Mart Krupovic, a virologist at the Pasteur Institute in Paris who was involved in both the ICTV decisions and Koonin’s taxonomy team.As to how the realms arose, virologists have several ideas. Perhaps they descended from independent genetic elements at the dawn of life on Earth, before cells even took shape. Maybe they escaped or ‘devolved’ from whole cells, ditching most of the cellular machinery for a minimal lifestyle. Koonin and Krupovic favour a hybrid hypothesis in which those primordial genetic elements stole genes from cellular life to build their virus particles. Because there are multiple origins for viruses, it’s possible there are multiple ways they’ve originated, says Kuhn, who also served on the ICTV committee and worked on the new taxonomy proposal.Thus, although the viral and cellular trees of life are distinct, the branches touch, and genes pass between the two. Whether viruses count as being ‘alive’ depends on your personal definition of life. Many researchers do not consider them to be living things, but others disagree. “I tend to believe that they are living,” says Hiroyuki Ogata, a bioinformatician working on viruses at Kyoto University in Japan. “They are evolving, they have genetic material composed of DNA and RNA, and they are very important in the evolution of all life.”The current classification is widely recognized as just the first attempt, and some virologists say it’s a bit of a mess. A score of families still lack links to any realm. “The good point is, we are trying to put some order in that mess,” says Martinez-Garcia.World changers With the total mass of viruses on Earth equivalent to that of 75 million blue whales, scientists are certain they make a difference to food webs, ecosystems and even the planet’s atmosphere. The accelerating discovery of new viruses “has revealed a watershed of new ways viruses directly impact ecosystems”, says Matthew Sullivan, an environmental virologist at Ohio State University in Columbus. But scientists are still struggling to quantify how much of an impact they have.“We don’t have a very simple story around here at the moment,” says Ogata. In the ocean, viruses can burst out of their microbial hosts, releasing carbon to be recycled by others that eat the host’s innards and then produce carbon dioxide. But, more recently, scientists have also come to appreciate that popped cells often clump together and sink to the bottom of the ocean, sequestering carbon away from the atmosphere.

    Viral genomes collected from thawing permafrost at Stordalen Mire in Sweden have genes that could help break down and release carbon.Credit: Bob Gibbons/Alamy

    On land, thawing permafrost is a major source of carbon, says Sullivan, and viruses seem to be instrumental in carbon release from microbes in that environment. In 2018, he and his colleagues described 1,907 viral genomes and fragments collected from thawing permafrost in Sweden, including genes for proteins that might influence how carbon compounds break down and, potentially, become greenhouse gases10.Viruses can also influence other organisms by stirring up their genomes. For example, when viruses transfer antibiotic-resistance genes from one bacterium to another, drug-resistant strains can take over. Over time, this kind of transfer can create major evolutionary shifts in a population, says Camarillo-Guerrero. And not just in bacteria — an estimated 8% of human DNA is of viral origin. For example, our mammalian ancestors acquired a gene essential for placental development from a virus.For many questions about viral lifestyles, scientists will need more than just genomes. They will need to find the virus’s hosts. A virus itself might carry clues: it could be toting a recognizable bit of host genetic material in its own genome, for example.Martinez-Garcia and his colleagues used single-cell genomics to identify the microbes that contained the newly discovered 37-F6 virus. The host, too, is one of the most abundant and diverse organisms in the sea, a bacterium known as Pelagibacter11. In some waters, Pelagibacter makes up half the cells present. If just this one type of virus were to suddenly disappear, says Martinez-Garcia, ocean life would be thrown wildly off balance.To understand a virus’s full impact, scientists need to work out how it changes its host, says Alexandra Worden, an evolutionary ecologist at the GEOMAR Helmholtz Centre for Ocean Research in Kiel, Germany. She’s studying giant viruses that carry genes for light-harvesting proteins called rhodopsins. Theoretically, these genes could be useful to the hosts — for purposes such as energy transfer or signalling — but the sequences can’t confirm that. To find out what’s going on with these rhodopsin genes, Worden plans to culture the host and virus together, and study how the pair function in the combined, ‘virocell’ state. “Cell biology is the only way you can say what that true role is, how does this really affect the carbon cycle,” she says.Back in Florida, Breitbart hasn’t cultured her spider viruses, but she’s learnt some more about them. The pair of viruses belong to a category Breitbart calls mind-boggling for their tiny, circular genomes, encoding just one gene for their protein coat and one for their replication protein. One of the viruses is found only in the spider’s body, never its legs, so she thinks it’s actually infecting some creature the spider eats. The other virus is found throughout the spider’s body, and in its eggs and spiderlings, so she thinks it’s transmitted from parent to offspring12. It doesn’t seem to be doing them any harm, as far as Breitbart can tell.With viruses, “finding them’s actually the easy part”, she says. Picking apart how viruses influence host life cycles and ecology is much trickier. But first, virologists must answer one of the toughest questions of all, Breitbart says: “How do you pick which one to study?”

    Nature 595, 22-25 (2021)
    doi: https://doi.org/10.1038/d41586-021-01749-7

    References1.Rosario, K. et al. PeerJ 6, e5761 (2018).PubMed 
    Article 

    Google Scholar 
    2.Schulz, F. et al. Science 356, 85–85 (2017).PubMed 
    Article 

    Google Scholar 
    3.Schulz, F. et al. Nature 578, 432–436 (2020).PubMed 
    Article 

    Google Scholar 
    4.Camarillo-Guerrero, L. F., Almeida, A., Rangel-Pineros, G., Finn, R. D. & Lawley, T. D. Cell 184, 1098–1109.e9 (2021).PubMed 
    Article 

    Google Scholar 
    5.Callanan, J. et al. Sci. Adv. 6, eaay591 (2020).Article 

    Google Scholar 
    6.Martinez-Hernandez, F. et al. Nature Commun. 8, 15892 (2017).PubMed 
    Article 

    Google Scholar 
    7.Simmonds, P. et al. Nature Rev. Microbiol. 15, 161–168 (2017).PubMed 
    Article 

    Google Scholar 
    8.International Committee on Taxonomy of Viruses Executive Committee. Nature Microbiol. 5, 668–674 (2020).PubMed 
    Article 

    Google Scholar 
    9.Koonin, E. V. et al. Microbiol. Mol. Biol. Rev. 84, e00061-19 (2020).PubMed 
    Article 

    Google Scholar 
    10.Emerson, J. B. et al. Nature Microbiol. 3, 870–770 (2018).PubMed 
    Article 

    Google Scholar 
    11.Martinez-Hernandez, F. et al. ISME J. 13, 232–236 (2019).PubMed 
    Article 

    Google Scholar 
    12.Rosario, K., Mettel, K. A., Greco, A. M. & Breitbart, M. J. Gen. Virol. 100, 1253–1265 (2019).PubMed 
    Article 

    Google Scholar 
    Download references

    Related Articles

    Profile of a killer: the complex biology powering the coronavirus pandemic

    Should virus-naming rules change during a pandemic? The question divides virologists

    The new scope of virus taxonomy: partitioning the virosphere into 15 hierarchical ranks

    Subjects

    Virology

    Evolution

    Ecology

    Latest on:

    Virology

    Age-related immune response heterogeneity to SARS-CoV-2 vaccine BNT162b2
    Article 30 JUN 21

    Untangling introductions and persistence in COVID-19 resurgence in Europe
    Article 30 JUN 21

    SARS-CoV-2 mRNA vaccines induce persistent human germinal centre responses
    Article 28 JUN 21

    Evolution

    Untangling introductions and persistence in COVID-19 resurgence in Europe
    Article 30 JUN 21

    From the archive
    News & Views 29 JUN 21

    Mysterious skull fossils expand human family tree — but questions remain
    News 25 JUN 21

    Ecology

    China’s economic approach to protecting its ecology
    Spotlight 29 JUN 21

    Red light, green light: both signal ‘go’ to deadly algae
    Research Highlight 24 JUN 21

    Limited potential for bird migration to disperse plants to cooler latitudes
    Article 23 JUN 21

    Jobs from Nature Careers

    All jobs

    Advanced Research Assistant
    Wellcome Trust Sanger Institute
    Cambridge, United Kingdom

    JOB POST

    Assistant or Associate Professor of Molecular Pharmacology and Therapeutics
    Columbia University Medical Center (CUMC), CU
    New York, United States

    JOB POST

    154308 Tenure Track Assistant Professor or Associate Professor at the Department of Biomedical Sciences
    University of Copenhagen (UCPH)
    Copenhagen, Denmark

    JOB POST

    Research Associate in Synthetic Genomics
    The University of Manchester
    Manchester, United Kingdom

    JOB POST

    Nature Briefing
    An essential round-up of science news, opinion and analysis, delivered to your inbox every weekday.

    Email address

    Yes! Sign me up to receive the daily Nature Briefing email. I agree my information will be processed in accordance with the Nature and Springer Nature Limited Privacy Policy.

    Sign up More

  • in

    Dinosaur biodiversity declined well before the asteroid impact, influenced by ecological and environmental pressures

    1.Weishampel, D. B., Dodson, P. & Osmólska, H. The Dinosauria 2nd edn (University of California Press, 2004).2.Fastovsky, D. E. & Weishampel, D. B. The Evolution and Extinction of the Dinosaurs (Cambridge University Press, 2005).3.Brusatte, S. L. et al. The extinction of the dinosaurs. Biol. Rev. 90, 628–642 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    4.Alvarez, L. W., Alvarez, W., Asaro, F. & Michel, H. V. Extraterrestrial cause for the Cretaceous-Tertiary extinction. Science 208, 1095–1108 (1980).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    5.Chiarenza, A. A. et al. Asteroid impact, not volcanism, caused the end-Cretaceous dinosaur extinction. Proc. Natl Acad. Sci. USA 117, 17084–17093 (2020).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    6.Schulte, P. et al. The Chicxulub asteroid impact and mass extinction at the Cretaceous-Paleogene boundary. Science 327, 1214–1218 (2010).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    7.Russell, D. A. The gradual decline of the dinosaurs—fact or fallacy? Nature 307, 360–361 (1984).ADS 
    Article 

    Google Scholar 
    8.Sloan, R. E., Rigby, J. K., Van Valen, L. M. & Gabriel, D. Gradual dinosaur extinction and simultaneous ungulate radiation in the Hell Creek Formation. Science 232, 629–633 (1986).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    9.Sheehan, P. M., Fastovsky, D. E., Hoffmann, R. G., Berghaus, C. B. & Gabriel, D. L. Sudden extinction of the dinosaurs: Latest Cretaceous, upper Great Plains, USA. Science 254, 835–839 (1991).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    10.Sakamoto, M., Benton, M. J. & Venditti, C. Dinosaurs in decline tens of millions of years before their final extinction. Proc. Natl Acad. Sci. USA 113, 5036–5040 (2016).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    11.Chiarenza, A. A. et al. Ecological niche modelling does not support climatically-driven dinosaur diversity decline before the Cretaceous/Paleogene mass extinction. Nat. Commun. 10, 1091 (2019).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    12.Russell, L. S. Body temperature of dinosaurs and its relationships to their extinction. J. Paleontol. 39, 497–501 (1965).
    Google Scholar 
    13.Brusatte, S. L., Butler, R. J., Prieto-Márquez, A. & Norell, M. A. Dinosaur morphological diversity and the end-Cretaceous extinction. Nat. Commun. 3, 804 (2012).ADS 
    PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    14.Benson, R. B. J. et al. Rates of dinosaur body mass evolution indicate 170 million years of sustained ecological innovation on the avian stem lineage. PLoS Biol. 12, e1001853 (2014).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    15.Rezende, E. L., Bacigalupe, L. D., Nespolo, R. F. & Bozinovic, F. Shrinking dinosaurs and the evolution of endothermy in birds. Sci. Adv. 6, eaaw4486 (2020).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    16.Lloyd, G. T. et al. Dinosaurs and the Cretaceous Terrestrial Revolution. Proc. R. Soc. B Biol. Sci. 275, 2483–2490 (2008).Article 

    Google Scholar 
    17.Gates, T. A., Prieto-Márquez, A. & Zanno, L. E. Mountain building triggered Late Cretaceous North American megaherbivore dinosaur radiation. PLoS ONE 7, e42135 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    18.Loewen, M. A., Irmis, R. B., Sertich, J. J. W., Currie, P. J. & Sampson, S. D. Tyrant dinosaur evolution tracks the rise and fall of late Cretaceous oceans. PLoS ONE 8, e79420 (2013).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    19.Archibald, J. D. et al. Cretaceous extinctions: Multiple causes. Science 328, 973 (2010).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    20.Mitchell, J. S., Roopnarine, P. D. & Angielczyk, K. D. Late Cretaceous restructuring of terrestrial communities facilitated the end-Cretaceous mass extinction in North America. Proc. Natl Acad. Sci. USA 109, 18857–18861 (2012).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Schoene, B. et al. U-Pb constraints on pulsed eruption of the Deccan Traps across the end-Cretaceous mass extinction. Science 363, 862–866 (2019).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    22.Sprain, C. J. et al. The eruptive tempo of Deccan volcanism in relation to the Cretaceous-Paleogene boundary. Science 363, 866–870 (2019).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    23.Hull, P. M. et al. On impact and volcanism across the Cretaceous-Paleogene boundary. Science 367, 266–272 (2020).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    24.Landman, N. H. et al. Ammonite extinction and nautilid survival at the end of the Cretaceous. Geology 42, 707–710 (2014).ADS 
    CAS 
    Article 

    Google Scholar 
    25.Longrich, N. R., Martill, D. M. & Andres, B. Late Maastrichtian pterosaurs from North Africa and mass extinction of Pterosauria at the Cretaceous-Paleogene boundary. PLoS Biol. 16, e2001663 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    26.Longrich, N. R., Tokaryk, T. & Field, D. J. Mass extinction of birds at the Cretaceous-Paleogene (K-Pg) boundary. Proc. Natl Acad. Sci. USA 108, 15253–15257 (2011).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    27.Longrich, N. R., Bhullar, B.-A. S. & Gauthier, J. A. Mass extinction of lizards and snakes at the Cretaceous-Paleogene boundary. Proc. Natl Acad. Sci. USA 109, 21396–21401 (2012).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    28.Fastovsky, D. E. et al. Shape of Mesozoic dinosaur richness. Geology 32, 877–880 (2004).ADS 
    Article 

    Google Scholar 
    29.Archibald, J. D. in Volcanism, Impacts, and Mass Extinctions: Causes and Effects (eds. Keller, G. & Kerr, A. C.) 213–224 (The Geological Society of America Special Paper 505, 2014).30.Wang, S. C. & Dodson, P. Estimating the diversity of dinosaurs. Proc. Natl Acad. Sci. USA 103, 13601–13605 (2006).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    31.Starrfelt, J. & Liow, L. H. How many dinosaur species were there? Fossil bias and true richness estimated using a Poisson sampling model. Philos. Trans. R. Soc. B Biol. Sci. 371, 20150219 (2016).Article 
    CAS 

    Google Scholar 
    32.Bonsor, J. A., Barrett, P. M., Raven, T. J. & Cooper, N. Dinosaur diversification rates were not in decline prior to the K-Pg boundary. R. Soc. Open Sci. 7, 201195 (2020).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    33.Benton, M. J., Wills, M. A. & Hitchin, R. Quality of the fossil record through time. Nature 403, 534–537 (2000).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    34.Alroy, J. et al. Effects of sampling standardization on estimates of Phanerozoic marine diversification. Proc. Natl Acad. Sci. USA 98, 6261–6266 (2001).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    35.Alroy, J. et al. Phanerozoic trends in the global diversity of marine invertebrates. Science 321, 97–100 (2008).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    36.Close, R. A., Evers, S. W., Alroy, J. & Butler, R. J. How should we estimate diversity in the fossil record? Testing richness estimators using sampling-standardised discovery curves. Methods Ecol. Evol. 9, 1386–1400 (2018).Article 

    Google Scholar 
    37.Silvestro, D., Salamin, N., Antonelli, A. & Meyer, X. Improved estimation of macroevolutionary rates from fossil data using a Bayesian framework. Paleobiology 45, 546–570 (2019).Article 

    Google Scholar 
    38.Close, R. A., Benson, R. B. J., Saupe, E. E., Clapham, M. E. & Butler, R. J. The spatial structure of Phanerozoic marine animal diversity. Science 368, 420–424 (2020).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    39.Benton, M. J. Scientific methodologies in collision: The history of the study of the extinction of the dinosaurs. Evol. Biol. 24, 371–400 (1990).
    Google Scholar 
    40.Butler, R. J., Benson, R. B. J., Carrano, M. T., Mannion, P. D. & Upchurch, P. Sea level, dinosaur diversity and sampling biases: Investigating the ‘common cause’ hypothesis in the terrestrial realm. Proc. R. Soc. B Biol. Sci. 278, 1165–1170 (2011).Article 

    Google Scholar 
    41.Zaffos, A., Finnegan, S. & Peters, S. E. Plate tectonic regulation of global marine animal diversity. Proc. Natl Acad. Sci. USA 114, 5653–5658 (2017).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.East, M., Müller, R. D., Williams, S., Zahirovic, S. & Heine, C. Subduction history reveals Cretaceous slab superflux as a possible cause for the mid-Cretaceous plume pulse and superswell events. Gondwana Res. 79, 125–139 (2020).ADS 
    Article 

    Google Scholar 
    43.Grasby, S. E., Them, T. R., Chen, Z., Yin, R. & Ardakani, O. H. Mercury as a proxy for volcanic emissions in the geologic record. Earth Sci. Rev. 196, 102880 (2019).CAS 
    Article 

    Google Scholar 
    44.Miller, K. G. et al. The Phanerozoic record of global sea level change. Science 310, 1293–1298 (2005).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    45.Ray, D. C. et al. The magnitude and cause of short-term eustatic Cretaceous sea-level change: a synthesis. Earth Sci. Rev. 197, 102901 (2019).Article 

    Google Scholar 
    46.Coiffard, C., Gomez, B., Daviero-Gomez, V. & Dilcher, D. L. Rise to dominance of angiosperm pioneers in European Cretaceous environments. Proc. Natl Acad. Sci. USA 109, 20955–20959 (2012).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    47.Chaboureau, A.-C., Sepulchre, P., Donnadieu, Y. & Franc, A. Tectonic-driven climate change and the diversification of angiosperms. Proc. Natl Acad. Sci. USA 111, 14066–14070 (2014).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    48.Magallón, S., Gómez-Acevedo, S., Sánchez-Reyes, L. L. & Hernández-Hernández, T. A metacalibrated time-tree documents the early rise of flowering plant phylogenetic diversity. N. Phytol. 207, 437–453 (2015).Article 

    Google Scholar 
    49.Magallón, S., Sánchez-Reyes, L. L. & Gómez-Acevedo, S. L. Thirty clues to the exceptional diversification of flowering plants. Ann. Bot. 123, 491–503 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    50.Meredith, R. W. et al. Impacts of the Cretaceous terrestrial revolution and KPg extinction on mammal diversification. Science 334, 521–524 (2011).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    51.Grossnickle, D. M. & Newham, E. Therian mammals experience an ecomorphological radiation during the Late Cretaceous and selective extinction at the K–Pg boundary. Proc. R. Soc. B Biol. Sci. 283, 20160256 (2016).Article 

    Google Scholar 
    52.Liu, L. et al. Genomic evidence reveals a radiation of placental mammals uninterrupted by the KPg boundary. Proc. Natl Acad. Sci. USA 114, E7282–E7290 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    53.Arbour, V. M., Zanno, L. E. & Gates, T. A. Ankylosaurian dinosaur palaeoenvironmental associations were influenced by extirpation, sea-level fluctuation, and geodispersal. Palaeogeogr. Palaeoclimatol. Palaeoecol. 449, 289–299 (2016).Article 

    Google Scholar 
    54.Tennant, J. P., Mannion, P. D. & Upchurch, P. Sea level regulated tetrapod diversity dynamics through the Jurassic/Cretaceous interval. Nat. Commun. 7, 12737 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    55.Silvestro, D., Schnitzler, J., Liow, L. H., Antonelli, A. & Salamin, N. Bayesian estimation of speciation and extinction from incomplete fossil occurrence data. Syst. Biol. 63, 349–367 (2014).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    56.Silvestro, D., Antonelli, A., Salamin, N. & Quental, T. B. The role of clade competition in the diversification of North American canids. Proc. Natl Acad. Sci. USA 112, 8684–8689 (2015).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    57.Lehtonen, S. et al. Environmentally driven extinction and opportunistic origination explain fern diversification patterns. Sci. Rep. 7, 4831 (2017).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    58.Condamine, F. L., Romieu, J. & Guinot, G. Climate cooling and clade competition likely drove the decline of lamniform sharks. Proc. Natl Acad. Sci. USA 116, 20584–20590 (2019).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    59.Signor, P. W. & Lipps, J. H. in Geological Implications of Impacts of Large Asteroids and Comets on The Earth (eds. Silver, L. T. & Schultz, P. H.) vol. 190, 291–296 (Geological Society of America Special Publication, 1982).60.Benson, R. B. J. Dinosaur macroevolution and macroecology. Annu. Rev. Ecol. Evol. Syst. 49, 379–408 (2018).Article 

    Google Scholar 
    61.Dean, C. D., Chiarenza, A. A. & Maidment, S. C. R. Formation binning: a new method for increased temporal resolution in regional studies, applied to the Late Cretaceous dinosaur fossil record of North America. Palaeontology 63, 881–901 (2020).Article 

    Google Scholar 
    62.Moen, D. & Morlon, H. Why does diversification slow down? Trends Ecol. Evol. 29, 190–197 (2014).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    63.Condamine, F. L., Rolland, J. & Morlon, H. Assessing the causes of diversification slowdowns: Temperature-dependent and diversity-dependent models receive equivalent support. Ecol. Lett. 22, 1900–1912 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    64.Prieto-Márquez, A., Dalla Vecchia, F. M., Gaete, R. & Galobart, À. Diversity, relationships, and biogeography of the lambeosaurine dinosaurs from the European archipelago, with description of the new aralosaurin Canardia garonnensis. PLoS ONE 8, e69835 (2013).65.Prieto-Márquez, A., Fondevilla, V., Sellés, A. G., Wagner, J. R. & Galobart, À. Adynomosaurus arcanus, a new lambeosaurine dinosaur from the Late Cretaceous Ibero-Armorican Island of the European archipelago. Cretac. Res. 96, 19–37 (2019).Article 

    Google Scholar 
    66.Longrich, N. R., Suberbiola, X. P., Pyron, R. A. & Jalil, N.-E. The first duckbill dinosaur (Hadrosauridae: Lambeosaurinae) from Africa and the role of oceanic dispersal in dinosaur biogeography. Cretac. Res. 120, 104678 (2021).Article 

    Google Scholar 
    67.Kobayashi, Y., Takasaki, R., Kubota, K. & Fiorillo, A. R. A new basal hadrosaurid (Dinosauria: Ornithischia) from the latest Cretaceous Kita-ama Formation in Japan implies the origin of hadrosaurids. Sci. Rep. 11, 8547 (2021).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    68.Stubbs, T. L., Benton, M. J., Elsler, A. & Prieto-Márquez, A. Morphological innovation and the evolution of hadrosaurid dinosaurs. Paleobiology 45, 347–362 (2019).Article 

    Google Scholar 
    69.Reest, A. J. van der & Currie, P. J. Troodontids (Theropoda) from the Dinosaur Park Formation, Alberta, with a description of a unique new taxon: Implications for deinonychosaur diversity in North America. Can. J. Earth Sci. 54, 919–935 (2017).70.Hartman, S. et al. A new paravian dinosaur from the Late Jurassic of North America supports a late acquisition of avian flight. PeerJ 7, e7247 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    71.Horner, J. R., Varricchio, D. J. & Goodwin, M. B. Marine transgressions and the evolution of Cretaceous dinosaurs. Nature 358, 59–61 (1992).ADS 
    Article 

    Google Scholar 
    72.O’Brien, C. L. et al. Cretaceous sea-surface temperature evolution: Constraints from TEX86 and planktonic foraminiferal oxygen isotopes. Earth Sci. Rev. 172, 224–247 (2017).ADS 
    Article 
    CAS 

    Google Scholar 
    73.Huber, B. T., MacLeod, K. G., Watkins, D. K. & Coffin, M. F. The rise and fall of the Cretaceous Hot Greenhouse climate. Glob. Planet. Change 167, 1–23 (2018).ADS 
    Article 

    Google Scholar 
    74.Mannion, P. D. et al. A temperate palaeodiversity peak in Mesozoic dinosaurs and evidence for Late Cretaceous geographical partitioning. Glob. Ecol. Biogeogr. 21, 898–908 (2012).Article 

    Google Scholar 
    75.Forster, A., Schouten, S., Baas, M. & Damsté, J. S. S. Mid-Cretaceous (Albian–Santonian) sea surface temperature record of the tropical Atlantic Ocean. Geology 35, 919–922 (2007).ADS 
    Article 

    Google Scholar 
    76.O’Connor, L. K. et al. Late Cretaceous temperature evolution of the southern high latitudes: a TEX86 perspective. Paleoceanogr. Paleoclimatol. 34, 436–454 (2019).ADS 
    Article 

    Google Scholar 
    77.Linnert, C. et al. Evidence for global cooling in the Late Cretaceous. Nat. Commun. 5, 4194 (2014).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    78.Crane, P. R. & Lidgard, S. Angiosperm diversification and paleolatitudinal gradients in Cretaceous floristic diversity. Science 246, 675–678 (1989).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    79.Condamine, F. L., Silvestro, D., Koppelhus, E. B. & Antonelli, A. The rise of angiosperms pushed conifers to decline during global cooling. Proc. Natl Acad. Sci. USA 117, 28867–28875 (2020).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    80.Condamine, F. L., Rolland, J. & Morlon, H. Macroevolutionary perspectives to environmental change. Ecol. Lett. 16, 72–85 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    81.Silvestro, D., Cascales-Miñana, B., Bacon, C. D. & Antonelli, A. Revisiting the origin and diversification of vascular plants through a comprehensive Bayesian analysis of the fossil record. N. Phytol. 207, 425–436 (2015).Article 

    Google Scholar 
    82.Prokoph, A., Shields, G. A. & Veizer, J. Compilation and time-series analysis of a marine carbonate δ18O, δ13C, 87Sr/86Sr and δ34S database through Earth history. Earth Sci. Rev. 87, 113–133 (2008).ADS 
    CAS 
    Article 

    Google Scholar 
    83.Miller, K. G. et al. The phanerozoic record of global sea-level change. Science 310, 1293–1298 (2005).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    84.Barrett, P. M. Paleobiology of herbivorous dinosaurs. Annu. Rev. Earth Planet. Sci. 42, 207–230 (2014).ADS 
    CAS 
    Article 

    Google Scholar 
    85.Grady, J. M., Enquist, B. J., Dettweiler-Robinson, E., Wright, N. A. & Smith, F. A. Evidence for mesothermy in dinosaurs. Science 344, 1268–1272 (2014).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    86.Eagle, R. A. et al. Isotopic ordering in eggshells reflects body temperatures and suggests differing thermophysiology in two Cretaceous dinosaurs. Nat. Commun. 6, 8296 (2015).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    87.Paladino, F. V., Dodson, P., Hammond, J. K. & Spotila, J. R. Temperature-dependent sex determination in dinosaurs? Implications for population dynamics and extinction. in Paleobiology of the Dinosaurs (ed. Farlow, J. O.) vol. 238, 63–70 (Geological Society of America Special Papers, 1989).88.Vavrek, M. J. & Larsson, H. C. E. Low beta diversity of Maastrichtian dinosaurs of North America. Proc. Natl Acad. Sci. USA 107, 8265–8268 (2010).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    89.Dunne, J. A., Williams, R. J. & Martinez, N. D. Network structure and biodiversity loss in food webs: Robustness increases with connectance. Ecol. Lett. 5, 558–567 (2002).Article 

    Google Scholar 
    90.Brodie, J. F. et al. Secondary extinctions of biodiversity. Trends Ecol. Evol. 29, 664–672 (2014).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    91.Fraser, D. et al. Investigating biotic interactions in deep time. Trends Ecol. Evol. 36, 61–75 (2021).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    92.Mallon, J. C. Competition structured a Late Cretaceous megaherbivorous dinosaur assemblage. Sci. Rep. 9, 15447 (2019).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    93.Benton, M. J. Progress and competition in macroevolution. Biol. Rev. 62, 305–338 (1987).Article 

    Google Scholar 
    94.Fricke, H. C. & Pearson, D. A. Stable isotope evidence for changes in dietary niche partitioning among hadrosaurian and ceratopsian dinosaurs of the Hell Creek Formation, North Dakota. Paleobiology 34, 534–552 (2008).Article 

    Google Scholar 
    95.Mallon, J. C. & Anderson, J. S. Skull ecomorphology of megaherbivorous dinosaurs from the Dinosaur Park Formation (Upper Campanian) of Alberta, Canada. PLoS ONE 8, e67182 (2013).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    96.Nordén, K. K., Stubbs, T. L., Prieto-Márquez, A. & Benton, M. J. Multifaceted disparity approach reveals dinosaur herbivory flourished before the end-Cretaceous mass extinction. Paleobiology 44, 620–637 (2018).Article 

    Google Scholar 
    97.Lyson, T. R. & Longrich, N. R. Spatial niche partitioning in dinosaurs from the latest Cretaceous (Maastrichtian) of North America. Proc. R. Soc. B Biol. Sci. 278, 1158–1164 (2011).Article 

    Google Scholar 
    98.Li, Z. et al. Ultramicrostructural reductions in teeth: Implications for dietary transition from non-avian dinosaurs to birds. BMC Evol. Biol. 20, 46 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    99.Cau, A. et al. Synchrotron scanning reveals amphibious ecomorphology in a new clade of bird-like dinosaurs. Nature 552, 395–399 (2017).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    100.Cau, A. The body plan of Halszkaraptor escuilliei (Dinosauria, Theropoda) is not a transitional form along the evolution of dromaeosaurid hypercarnivory. PeerJ 8, e8672 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    101.Fowler, D. W., Freedman, E. A., Scannella, J. B. & Kambic, R. E. The predatory ecology of Deinonychus and the origin of flapping in birds. PLoS ONE 6, e28964 (2011).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    102.Frederickson, J. A., Engel, M. H. & Cifelli, R. L. Ontogenetic dietary shifts in Deinonychus antirrhopus (Theropoda; Dromaeosauridae): Insights into the ecology and social behavior of raptorial dinosaurs through stable isotope analysis. Palaeogeogr. Palaeoclimatol. Palaeoecol. 552, 109780 (2020).Article 

    Google Scholar 
    103.O’Connor, J. et al. Microraptor with ingested lizard suggests non-specialized digestive function. Curr. Biol. 29, 2423–2429 (2019).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    104.King, J. L., Sipla, J. S., Georgi, J. A., Balanoff, A. M. & Neenan, J. M. The endocranium and trophic ecology of Velociraptor mongoliensis. J. Anat. 237, 861–869 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    105.Owocki, K., Kremer, B., Cotte, M. & Bocherens, H. Diet preferences and climate inferred from oxygen and carbon isotopes of tooth enamel of Tarbosaurus bataar (Nemegt Formation, Upper Cretaceous, Mongolia). Palaeogeogr. Palaeoclimatol. Palaeoecol. 537, 109190 (2020).Article 

    Google Scholar 
    106.Dalman, S. & Lucas, S. New evidence for cannibalism in tyrannosaurid dinosaurs from the Late Cretaceous of New Mexico. N. Mex. Mus. Nat. Hist. Sci. Bull. 82, 39–56 (2021).
    Google Scholar 
    107.Frederickson, J. A., Engel, M. H. & Cifelli, R. L. Niche partitioning in theropod dinosaurs: Diet and habitat preference in predators from the uppermost Cedar Mountain Formation (Utah, U.S.A.). Sci. Rep. 8, 17872 (2018).108.Hassler, A. et al. Calcium isotopes offer clues on resource partitioning among Cretaceous predatory dinosaurs. Proc. R. Soc. B Biol. Sci. 285, 20180197 (2018).109.Schroeder, K., Lyons, S. K. & Smith, F. A. The influence of juvenile dinosaurs on community structure and diversity. Science 371, 941–944 (2021).110.Currie, P. J., Badamgarav, D., Koppelhus, E. B., Sissons, R. & Vickaryous, M. K. Hands, feet, and behaviour in Pinacosaurus (Dinosauria: Ankylosauridae). Acta Palaeontol. Polon. 56, 489–504 (2011).Article 

    Google Scholar 
    111.Burns, M. E., Currie, P. J., Sissons, R. L. & Arbour, V. M. Juvenile specimens of Pinacosaurus grangeri Gilmore, 1933 (Ornithischia: Ankylosauria) from the Late Cretaceous of China, with comments on the specific taxonomy of Pinacosaurus. Cretac. Res. 32, 174–186 (2011).Article 

    Google Scholar 
    112.Burns, M. E., Tumanova, T. A. & Currie, P. J. Postcrania of juvenile Pinacosaurus grangeri (Ornithischia: Ankylosauria) from the Upper Cretaceous Alagteeg Formation, Alag Teeg, Mongolia: Implications for ontogenetic allometry in ankylosaurs. J. Paleontol. 89, 168–182 (2015).113.Botfalvai, G., Prondvai, E. & Ősi, A. Living alone or moving in herds? A holistic approach highlights complexity in the social lifestyle of Cretaceous ankylosaurs. Cretac. Res. 118, 104633 (2021).Article 

    Google Scholar 
    114.Arbour, V. M. & Zanno, L. E. The evolution of tail weaponization in amniotes. Proc. R. Soc. B Biol. Sci. 285, 20172299 (2018).Article 

    Google Scholar 
    115.Arbour, V. M. & Zanno, L. E. Tail weaponry in ankylosaurs and glyptodonts: An example of a rare but strongly convergent phenotype. Anat. Rec. 303, 988–998 (2020).Article 

    Google Scholar 
    116.Van Valen, L. A new evolutionary law. Evol. Theory 1, 1–30 (1973).117.Hagen, O., Andermann, T., Quental, T. B., Antonelli, A. & Silvestro, D. Estimating age-dependent extinction: Contrasting evidence from fossils and phylogenies. Syst. Biol. 67, 458–474 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    118.Finnegan, S., Payne, J. L. & Wang, S. C. The Red Queen revisited: Reevaluating the age selectivity of Phanerozoic marine genus extinctions. Paleobiology 34, 318–341 (2008).Article 

    Google Scholar 
    119.Doran, N. A., Arnold, A. J., Parker, W. C. & Huffer, F. W. Is extinction age dependent? PALAIOS 21, 571–579 (2006).ADS 
    Article 

    Google Scholar 
    120.Larson, D. W., Brown, C. M. & Evans, D. C. Dental disparity and ecological stability in bird-like dinosaurs prior to the end-Cretaceous mass extinction. Curr. Biol. 26, 1325–1333 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    121.Romano, M. Disparity versus diversity in ankylosaurid dinosaurs: Explored morphospace indicates two separate evolutive radiations. Rend. Online Soc. Geol. It. 53, 2–8 (2021).122.Turner, A. H., Montanari, S. & Norell, M. A. A new dromaeosaurid from the Late Cretaceous Khulsan locality of Mongolia. Am. Mus. Novitat. 2020, 1–48 (2021).123.Maryańska, T. & Osmólska, H. Pachycephalosauria, a new suborder of ornithischian dinosaurs. Palaeontol. Polon. 30, 45–102 (1974).
    Google Scholar 
    124.Sereno, P. C. National Geographic Research: Phylogeny of the bird-hipped dinosaurs (Order Ornithischia). Natl Geogr. Res. 2, 234–256 (1986). https://d3qi0qp55mx5f5.cloudfront.net/paulsereno/i/docs/86-NGRes-PhyloOrnithis_1.pdf?mtime=1591821557.125.Sullivan, R. M. A taxonomic review of the Pachycephalosauridae (Dinosauria: Ornithischia). N. Mex. Mus. Nat. Hist. Sci. Bull. 35, 347–365 (2006).
    Google Scholar 
    126.Lee, M. S. Y., Cau, A., Naish, D. & Dyke, G. J. Morphological clocks in paleontology, and a mid-cretaceous origin of crown aves. Syst. Biol. 63, 442–449 (2014).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    127.Arbour, V. M. & Evans, D. C. A new ankylosaurine dinosaur from the Judith River Formation of Montana, USA, based on an exceptional skeleton with soft tissue preservation. R. Soc. Open Sci. 4, 161086 (2017).128.McDonald, A. T., Wolfe, D. G. & Dooley, A. C. Jr A new tyrannosaurid (Dinosauria: Theropoda) from the Upper Cretaceous Menefee Formation of New Mexico. PeerJ 6, e5749 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    129.Longrich, N. R. & Field, D. J. Torosaurus is not Triceratops: Ontogeny in chasmosaurine ceratopsids as a case study in dinosaur taxonomy. PLoS ONE 7, e32623 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    130.Larson, P. L. in Tyrannosaurid Paleobiology (eds. Parrish, J. M., Molnar, R. A., Currie, P. J. & Koppelhus, E. B.) 15–54 (Indiana University Press, 2013).131.Yun, C. Evidence points out that ‘Nanotyrannus’ is a juvenile Tyrannosaurus rex. PeerJ 3, e1052 (2015).Article 

    Google Scholar 
    132.Brusatte, S. L. et al. Dentary groove morphology does not distinguish ‘Nanotyrannus’ as a valid taxon of tyrannosauroid dinosaur. Comment on: “Distribution of the dentary groove of theropod dinosaurs: Implications for theropod phylogeny and the validity of the genus Nanotyrannus Bakker et al., 1988. Cretac. Res. 65, 232–237 (2016).Article 

    Google Scholar 
    133.Schmerge, J. D. & Rothschild, B. M. When a groove is not a groove: Clarification of the appearance of the dentary groove in tyrannosauroid theropods and the distinction between Nanotyrannus and Tyrannosaurus. Reply to Comment on: “Distribution of the dentary groove of theropod dinosaurs: Implications for theropod phylogeny and the validity of the genus Nanotyrannus Bakker et al., 1988. Cretac. Res. 65, 238–243 (2016).Article 

    Google Scholar 
    134.Xu, X., Zhou, Z., Sullivan, C., Wang, Y. & Ren, D. An updated review of the Middle-Late Jurassic Yanliao biota: Chronology, taphonomy, paleontology and paleoecology. Acta Geol. Sin. 90, 2229–2243 (2016).Article 

    Google Scholar 
    135.Cau, A., Brougham, T. & Naish, D. The phylogenetic affinities of the bizarre Late Cretaceous Romanian theropod Balaur bondoc (Dinosauria, Maniraptora): Dromaeosaurid or flightless bird? PeerJ 3, e1032 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    136.Agnolin, F. L. & Motta, M. J. Paravian phylogeny and the dinosaur-bird transition: An overview. Front. Earth Sci. 6, 252 (2019).ADS 
    Article 

    Google Scholar 
    137.Pei, R. et al. Potential for powered flight neared by most close avialan relatives, but few crossed Its thresholds. Curr. Biol. 30, 4033–4046 (2020).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    138.Foth, C. & Rauhut, O. W. M. Re-evaluation of the Haarlem Archaeopteryx and the radiation of maniraptoran theropod dinosaurs. BMC Evol. Biol. 17, 236 (2017).139.Rauhut, O. W., Tischlinger, H. & Foth, C. A non-archaeopterygid avialan theropod from the Late Jurassic of southern Germany. eLife 8, e43789 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    140.Lefèvre, U. et al. A new Jurassic theropod from China documents a transitional step in the macrostructure of feathers. Sci. Nat. 104, 74 (2017).Article 
    CAS 

    Google Scholar 
    141.Shen, C. et al. A new troodontid dinosaur from the Lower Cretaceous Yixian formation of Liaoning province. China Acta Geol. Sin. 91, 763–780 (2017).Article 

    Google Scholar 
    142.Arbour, V. M. & Currie, P. J. Euoplocephalus tutus and the diversity of ankylosaurid dinosaurs in the Late Cretaceous of Alberta, Canada, and Montana, USA. PLoS ONE 8, e62421 (2013).143.Arbour, V. M. & Currie, P. J. Systematics, phylogeny and palaeobiogeography of the ankylosaurid dinosaurs. J. Syst. Palaeontol. 14, 385–444 (2016).Article 

    Google Scholar 
    144.Arbour, V. M., Currie, P. J. & Badamgarav, D. The ankylosaurid dinosaurs of the Upper Cretaceous Baruungoyot and Nemegt formations of Mongolia. Zool. J. Linn. Soc. 172, 631–652 (2014).
    Google Scholar 
    145.Arbour, V. M. et al. A new ankylosaurid dinosaur from the Upper Cretaceous (Kirtlandian) of New Mexico with implications for ankylosaurid diversity in the Upper Cretaceous of Western North America. PLoS ONE 9, e108804 (2014).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    146.Gradstein, F. M., Ogg, J. G., Schmitz, M. D. & Ogg, G. M. The Geologic Time Scale 2012 (Elsevier B.V., 2012).147.Brown, C. M. & Henderson, D. M. A new horned dinosaur reveals convergent evolution in cranial ornamentation in Ceratopsidae. Curr. Biol. 25, 1641–1648 (2015).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    148.Jerzykiewicz, T., Currie, P. J., Fanti, F. & Lefeld, J. Lithobiotopes of the Nemegt Gobi Basin. Can. J. Earth Sci. https://doi.org/10.1139/cjes-2020-0148 (2021).149.Silvestro, D., Salamin, N. & Schnitzler, J. PyRate: A new program to estimate speciation and extinction rates from incomplete fossil data. Methods Ecol. Evol. 5, 1126–1131 (2014).Article 

    Google Scholar 
    150.Rambaut, A. R., Drummond, A. J., Xie, D., Baele, G. & Suchard, M. A. Posterior summarization in Bayesian phylogenetics using Tracer 1.7. Syst. Biol. 67, 901–904 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    151.Brusatte, S. L. et al. Tyrannosaur paleobiology: New research on ancient exemplar organisms. Science 329, 1481–1485 (2010).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    152.Ryan, M. J., Chinnery-Allgeier, B. J. & Eberth, D. A. New Perspectives on Horned Dinosaurs (Indiana University Press, 2010).153.Xu, X., Wang, K., Zhao, X. & Li, D. First ceratopsid dinosaur from China and its biogeographical implications. Chin. Sci. Bull. 55, 1631–1635 (2010).CAS 
    Article 

    Google Scholar 
    154.Hannisdal, B. & Peters, S. E. Phanerozoic Earth system evolution and marine biodiversity. Science 334, 1121–1124 (2011).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    155.Liow, L. H., Reitan, T. & Harnik, P. G. Ecological interactions on macroevolutionary time scales: Clams and brachiopods are more than ships that pass in the night. Ecol. Lett. 18, 1030–1039 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    156.Erwin, D. H. Climate as a driver of evolutionary change. Curr. Biol. 19, R575–R583 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    157.Mayhew, P. J., Bell, M. A., Benton, T. G. & McGowan, A. J. Biodiversity tracks temperature over time. Proc. Natl Acad. Sci. USA 109, 15141–15145 (2012).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    158.Zachos, J., Pagani, M., Sloan, L., Thomas, E. & Billups, K. Trends, rythms, and aberration in global climate 65 Ma to present. Science 292, 686–693 (2001).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    159.Zachos, J. C., Dickens, G. R. & Zeebe, R. E. An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature 451, 279–283 (2008).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    160.Cramer, B. S., Toggweiler, J. R., Wright, J. D., Katz, M. E. & Miller, K. G. Ocean overturning since the late cretaceous: Inferences from a new benthic foraminiferal isotope compilation. Paleoceanography 24, 1–14 (2009).Article 

    Google Scholar 
    161.Barba-Montoya, J., Reis, M., Schneider, H., Donoghue, P. C. J. & Yang, Z. Constraining uncertainty in the timescale of angiosperm evolution and the veracity of a Cretaceous Terrestrial Revolution. N. Phytol. 218, 819–834 (2018).Article 

    Google Scholar 
    162.Zhang, M., Dai, S., Du, B., Ji, L. & Hu, S. Mid-Cretaceous hothouse climate and the expansion of early angiosperms. Acta Geol. Sin. 92, 2004–2025 (2018).Article 

    Google Scholar  More

  • in

    Handling of spurious sequences affects the outcome of high-throughput 16S rRNA gene amplicon profiling

    Filtering threshold for handling spurious sequencesWe first used bacterial communities of known composition (simplified communities) to assess the occurrence of spurious taxa and to determine at which relative abundances they begin to appear. To propose a cutoff that is potentially applicable to different 16S rRNA gene amplicon studies, we included reference data obtained with different variable regions and sequencing pipelines and originating from both in vitro an in vivo communities varying in number and type of species (max. 58) (Tables 1 and 2). To determine a filtering threshold that allowed exclusion of most spurious taxa, we recorded the relative abundance of the first spurious OTU occurring in each of the reference community datasets (Fig. 2a). Median values of approx. 0.12% relative abundance were observed (Fig. 2b). Besides one outlier in the mock communities (0.44% relative abundance), all values were below 0.25% relative abundance.Fig. 2: Determination of filtering thresholds using artificial communities of known composition in vitro (mock; n = 9 different types; 21 replicates in total) and in mice (gnotobiotes; n = 4 different communities; 28 mice in total).a Example of the occurrence of all molecular species detected without filtering in the gut of a gnotobiotic mouse [49]. The arrow indicates the position of the first spurious molecular species, all following taxa being considered as having a high risk of being spurious (light gray bars in the enlarged inset). b Distribution of the relative abundances of first occurring spurious molecular species (as shown in panel a) across all mock communities and samples from gnotobiotes. The orange dashes on the y-axis indicate the consensus threshold of 0.25% relative abundance, above which no spurious taxa occurred with the exception of one outlier in a mock community at a relative abundance of 0.44%. c Comparison of various standard filtering cutoffs (see explanations in the text) in terms of spurious taxa (i.e., those molecular species not matching sequences of the known species contained in the artificial communities). d Corresponding percentages of positive hits retained by the different filtering strategies, with positive hits being defined as the reference sequences found in the respective amplicon datasets. e Percentage of spurious taxa and positive hits in the same reference communities using the DADA2 pipeline for analysis based on amplicon sequence variants (ASVs) [6]. f Effect of filtering thresholds at increments of 0.05% relative abundance on the detection of spurious taxa and positive hits in all mock and gnotobiotic datasets for OTUs (upper panel) and ASVs (lower panel). Lines correspond to mean values; ribbons represent standard deviations.Full size imageWithout any filtering, sequence clustering generated an average of 508 ± 355 OTUs (min. 52; max. 1081) per mock community (10–58 target species in theory) and 105 ± 50 OTUs (min, 55; max. 215) per gnotobiotic community (4–12 target species in theory). Up to 87% of these OTUs were spurious (i.e., they did not match the expected classification of species contained in the corresponding artificial community) (Fig. 2c). On average, the proportion of spurious OTUs in both the mock communities and samples from gnotobiotic mice was slightly lower after removing singletons, although this did not reach statistical significance (50.8 vs. 64.3%, p = 0.227; 57.5% vs. 65.7%; p = 0.70, pairwise comparison by t-test, including Benjamini–Hochberg correction following ANOVA). Interestingly, the proportion of spurious molecular species was higher in gnotobiotic mice independent of filtering (p  0.50) (Fig. 2d). Note that the diversity of reference communities in the gnotobiotic mice was relatively low (4–12 members; Table 2), resulting in a marked drop in the percentage of positive hit (8–25%) when even just one true member is excluded after filtering because of its low relative abundance (which is an expectable event considering a classical, exponentially decreasing distribution of species occurrence in gut environments).We next employed the widely used ASV analysis approach to confirm the aforementioned results. Processing of the same simplified communities generated a total number of 42 ± 25 ASVs (min. 16; max. 98) for mock communities (10–58 target species) and 14 ± 8 ASVs (min. 4; max. 25) for gnotobiotes (4–12 target species). Altogether, a marked decrease in spurious taxa was observed compared with OTU clustering, with an average of 8.6 ± 11.8 and 4.4 ± 6.4% spurious sequences for mock and gnotobiotic communities, respectively (comparison of purple box plots in Fig. 2e, top panels, and Fig. 2c). Of note, the DADA2 pipeline used for the ASV approach does not infer sequence variants that are only supported by a single read (singletons) due to a lack of confidence in their existence relative to sequencing errors. Consequently, data corresponding to “no filtering” with the OTU-based approach were not generated. On average, the first spurious ASV occurred at a relative abundance of 0.10 ± 0.32%. By applying the cutoff of 0.25% relative abundance, spurious sequences were completely removed (except for three outlying samples), albeit with a slight drop in positive hits for both mock and gnotobiotic communities (Fig. 2e).To obtain a more comprehensive view on how filtering thresholds affect the detection of spurious taxa, all datasets (mock and gnotobiotic mice) were processed using a range of relative abundance filtering thresholds (from 0 to 0.5% at increments of 0.05%) after either OTU- or ASV-based processing of raw sequence reads (Fig. 2f). These data indicate that filtering thresholds between 0.1 and 0.3% are appropriate to reduce the occurrence of spurious taxa to 600 of the 678 spurious OTUs occurred in fewer than five of the ten sequencing runs tested, with approximately 450 of them occurring in only one run (Fig. 3c). This observation indicates that the majority of spurious taxa are sporadic cross-contaminations rather than generalist artifacts across sequencing runs, suggesting that fully independent technical replicates would improve data quality. Although most of the spurious taxa were characterized by relative abundances between 0.25 and 2% in the IMNGS-amplicon datasets tested, they represented very dominant populations in a few samples (Fig. 3d).Fig. 3: Origin and occurrence of spurious taxa.a Taxonomic profile and ecological distribution. Inner ring: SILVA-based classification of all non-redundant spurious molecular species at the phylum and family level. Outer colored ring: sample type characterized by the highest prevalence for the given taxon. Outer bars: corresponding highest prevalence values. Only samples with relative abundances >0.25% for any given OTU were counted as positive for prevalence calculation. The total numbers of samples considered were: human, 46,153; soil, 29,864; freshwater, 13,977; mouse, 10,409; marine, 8478. b Distribution of the spurious taxa across sample types. The exclusivity of each OTU for any given sample type was assessed using a Z-test: those assumed to be non-specific for any given sample type appear in red (p 0.25% in at least one replicate were kept). Richness was calculated using ampvis2 [29]. Applying the 0.25% cutoff decreased the number of observed ASVs from 408 ± 71 to 139 ± 5 and, more importantly, the IQR from 101 to 7 (Fig. 6b). Unweighted UniFrac distances within and between runs as calculated using ampvis2 were also compared before and after filtering. Sequences were aligned using MAFFT [30] and phylogeny was inferred using FastTree. Whilst the community makeup in the soil sample varied substantially between sequencing runs without additional filtering, the 0.25% cutoff reduced this variation to the level observed within runs without filtering (Fig. 6c). Replicates within a run were very similar after applying the 0.25% cutoff. Altogether, these data serve as an independent confirmation that stringent filtering delivers more stable values obtained for the exact same sample sequenced in replicates across several sequencing runs. More

  • in

    Insights into rumen microbial biosynthetic gene cluster diversity through genome-resolved metagenomics

    2,809 draft MAGs from the rumen ecosystemWe amassed 3.2 terabase pairs (Tbp) of data from 346 publicly available and 66 new rumen metagenome datasets (Supplementary Table 1). The metagenomes were from cattle (312 samples, 2.1 Tbp), sheep (75 samples, 888.4 gigabase pairs (Gbp)), moose (9 samples, 108.8 Gbp), deer (8 samples, 62.9 Gbp), and bison (8 samples, 52.3 Gbp). Metagenomes were assembled independently to reduce the influence of strain variation and improve the recovery of closely related genomes18,19. Following refinement, dereplication, and filtering of resulting population genomes, we identified 2,809 nonredundant MAGs satisfying the following criteria: dRep20 genome quality score ≥60, ≥75% complete, ≤10% contamination, N50 ≥5 kbp, and ≤500 contigs.The median estimated completeness and contamination of the MAGs were 89.7% and 0.9%, respectively (Fig. 1a and Supplementary Data 1). Further, recovered MAGs had a median genome size of 2.2 Mbp, a median of 131 contigs, and a median N50 of 28.3 kbp (Fig. 1b). The proposed minimum information about a MAG (MIMAG) specifies high-quality draft genomes to have an estimated ≥90% completeness, ≤5% contamination, at least 18 tRNAs, and contain 23S, 16S, and 5S rRNA genes21. It remains challenging to reconstruct rRNA genes from short metagenomic reads due to the high sequence similarity of rRNA genes in closely related species. As a result, despite high estimated completeness and low contamination rates, only 20 MAGs meet the MIMAG standards for a high-quality draft genome. We identified a 16S rRNA gene in 197 of the MAGs. The remaining MAGs are characterized as medium-quality MAGs under the MIMAG standards.Fig. 1: Genomic properties of 2,809 rumen MAGs.a CheckM completeness and contamination estimates for the 2,809 population genomes recovered from rumen metagenomes. The size of the point on the scatter plot corresponds to the dRep genome quality score, where Quality = Completeness − (5 ⋅ Contamination) + (Contamination ⋅ (Strain Heterogeneity/100)) + 0.5 ⋅ (({mathrm{log}},)(N50). The reported MAGs meet the following minimum criteria: genome quality score ≥60, ≥75% complete, ≤10% contamination, N50 ≥5 kbp, and ≥500 contigs. b The frequency distribution of the number of contigs and genome sizes of reconstructed MAGs.Full size imageThe majority of bacterial MAGs belonged to phyla Firmicutes or Bacteroidota (2,326; Fig. 2a and Supplementary Data 1). Additionally, we assembled 12 bacterial genomes from the superphylum Patescibacteria. At lower taxonomic ranks, Lachnospiraceae (415) and Prevotella (398) were the dominant family and genus identified among the assembled bacterial genomes. The most prevalent archaeal family and genus were Methanobacteriaceae (45) and Methanobrevibacter (35), respectively (Fig. 2b). The recovered MAGs represent several new taxonomic lineages, as four genomes could not be classified at the rank of order, 16 at the rank of family, and 243 at the genus rank.Fig. 2: Phylogenetic relationships and coverage patterns of near-complete MAGs.a Phylogenomic analysis of 1,163 near-complete (≥90% complete, ≤5% contamination, and N50 ≥15 kbp) bacterial MAGs and (b) 20 near-complete archaeal MAGs inferred from the concatenation of phylogenetically informative proteins. Layers below the genomic trees designate bacterial phylum or archaeal genus based on GTDB taxonomic assignments, genomic size (0–5 Mbp), and the mean number of bases with ≥1× coverage in a rumen metagenomic dataset (layer color indicates the ruminant the data was collected from). The mean number of bases with ≥1× coverage was used as input for hierarchical clustering of rumen metagenomic datasets based on Euclidean distance and Ward linkage. The bacterial and archaeal phylogenetic trees are provided as Supplementary Data 6 and Supplementary Data 7, respectively.Full size imageSpecies-level overlap between reference genomes, the Hungate1000 Collection, and rumen MAGsTo further characterize the assembled genomes, we compared the MAGs to other rumen-specific genome collections, specifically genomes generated from the Hungate1000 project3 and MAGs identified from the Stewart et al. studies4,5. We clustered genomes based on approximate species-level thresholds (≥95% ANI) and calculated the intersection between MAGs in the current study and the Hungate1000 Collection (410 genomes)3, MAGs from Stewart et al. (4,941 genomes)4,5, and a dereplicated genome collection from the GTDB (22,441 genomes, see Methods)22, which includes reference isolate genomes and some environmental MAGs23. It should be noted that we used the raw data from the first of the Stewart et al. studies4 (Supplementary Table 1), but with different assembly and binning approaches. Approximately one-third of the MAGs (1,007) did not exhibit ≥95% ANI with a genome in the GTDB, Stewart et al. MAGs, or the Hungate1000 isolates (Fig. 3a). When considering the pairwise intersections between the datasets, 98 (3.5%), 933 (33.2%), and 1,438 (51.2%) of the MAGs in the current study had ≥95% ANI with a genome in the Hungate1000 Collection3, GTDB22, and Stewart et al.4,5, respectively. One hundred twenty-one (29.5%), 552 (2.5%), and 3,125 (63.2%) of the genomes from the Hungate1000 Collection3, GTDB22, and Stewart et al.4,5 displayed ≥95% ANI with a MAG from the current study. Together, these results indicate that we recovered a majority of previous rumen genomic diversity with additional lineages not previously identified in other major rumen genomic collections.Fig. 3: Genomes sharing ≥95% ANI between databases and the characterization of rumen-specific 95% ANI clusters.a The approximate number of species overlapping amongst rumen-specific and reference genomic datasets. Genomes demonstrating ≥95% ANI were considered to be shared between two datasets. Presented are a subset of intersections in which a MAG from the current study was the query genome. b The number of genomes comprising each of the 3,541 95% ANI clusters generated from 8,160 rumen microbial genomes in the current study, the Hungate1000 Collection3, and Stewart et al. studies4, 5. c Rarefaction analysis based on subsampling 95% ANI clusters at steps of 500 genomes indicates the 8,160 genomes from recently published rumen genomic collections still only represent a fraction of expected microbial species diversity in the rumen ecosystem. Phylogenomic relationships of the 1,781 near-complete bacterial (d) and 35 near-complete archaeal (e) representative genomes with the highest dRep genome quality score from the 3,541 95% ANI clusters generated from 8,160 rumen-specific genomes. Near-complete genomes were defined as being ≥90% complete, having ≤5% contamination, and contig N50 ≥15 kbp. Layers surrounding the genomic trees indicate the bacterial phyla or archaeal genera and the log normalized number of genomes from each rumen genomic collection belonging to the same 95% ANI cluster. The bacterial and archaeal phylogenetic trees are provided as Supplementary Data 8 and Supplementary Data 9, respectively.Full size imageWe applied an additional clustering approach to identify the approximate number of species represented by the rumen-specific genomes assembled in this study, in the Hungate1000 Collection3, and Stewart et al.4,5. A 95% ANI threshold yielded 3,541 clusters from the combination of the datasets (Supplementary Data 2). Of the 3,541 clusters, 2,024 contained a MAG from the current study, and 1,135 were composed exclusively of MAGs from the current study. In comparison, 2,175 and 286 clusters were comprised of genomes from Stewart et al.4,5 and the Hungate1000 Collection3, respectively. The majority of 95% ANI clusters (2,166) are only comprised of a single genome (Fig. 3b). Furthermore, a rarefaction curve suggests the 8,160 genomes from the genomic collections analyzed here only represent a fraction of the estimated microbial species diversity in the rumen (Fig. 3c). The genome with the best dRep score from each cluster was used to generate a phylogenetic tree highlighting the species diversity within each rumen genomic collection and represents the vast diversity of rumen bacterial (Fig. 3d) and archaeal (Fig. 3e) genomes published to date.As stated previously, the median genome size of reconstructed MAGs was 2.2 Mbp, smaller than the median size of genomes from the Hungate1000 project (3.1 Mbp)3. To provide an assessment at a finer resolution, genome sizes of MAGs and Hungate1000 genomes3 belonging to the same 95% ANI cluster were compared (Supplementary Fig. 1). Adjusted sizes of MAGs and Hungate1000 genomes that are ≥95% complete displayed a regression coefficient of 0.96 with a slope of 0.86, indicating the binning process likely did not lead to extensive losses and systematic biases in the reconstructed genomes. Instead, it further highlights that current culturing approaches have not brought large portions of rumen microbial diversity into culture and putatively supports previous findings from the human gut that revealed genome-reduction in uncultured bacteria24.Rumen metagenome classification rates using reference and rumen-specific genomesUtilizing an approach similar to Stewart et al.4,5, we investigated the influence of MAGs on rates of metagenomic read classification. The baseline for read classification was the standard Kraken database containing bacterial, archaeal, fungal, and protozoal RefSeq genomes25. Each rumen-specific dataset was incrementally added to the Kraken RefSeq genomic database in the following order to build new databases: the Hungate1000 Collection3, MAGs from Stewart et al.4,5, and MAGs from the current study. Each individual and collective database was used for classification of sample reads that underpinned metagenomic binning and from a rumen metagenomic dataset not used in the reconstruction of MAGs26. MAGs from the current work classified more reads from deer, moose, and sheep metagenomes, while the more numerous MAGs from Stewart et al.4,5 classified more reads from bison and cattle metagenomes (Supplementary Fig. 2a). The addition of MAGs improves classification relative to databases primarily based on cultured isolates, like the Hungate1000 Collection3 (Supplementary Fig. 2b). Using the combination of all reference and rumen-specific genomes, the median classification rate on an independent set of cattle metagenomes was 62.6%.Phylogenetic characterization of biosynthetic gene clustersMicrobial genome mining is a powerful tool for natural product discovery. We sought to explore the extent of secondary metabolite diversity coded by the MAGs in the current study, the Hungate1000 Collection3, and Stewart et al. MAGs4,5. We identified 14,814 BGCs encoded by the 8,160 rumen-specific genomes using antiSMASH27 (Fig. 4a and Supplementary Data 3). The majority of BGCs were NRPS (5,346), followed by aryl polyenes (2,800), sactipeptides (2,126), and bacteriocins (1,943). Only a few PKS were identified (75). Firmicutes harbored the vast majority of clusters for NRPS, sactipeptide, lantipeptide, lassopeptide, and bacteriocin synthesis (Fig. 4b). At lower taxonomic ranks, DTU089 (979), Bacteroidaceae (934), and Lachnospiraceae (923) coded for the bulk of NRPS gene clusters. Moreover, Acidaminococcaceae genomes contained 21.2% of identified bacteriocins and Ruminococcus spp. possessed the bulk of sactipeptides and lantipeptides. Archaea were predicted to code 737 BGCs, including an average of 3.8 NRPS gene clusters per genome (Fig. 4a).Fig. 4: Characterization of BGCs from 8,160 rumen genomes and MAGs.a Number and types of BGCs identified from select phyla in genomes from the Hungate1000 Collection3, Stewart et al. studies4, 5, and the current study. b Phylogenomic analysis of 1,766 near-complete Firmicutes genomes inferred from the concatenation of phylogenetically informative proteins. The inner layer surrounding the genomic tree designates taxonomic annotations, while the remaining layers depict the log normalized number of BGCs in the genome with the ascribed function. Bacterial class and order labels are displayed for those lineages in which more than 50 genomes were identified. Near-complete genomes were defined as being ≥90% complete, having ≤5% contamination, and contig N50 ≥15 kbp. The phylogenetic tree is provided as Supplementary Data 10. c A relational network of NRPS gene clusters in Firmicutes, Bacteroidota, and Euryarchaeota highlights the similarity of NRPS BGCs from Euryarchaeota and Firmicutes. Edge weight represents the similarity of two BGCs, as determined by BiG-SCAPE (i.e. darker edges demonstrate more similarity between two BGCs). Edges are only shown for BGCs with ≥0.3 BiG-SCAPE similarity. Nodes from each phylum are duplicated to illustrate intra-phylum relationships and nodes along a given axis are ordered alphabetically by taxonomic family. d The association between genome phylogeny and the similarity of NRPS gene clusters coded by near-complete Euryarchaeota genomes. BGCs designated as NRPS were clustered with BiG-SCAPE. The relationship between NRPS clusters was portrayed through the hierarchical clustering of pairwise inter-cluster similarities. The number of NRPS clusters coded by each genome (range of 0–3) is presented alongside the assigned genus. A group of Methanobrevibacter genomes, likely of the same species (≥95% ANI), possessed very similar NRPS clusters (highlighted in blue). Yet, phylogenetically closely related genomes, belonging to two different 95% ANI clusters, did not code for any identified NRPS gene clusters (highlighted in red). The phylogenetic tree is based on the concatenation of 122 phylogenetically informative archaeal proteins and is available as Supplementary Data 11.Full size imageNRPS exhibit high molecular and structural diversity resulting in a wide array of biological activities. The diversity of NRPS, combined with their proteolytic stability and selective bioactivity, has resulted in the development of many NRPS as antimicrobials and other therapeutic agents28. Given the prevalence of NRPS among the recovered MAGs (Fig. 4a), the peptides appear to be important bioactive metabolites in the rumen. To gain fundamental insight into the phylogenetic diversity of rumen NRPS, we built a network based on BGC similarity using BiG-SCAPE29. BiG-SCAPE uses protein domain content, order, copy number, and sequence identity to calculate a distance metric. We assessed the similarity of NRPS gene clusters identified in Firmicutes, Bacteroidota, and Euryarchaeota, as these three phyla coded for 96.4% of assembled NRPS gene clusters from rumen genomes. With a BiG-SCAPE similarity threshold of 0.3, the resulting network consisted of 3,436 nodes (NRPS BGCs on contigs ≥10 kbp) and 79,112 edges (Fig. 4c and Supplementary Data 4). As expected, the network analysis depicted high inter- and intra-phylum genetic diversity among the NRPS gene clusters. The median intra-phylum, -family, and -genus similarity was 0.40, 0.44, and 0.46, respectively, while the median inter-phylum, -family, and -genus similarity was 0.32, 0.34, and 0.34, respectively. Further, only 2.6% of edges were inter-phylum and 69.0% were intra-family. Of the 6,594 Euryarchaeota edges, 8.1% were Euryarchaeota-Firmicutes (median similarity of 0.32) and 2.0% of edges were Euryarchaeota-Bacteroidota (median similarity of 0.31). To further examine the phylogenetic relationships of rumen Euryarchaeota NRPS, we clustered 265 NRPS gene clusters (≥10 kbp) from 85 near-complete Euryarchaeota genomes at a higher similarity threshold of 0.75, yielding 57 NRPS clusters (Fig. 4d). The distribution of NRPS clusters amongst the genomes suggests there exists a strong relationship between methanogen phylogeny and NRPS similarity. Only Methanobrevibacter genomes contain NRPS gene clusters, and genomes of the same species often possessed many of the same NRPS clusters (see genomes highlighted in blue in Fig. 4d). However, there are instances in which closely related methanogens code for a contrasting pattern of NRPS clusters or no NRPS clusters at all (see genomes highlighted in red in Fig. 4d).Bacteriocins likely serve as regulatory elements in complex microbial communities such as the rumen. Consequently, bacteriocins have been studied and characterized for their bactericidal activity and as agents that modulate the microbiota structure and function30. In particular, lanthipeptides, a class of ribosomally synthesized and post-translationally modified peptides (RiPPs) with thioether cross-linked amino acids31, are of pharmaceutical, preservative, and agricultural interest due to their strong antimicrobial properties against gram-positive pathogens31,32,33, low levels of antimicrobial resistance34, and stability35. We identified 195 rumen lanthipeptide BGCs from the Hungate1000 genomes and MAGs from Stewart et al. and the current study. Rumen lanthipeptide BGCs were clustered with 22,870 lanthipeptide BGCs from RefSeq genomes36,37 into gene cluster families (GCFs; groups of BGCs that may generate highly similar products). Clustering with BiG-SCAPE29 yielded 4,565 GCFs, 120 of which contained a rumen lanthipeptide. The 120 GCFs were composed of 519 lanthipeptide BGCs, where 324 were from RefSeq isolates and 195 from rumen genomes (Fig. 5a). The 324 RefSeq BGCs fell into only 18 GCFs. Lanthipeptides from the Hungate1000 isolates clustered into 36 GCFs, while rumen MAG lanthipeptides belonged to 92 GCFs, 82 of which were exclusively composed of MAG lanthipeptides. Together, this evidence suggests rumen MAGs code for diverse and novel lanthipeptides not represented in cultured isolates, including the Hungate Collection.Fig. 5: Phylogenetic diversity of 195 lanthipeptide BGCs coded by rumen genomes.a Network depicting the similarity between lanthipeptide BGCs identified from complete and draft isolate genomes in RefSeq and rumen genomes of the Hungate1000 collection, Stewart et al. MAGs, and MAGs from the current study. The BGCs were clustered into gene cluster families (GCFs) with BiG-SCAPE29. Only the GCFs containing a rumen genome and at least two BGCs were visualized. Nodes in the network represent BGCs and edges connect BGCs with BiG-SCAPE defined similarity ≥0.3. b Phylogenetic relationships of 120 near-complete rumen bacterial genomes coding for lanthipeptide BGCs. Near-complete genomes were defined as being ≥90% complete, having ≤5% contamination, and contig N50 ≥15 kbp. Layers surrounding the genomic trees indicate the bacterial phyla and family, if the genome is a MAG or Hungate Collection isolate, and the class of lanthipeptide, as predicted by antiSMASH27. Genomes without an indicated lanthipeptide class were not classified by antiSMASH. The phylogenetic tree is based on the concatenation of 120 phylogenetically informative bacterial proteins and is available as Supplementary Data 12.Full size imageWe sought to further examine the differences in rumen MAG lanthipeptides relative to isolates and the taxonomic diversity of rumen microbes coding for lanthipeptides. The 195 rumen lanthipeptides were mainly found in Firmicutes genomes, with a subset from Bacteroidota and Actinobacteriota (Fig. 5b). Fifty-two of the 55 lanthipeptides from the Hungate Collection isolates were from Firmicutes (94.5%). At the family-level, these 52 Firmicutes BGCs were distributed evenly between Lachnospiraceae and Streptococcaceae. In contrast, 19.2% and 8.6% of lanthipeptides from rumen MAGs belonged to Bacteroidota and Actinobacteriota, respectively. Lanthipeptides from MAGs were also found in Muribaculaceae and Oscillospiraceae. Moreover, 26.4% of rumen MAG lanthipeptides, compared to 3.6% of Hungate Collection isolates, were found in Eubacterium genomes. The majority of Eubacterium MAG lanthipeptides (62.1%) belonged to a single GCF, suggesting they code for very similar products. Lastly, antiSMASH predicted the bulk of the rumen lanthipeptides were Class II lanthipeptides, with fewer Class I and Class III types (Fig. 5b). Nearly all of the Class I lanthipeptides were from Hungate isolates. The above analysis of lanthipeptide diversity further supports that rumen MAGs code for novel secondary metabolites not represented in cultured isolates.We aligned previously published rumen metatranscriptome data from steers characterized as having high and low feed efficiency to the BGCs to demonstrate if the identified BGCs are active and to explore potential ecological roles of secondary metabolites. Despite data from the metatranscriptome study not being applied to reconstruct genomes in the current study, we identified the expression of 554 gene clusters from rumen-specific genomes in the 20 metatranscriptomes (≥100 aligned reads). Metatranscriptome read count data were normalized independently for each genome to better account for the variation in taxonomic composition across samples38. Genome-specific normalization resulted in the identification of 17 differentially expressed gene clusters between steers with high and low feed efficiency (DESeq239 false discovery rate adjusted P  More

  • in

    Decline in symbiont-dependent host detoxification metabolism contributes to increased insecticide susceptibility of insects under high temperature

    1.Bálint M, Domisch S, Engelhardt CHM, Haase P, Lehrian S, Sauer J, et al. Cryptic biodiversity loss linked to global climate change. Nat Clim Chang. 2011;1:313–8.Article 

    Google Scholar 
    2.Parmesan C, Yohe G. A globally coherent fingerprint of climate change impacts across natural systems. Nature. 2003;421:37–42.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    3.Blois JL, Zarnetske PL, Fitzpatrick MC, Finnegan S. Climate change and the past, present, and future of biotic interactions. Science. 2013;341:499–504.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    4.Haines A, Ebi K. The imperative for climate action to protect health. N. Engl J Med. 2019;380:263–73.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    5.Deutsch CA, Tewksbury JJ, Tigchelaar M, Battisti DS, Merrill SC, Huey RB, et al. Increase in crop losses to insect pests in a warming climate. Science. 2018;361:916–9.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    6.Kattwinkel M, Jan-Valentin K, Foit K, Liess M. Climate change, agricultural insecticide exposure, and risk for freshwater communities. Ecol Appl. 2011;21:2068–81.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    7.Moe SJ, De Schamphelaere K, Clements WH, Sorensen MT, Van den Brink PJ, Liess M. Combined and interactive effects of global climate change and toxicants on populations and communities. Environ Toxicol Chem. 2013;32:49–61.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    8.Potts SG, Biesmeijer JC, Kremen C, Neumann P, Schweiger O, Kunin WE. Global pollinator declines: trends, impacts and drivers. Trends Ecol Evol. 2010;25:345–53.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    9.Moran EV, Alexander JM. Evolutionary responses to global change: Lessons from invasive species. Ecol Lett. 2014;17:637–49.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    10.Harwood AD, You J, Lydy MJ. Temperature as a toxicity identification evaluation tool for pyrethroid insecticides: toxicokinetic confirmation. Environ Toxicol Chem. 2009;28:1051–8.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    11.Guo L, Su M, Liang P, Li S, Chu D. Effects of high temperature on insecticide tolerance in whitefly Bemisia tabaci (Gennadius) Q biotype. Pestic Biochem Physiol. 2018;150:97–104.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    12.Mao K, Jin R, Li W, Ren Z, Qin X, He S, et al. The influence of temperature on the toxicity of insecticides to Nilaparvata lugens (Stål). Pestic Biochem Physiol. 2019;156:80–86.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    13.Verheyen J, Delnat V, Stoks R. Increased daily temperature fluctuations overrule the ability of gradual thermal evolution to offset the increased pesticide toxicity under global warming. Environ Sci Technol. 2019;53:4600–8.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    14.Moran NA. Symbiosis as an adaptive process and source of phenotypic complexity. Proc Natl Acad Sci USA. 2007;104:8627–3863.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    15.Kikuchi Y, Hayatsu M, Hosokawa T, Nagayama A, Tago K, Fukatsu T. Symbiont-mediated insecticide resistance. Proc Natl Acad Sci USA. 2012;109:8618–22.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    16.Jones RM, Desai C, Darby TM, Luo L, Wolfarth AA, Scharer CD, et al. Lactobacilli modulate epithelial cytoprotection through the Nrf2 pathway. Cell Rep. 2015;12:1217–25.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    17.Cheng D, Guo Z, Riegler M, Xi Z, Liang G, Xu Y. Gut symbiont enhances insecticide resistance in a significant pest, the oriental fruit fly Bactrocera dorsalis (Hendel). Microbiome. 2017;5:13.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    18.Pang R, Chen M, Yue L, Xing K, Li T, Kang K, et al. A distinct strain of Arsenophonus symbiont decreases insecticide resistance in its insect host. PLoS Genet. 2018;14:e1007725.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    19.Kikuchi Y, Tada A, Musolin DL, Hari N, Hosokawa T, Fujisaki K, et al. Collapse of insect gut symbiosis under simulated climate change. mBio. 2016;7:e01578–16.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    20.Corbin C, Heyworth ER, Ferrari J, Hurst GDD. Heritable symbionts in a world of varying temperature. Heredity. 2017;118:10–20.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Jia FX, Yang MS, Yang WJ, Wang JJ. Influence of continuous high temperature conditions on Wolbachia infection frequency and the fitness of Liposcelis tricolor (Psocoptera: Liposcelididae). Environ Entomol. 2009;38:1365–72.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    22.Burke G, Fiehn O, Moran N. Effects of facultative symbionts and heat stress on the metabolome of pea aphids. ISME J. 2010;4:242–52.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    23.Fan Y, Wernegreen JJ. Can’t take the heat: high temperature depletes bacterial endosymbionts of ants. Micro Ecol. 2013;66:727–33.Article 

    Google Scholar 
    24.Hussain M, Akutse KS, Ravindran K, Lin Y, Bamisile BS, Qasim M, et al. Effects of different temperature regimes on survival of Diaphorina citri and its endosymbiotic bacterial communities. Environ Microbiol. 2017;19:3439–49.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    25.Engl T, Eberl N, Gorse C, Krüger T, Schmidt THP, Plarre R, et al. Ancient symbiosis confers desiccation resistance to stored grain pest beetles. Mol Ecol. 2018;27:2095–108.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    26.Zhang XJ, Yu XP, Chen JM. High Temperature effects on yeast-like endosymbiotes and pesticide resistance of the small brown planthopper, Laodelphax striatellus. Rice Sci. 2008;15:326–30.CAS 
    Article 

    Google Scholar 
    27.Zhang B, Zuo TQ, Li HG, Sun LJ, Wang SF, Zhang CY, et al. Effect of heat shock on the susceptibility of Frankliniella occidentalis (Thysanoptera: Thripidae) to insecticides. J Integr Agric. 2016;15:2309–18.CAS 
    Article 

    Google Scholar 
    28.Karimzadeh R, Javanshir M, Hejazi MJ. Individual and combined effects of insecticides, inert dusts and high temperatures on Callosobruchus maculatus (Coleoptera: Chrysomelidae). J Stored Prod Res. 2020;89:10693.Article 

    Google Scholar 
    29.Michigan State University. Arthropod Pesticide Resistance Database (APRD). East Lansing: Michigan State University; 2020. http://www.pesticideresistance.com/.30.Ju JF, Bing XL, Zhao DS, Guo Y, Xi Z, Hoffmann AA, et al. Wolbachia supplement biotin and riboflavin to enhance reproduction in planthoppers. ISME J. 2019;14:676–87.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    31.Zhang Y, Tang T, Li W, Cai T, Li J, Wan H. Functional profiling of the gut microbiomes in two different populations of the brown planthopper. Nilaparvata lugens J Asia Pac Entomol. 2018;21:1309–14.Article 

    Google Scholar 
    32.Ye YH, Seleznev A, Flores HA, Woolfit M, McGraw EA. Gut microbiota in Drosophila melanogaster interacts with Wolbachia but does not contribute to Wolbachia-mediated antiviral protection. J Invertebr Pathol. 2017;143:18–25.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    33.Yamada R, Floate KD, Riegler M, O’Neill SL. Male development time influences the strength of Wolbachia-induced cytoplasmic incompatibility expression in Drosophila melanogaster. Genetics. 2007;177:801–8.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    34.Wari D, Kabir MA, Mujiono K, Hojo Y, Shinya T, Tani A, et al. Honeydew-associated microbes elicit defense responses against brown planthopper in rice. J Exp Bot. 2019;70:1683–96.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    35.Miller ALE, Tindall K, Leonard BR. Bioassays for monitoring insecticide resistance. J Vis Exp. 2010;46:2129.
    Google Scholar 
    36.Zhang J, Zhang Y, Wang Y, Yang Y, Cang X, Liu Z. Expression induction of P450 genes by imidacloprid in Nilaparvata lugens: a genome-scale analysis. Pestic Biochem Physiol. 2016;132:59–64.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    37.Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2-ΔΔCT method. Methods. 2001;25:402–8.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    38.Noda H, Koizumi Y, Zhang Q, Deng K. Infection density of Wolbachia and incompatibility level in two planthopper species, Laodelphax striatellus and Sogatella furcifera. Insect Biochem Mol Biol. 2001;31:727–37.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    39.Martin M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 2011. https://doi.org/10.14806/ej.17.1.20040.Callahan BJ, McMurdie PJ, Rosen MJ, Han AW, Johnson AJA, Holmes SP. DADA2: high-resolution sample inference from illumina amplicon data. Nat Methods. 2016;13:581–3.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    41.Katoh K, Misawa K, Kuma KI, Miyata T. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 2002;30:3059–66.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    42.Bokulich NA, Kaehler BD, Rideout JR, Dillon M, Bolyen E, Knight R, et al. Optimizing taxonomic classification of marker-gene amplicon sequences with QIIME 2’s q2-feature-classifier plugin. Microbiome. 2018;6:90.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    43.Liu S, Ding Z, Zhang C, Yang B, Liu Z. Gene knockdown by intro-thoracic injection of double-stranded RNA in the brown planthopper, Nilaparvata lugens. Insect Biochem Mol Biol. 2010;40:666–71.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    44.Tai V, James ER, Nalep CA, Scheffrahn RH, Perlman SJ, Keelinga PJ. The role of host phylogeny varies in shaping microbial diversity in the hindguts of lower termites. Appl Environ Microbiol. 2015;81:1059–70.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    45.Bale JS, Hayward SAL. Insect overwintering in a changing climate. J Exp Biol. 2010;213:980–94.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    46.Rahmstorf S, Cazenave A, Church JA, Hansen JE, Keeling RF, Parker DE, et al. Recent climate observations compared to projections. Science. 2007;316:709.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    47.Radchuk V, Reed T, Teplitsky C, van de Pol M, Charmantier A, Hassall C, et al. Adaptive responses of animals to climate change are most likely insufficient. Nat Commun. 2019;10:3019.Article 
    CAS 

    Google Scholar 
    48.Iwamura T, Guzman-Holst A, Murray KA. Accelerating invasion potential of disease vector Aedes aegypti under climate change. Nat Commun. 2020;11:2130.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    49.Li J, Mao T, Wang H, Lu Z, Qu J, Fang Y, et al. The CncC/keap1 pathway is activated in high temperature-induced metamorphosis and mediates the expression of Cyp450 genes in silkworm, Bombyx mori. Biochem Biophys Res Commun. 2019;541:1045–50.Article 
    CAS 

    Google Scholar 
    50.Kalsi M, Palli SR. Transcription factor cap n collar C regulates multiple cytochrome P450 genes conferring adaptation to potato plant allelochemicals and resistance to imidacloprid in Leptinotarsa decemlineata (Say). Insect Biochem Mol Biol. 2017;83:1–12.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    51.Kalsi M, Palli SR. Transcription factors, CncC and Maf, regulate expression of CYP6BQ genes responsible for deltamethrin resistance in Tribolium castaneum. Insect Biochem Mol Biol. 2015;65:47–56.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    52.Misra JR, Lam G, Thummel CS. Constitutive activation of the Nrf2/Keap1 pathway in insecticide-resistant strains of Drosophila. Insect Biochem Mol Biol. 2013;43:1116–24.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    53.Tang B, Cheng Y, Li Y, Li W, Ma Y, Zhou Q, et al. Adipokinetic hormone regulates cytochrome P450-mediated imidacloprid resistance in the brown planthopper, Nilaparvata lugens. Chemosphere. 2020;259:127490.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    54.Cheng Y, Li Y, Li W, Song Y, Zeng R, Lu K. Inhibition of hepatocyte nuclear factor 4 confers imidacloprid resistance in Nilaparvata lugens via the activation of cytochrome P450 and UDP-glycosyltransferase genes. Chemosphere. 2021;263:128269.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    55.Li Y, Liu X, Wang N, Zhang Y, Hoffmann AA, Guo H. Background-dependent Wolbachia-mediated insecticide resistance in Laodelphax striatellus. Environ Microbiol. 2020;22:2653–63.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    56.Berticat C, Rousset F, Raymond M, Berthomieu A, Weill M. High Wolbachia density in insecticide-resistant mosquitoes. Proc R Soc B Biol Sci. 2002;269:1413–6.Article 

    Google Scholar 
    57.Zhang G, Hussain M, O’Neill SL, Asgari S. Wolbachia uses a host microRNA to regulate transcripts of a methyltransferase, contributing to dengue virus inhibition in Aedes aegypti. Proc Natl Acad Sci USA. 2013;110:10276–81.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    58.Bi J, Sehgal A, Williams JA, Wang YF. Wolbachia affects sleep behavior in Drosophila melanogaster. J Insect Physiol. 2018;107:81–88.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    59.Roughgarden J, Gilbert SF, Rosenberg E, Zilber-Rosenberg I, Lloyd EA. Holobionts as units of selection and a model of their population dynamics and evolution. Biol Theory. 2018;13:44–65.Article 

    Google Scholar 
    60.Pan X, Zhou G, Wu J, Bian G, Lu P, Raikhel AS, et al. Wolbachia induces reactive oxygen species (ROS)-dependent activation of the Toll pathway to control dengue virus in the mosquito Aedes aegypti. Proc Natl Acad Sci USA. 2012;109:E23–31.PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    61.Gong JT, Li Y, Li TP, Liang Y, Hu L, Zhang D, et al. Stable introduction of plant-virus-inhibiting Wolbachia into planthoppers for rice protection. Curr Biol. 2020;30:4837–45.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    62.Elzaki MEA, Li ZF, Wang J, Xu L, Liu N, Zeng RS, et al. Activiation of the nitric oxide cycle by citrulline and arginine restores susceptibility of resistant brown planthoppers to the insecticide imidacloprid. J Hazard Mater. 2020;396:122755.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    63.Werren JH. Biology of Wolbachia. Annu Rev Entomol. 1997;42:587–609.CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    64.Kokou F, Sasson G, Nitzan T, Doron-Faigenboim A, Harpaz S, Cnaani A, et al. Host genetic selection for cold tolerance shapes microbiome composition and modulates its response to temperature. Elife. 2018;77:e36398.Article 

    Google Scholar  More