More stories

  • in

    The occurrence and ecology of microbial chain elongation of carboxylates in soils

    1.
    Barker HA, Taha SM. Clostridium kluyverii, an organism concerned in the formation of caproic acid from ethyl alcohol. J Bacteriol. 1942;43:347–63.
    CAS  PubMed  PubMed Central  Article  Google Scholar 
    2.
    Angenent LT, Richter H, Buckel W, Spirito CM, Steinbusch KJJ, Plugge CM, et al. Chain elongation with reactor microbiomes: open-culture biotechnology to produce biochemicals. Environ Sci Technol. 2016;50:2796–810.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    3.
    Béchamp MA. Lettre de m. A. Béchamp a m. Dumas. Ann Chim Phys 1868;4:103–11.
    Google Scholar 

    4.
    Weimer PJ, Stevenson DM. Isolation, characterization, and quantification of Clostridium kluyveri from the bovine rumen. Appl Microbiol Biotechnol. 2012;94:461–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Kenealy WR, Waselefsky DM. Studies on the substrate range of Clostridium kluyveri – the use of propanol and succinate. Arch Microbiol. 1985;141:187–94.
    CAS  Article  Google Scholar 

    6.
    Barker HA, Kamen MD, Bornstein BT. The synthesis of butyric and caproic acids from ethanol and acetic acid by Clostridium kluyveri. Proc Natl Acad Sci USA. 1945;31:373–81.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    7.
    Bornstein BT, Barker HA. The energy metabolism of Clostridium kluyveri and the synthesis of fatty acids. J Biol Chem. 1948;172:659–69.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    8.
    Seedorf H, Fricke WF, Veith B, Bruggemann H, Liesegang H, Strittimatter A, et al. The genome of Clostridium kluyveri, a strict anaerobe with unique metabolic features. Proc Natl Acad Sci USA. 2008;105:2128–33.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Gonzalez-Cabaleiro R, Lema JM, Rodriguez J, Kleerebezem R. Linking thermodynamics and kinetics to assess pathway reversibility in anaerobic bioprocesses. Energy Environ Sci. 2013;6:3780–9.
    CAS  Article  Google Scholar 

    10.
    Spirito CM, Richter H, Rabaey K, Stams AJM, Angenent LT. Chain elongation in anaerobic reactor microbiomes to recover resources from waste. Curr Opin Biotechnol. 2014;27:115–22.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    11.
    Rittmann BE & McCarty PL. Environmental Biotechnology: Principles and Applications. McGraw-Hill Book Education: New York; 2001.

    12.
    Thauer RK, Jungermann K, Henninger H, Wenning J, Decker K. The energy metabolism of Clostridium kluyveri. Eur J Biochem. 1968;4:173–80.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    13.
    Stadtman ER, Barker HA. Fatty acid synthesis by enzyme preparations of Clostridium kluyveri. I. Preparation of cell-free extracts that catalyze the conversion of ethanol and acetate to butyrate and caproate. J Biol Chem. 1949;180:1085–93.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    14.
    Stadtman ER, Barker HA. Fatty acid synthesis by enzyme preparations of Clostridium kluyveri. VI. Reactions of acyl phosphates. J Biol Chem. 1950;184:769–93.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    15.
    Steinbusch KJJ, Hamelers HVM, Plugge CM, Buisman CJN. Biological formation of caproate and caprylate from acetate: fuel and chemical production from low grade biomass. Energy Environ Sci. 2011;4:216–24.
    CAS  Article  Google Scholar 

    16.
    Agler MT, Spirito CM, Usack JG, Werner JJ, Angenent LT. Chain elongation with reactor microbiomes: upgrading dilute ethanol to medium-chain carboxylates. Energy Environ Sci. 2012;5:8189–92.
    CAS  Article  Google Scholar 

    17.
    Cavalcante WD, Leitao RC, Gehring TA, Angenent LT, Santaella ST. Anaerobic fermentation for n-caproic acid production: A review. Process Biochem. 2017;54:106–19.
    CAS  Article  Google Scholar 

    18.
    De Groof V, Coma M, Arnot T, Leak DJ, Lanham AB. Medium chain carboxylic acids from complex organic feedstocks by mixed culture fermentation. Molecules 2019;24:398.
    PubMed Central  Article  CAS  Google Scholar 

    19.
    Schievano A, Sciarria TP, Vanbroekhoven K, De Wever H, Puig S, Andersen SJ, et al. Electro-fermentation – merging electrochemistry with fermentation in industrial applications. Trends Biotechnol. 2016;34:866–78.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    20.
    Jourdin L, Raes SMT, Buisman CJN, Strik D. Critical biofilm growth throughout unmodified carbon felts allows continuous bioelectrochemical chain elongation from CO2 up to caproate at high current density. Front Energy Res. 2018;6:7.
    Article  Google Scholar 

    21.
    Candry P, Huang SL, Carvajal-Arroyo JM, Rabaey K, Ganigue R. Enrichment and characterisation of ethanol chain elongating communities from natural and engineered environments. Sci Rep. 2020;10:1–10.
    Article  CAS  Google Scholar 

    22.
    Conrad R. Importance of hydrogenotrophic, aceticlastic and methylotrophic methanogenesis for methane production in terrestrial, aquatic and other anoxic environments: a mini review. Pedosphere 2020;30:25–39.
    Article  Google Scholar 

    23.
    Rui JP, Peng JJ, Lu YH. Succession of bacterial populations during plant residue decomposition in rice field soil. Appl Environ Microbiol. 2009;75:4879–86.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    24.
    Tsutsuki K, Ponnamperuma FN. Behavior of anaerobic decomposition in submerged soils – effect of organic material amendment, soil properties, and temperature. Soil Sci Plant Nutr. 1987;33:13–33.
    CAS  Article  Google Scholar 

    25.
    Roy R, Kluber HD, Conrad R. Early initiation of methane production in anoxic rice soil despite the presence of oxidants. FEMS Microbiol Ecol. 1997;24:311–20.
    CAS  Article  Google Scholar 

    26.
    Adeleke R, Nwangburuka C, Oboirien B. Origins, roles and fate of organic acids in soils: a review. S Afr J Bot. 2017;108:393–406.
    CAS  Article  Google Scholar 

    27.
    Mohana Rangan S, Mouti A, LaPat-Polasko L, Lowry GV, Krajmalnik-Brown R, Delgado A. Synergistic zero-valent iron (Fe0) and microbiological trichloroethene and perchlorate reductions are determined by the concentration and speciation of Fe. Environ Sci Technol. 2020;54:14422–31.
    Article  CAS  Google Scholar 

    28.
    Delgado AG, Kang D-W, Nelson KG, Fajardo-Williams D, Miceli JF, III, Done HY, et al. Selective enrichment yields robust ethene-producing dechlorinating cultures from microcosms stalled at cis-dichloroethene. PLoS ONE. 2014;9:e100654.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    29.
    Delgado AG, Fajardo-Williams D, Popat SC, Torres CI, Krajmalnik-Brown R. Successful operation of continuous reactors at short retention times results in high-density, fast-rate Dehalococcoides dechlorinating cultures. Appl Microbiol Biotechnol. 2014;98:2729–37.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    30.
    Chen TF, Delgado AG, Yavuz BM, Maldonado J, Zuo Y, Kamath R, et al. Interpreting interactions between ozone and residual petroleum hydrocarbons in soil. Environ Sci Technol. 2017;51:506–13.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    31.
    Esquivel-Elizondo S, Miceli J, Torres CI, Krajmalnik-Brown R. Impact of carbon monoxide partial pressures on methanogenesis and medium chain fatty acids production during ethanol fermentation. Biotechnol Bioeng. 2018;115:341–50.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    32.
    Delgado AG, Fajardo-Williams D, Kegerreis KL, Parameswaran P, Krajmalnik-Brown R. Impact of ammonium on syntrophic organohalide-respiring and fermenting microbial communities. mSphere. 2016;1:e00053–16.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    33.
    Delgado AG, Fajardo-Williams D, Bondank E, Esquivel-Elizondo S, Krajmalnik-Brown R. Coupling bioflocculation of Dehalococcoides mccartyi to high-rate reductive dehalogenation of chlorinated ethenes. Environ Sci Technol. 2017;51:11297–307.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    34.
    Esquivel-Elizondo S, Delgado AG, Krajmalnik-Brown R. Evolution of microbial communities growing with carbon monoxide, hydrogen, and carbon dioxide. FEMS Microbiol Ecol. 2017;93:fix076.
    Article  CAS  Google Scholar 

    35.
    Xiaoyu Z, Yong T, Cheng L, Xiangzhen L, Na W, Wenjie Z, et al. The synthesis of n-caproate from lactate: a new efficient process for medium-chain carboxylates production. Sci Rep. 2015;5:14360.
    Article  CAS  Google Scholar 

    36.
    Caporaso JG, Christian LL, William AW, Donna B-L, James H, Noah F, et al. Ultra-high-throughput microbial community analysis on the Illumina HiSeq and MiSeq platforms. ISME J. 2012;6:1621–24.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    37.
    Masella A, Bartram A, Truszkowski J, Brown D, Neufeld J. PANDAseq: paired-end assembler for illumina sequences. BMC Bioinform. 2012;13:31.
    CAS  Article  Google Scholar 

    38.
    Bolyen E, Rideout JR, Dillon MR, Bokulich NA, Abnet CC, Al-Ghalith GA, et al. Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2. Nat Biotechnol. 2019;37:852–57.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    39.
    Callahan BJ, McMurdie PJ, Holmes SP. Exact sequence variants should replace operational taxonomic units in marker-gene data analysis. ISME J. 2017;11:2639–43.
    PubMed  PubMed Central  Article  Google Scholar 

    40.
    Quast C, Pruesse E, Yilmaz P, Gerken J, Schweer T, Yarza P, et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 2013;41:D590–D6.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    41.
    Robeson MS, O’Rourke DR, Kaehler BD, Ziemski M, Dillon MR, Foster JT, et al. RESCRIPt: Reproducible sequence taxonomy reference database management for the masses. bioRxiv. 2020; https://doi.org/10.1101/2020.10.05.326504.

    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.
    Camacho C, Coulouris G, Avagyan V, Ma N, Papadopoulos J, Bealer K, et al. BLAST plus: architecture and applications. BMC Bioinform. 2009;10:1.
    Article  CAS  Google Scholar 

    44.
    Kusel K, Drake HL. Acetate synthesis in soil from a Bavarian beech forest. Appl Environ Microbiol. 1994;60:1370–3.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    45.
    Kusel K, Drake HL. Effects of environmental parameters on the formation and turnover of acetate by forest soils. Appl Environ Microbiol. 1995;61:3667–75.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    46.
    Duddleston KN, Kinney MA, Kiene RP, Hines ME. Anaerobic microbial biogeochemistry in a northern bog: Acetate as a dominant metabolic end product. Glob Biogeochem Cycles. 2002;16:11.1–9.
    Article  CAS  Google Scholar 

    47.
    Thebrath B, Mayer HP, Conrad R. Bicarbonate-dependent production and methanogenic consumption of acetate in anoxic paddy soil. FEMS Microbiol Ecol. 1992;86:295–302.
    CAS  Article  Google Scholar 

    48.
    Delgado AG, Parameswaran P, Fajardo-Williams D, Halden RU, Krajmalnik-Brown R. Role of bicarbonate as a pH buffer and electron sink in microbial dechlorination of chloroethenes. Microb Cell Fact. 2012;11:128.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    49.
    Kucek LA, Spirito CM, Angenent LT. High n-caprylate productivities and specificities from dilute ethanol and acetate: chain elongation with microbiomes to upgrade products from syngas fermentation. Energy Environ Sci. 2016;9:3482–94.
    CAS  Article  Google Scholar 

    50.
    Volker AR, Gogerty DS, Bartholomay C, Hennen-Bierwagen T, Zhu HL, Bobik TA. Fermentative production of short-chain fatty acids in Escherichia coli. Microbiology 2014;160:1513–22.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    51.
    Grootscholten TIM, Steinbusch KJJ, Hamelers HVM, Buisman CJN. Chain elongation of acetate and ethanol in an upflow anaerobic filter for high rate MCFA production. Bioresour Technol. 2013;135:440–5.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    52.
    Reddy MV, Mohan SV, Chang YC. Medium-chain fatty acids (MCFA) production through anaerobic fermentation using Clostridium kluyveri: effect of ethanol and acetate. Appl Biochem Biotechnol. 2018;185:594–605.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    53.
    Scarborough MJ, Lawson CE, Hamilton JJ, Donohue TJ, Noguera DR. Metatranscriptomic and thermodynamic insights into medium-chain fatty acid production using an anaerobic microbiome. mSystems 2018;3:6.
    Article  Google Scholar 

    54.
    Bao S, Wang QY, Zhang PY, Zhang Q, Wu Y, Li F, et al. Effect of acid/ethanol ratio on medium chain carboxylate production with different VFAs as the electron acceptor: insight into carbon balance and microbial community. Energies 2019;12:3720.
    CAS  Article  Google Scholar 

    55.
    Spirito CM, Marzilli AM, Angenent LT. Higher substrate ratios of ethanol to acetate steered chain elongation toward n-caprylate in a bioreactor with product extraction. Environ Sci Technol. 2018;52:13438–47.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    56.
    Coma M, Vilchez-Vargas R, Roume H, Jauregui R, Pieper DH, Rabaey K. Product diversity linked to substrate usage in chain elongation by mixed-culture fermentation. Environ Sci Technol. 2016;50:6467–76.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    57.
    Janssen PH. Identifying the dominant soil bacterial taxa in libraries of 16S rRNA and 16S rRNA genes. Appl Environ Microbiol. 2006;72:1719–28.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    58.
    Spain AM, Krumholz LR, Elshahed MS. Abundance, composition, diversity and novelty of soil Proteobacteria. ISME J 2009;3:992–1000.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    59.
    Johnson JS, Spakowicz DJ, Hong BY, Petersen LM, Demkowicz P, Chen L, et al. Evaluation of 16S rRNA gene sequencing for species and strain-level microbiome analysis. Nat Commun. 2019;10:5029.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    60.
    Hollister EB, Forrest AK, Wilkinson HH, Ebbole DJ, Malfatti SA, Tringe SG, et al. Structure and dynamics of the microbial communities underlying the carboxylate platform for biofuel production. Appl Microbiol Biotechnol. 2010;88:389–99.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    61.
    Mackie RI, Aminov RI, Hu WP, Klieve AV, Ouwerkerk D, Sundset MA, et al. Ecology of uncultivated Oscillospira species in the rumen of cattle, sheep, and reindeer as assessed by microscopy and molecular approaches. Appl Environ Microbiol. 2003;69:6808–15.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    62.
    Ye TR, Cai HY, Liu X, Jiang HL. Dominance of Oscillospira and Bacteroides in the bacterial community associated with the degradation of high-concentration dimethyl sulfide under iron-reducing condition. Ann Microbiol. 2016;66:1199–206.
    CAS  Article  Google Scholar 

    63.
    Konikoff T, Gophna U. Oscillospira: a central, enigmatic component of the human gut microbiota. Trends Microbiol. 2016;24:523–4.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    64.
    Clarke RTJ. Niche in pasture-fed ruminants for the large rumen bacteria Oscillospira, Lampropedia, and Quin’s and Eadie’s ovals. Appl Environ Microbiol. 1979;37:654–7.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    65.
    Lee GH, Rhee MS, Chang DH, Lee J, Kim S, Yoon MH, et al. Oscillibacter ruminantium sp nov., isolated from the rumen of Korean native cattle. Int J Syst Evol Microbiol. 2013;63:1942–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    66.
    Iino T, Mori K, Tanaka K, Suzuki KI, Harayama S. Oscillibacter valericigenes gen. nov., sp nov., a valerate-producing anaerobic bacterium isolated from the alimentary canal of a Japanese corbicula clam. Int J Syst Evol Microbiol. 2007;57:1840–5.
    PubMed  Article  PubMed Central  Google Scholar 

    67.
    Gophna U, Konikoff T, Nielsen HB. Oscillospira and related bacteria – From metagenomic species to metabolic features. Environ Microbiol. 2017;19:835–41.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    68.
    Wang H-J, Dai K, Wang Y-Q, Wang H-F, Zhang F, Zeng RJ. Mixed culture fermentation of synthesis gas in the microfiltration and ultrafiltration hollow-fiber membrane biofilm reactors. Bioresour Technol. 2018;267:650–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    69.
    Fraj B, Ben Hania W, Postec A, Hamdi M, Ollivier B, Fardeau ML. Fonticella tunisiensis gen. nov., sp nov., isolated from a hot spring. Int J Syst Evol Microbiol. 2013;63:1947–50.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    70.
    Collins MD, Lawson PA, Willems A, Cordoba JJ, Fernandezgarayzabal J, Garcia P, et al. The phylogeny of the genus Clostridium – Proposal of 5 new genera and 11 new species combinations. Int J Syst Bacteriol. 1994;44:812–26.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    71.
    BS Jeon, Kim BC, Um Y, et al. BI. Production of hexanoic acid from D-galactitol by a newly isolated Clostridium sp. BS-1. Appl Microbiol Biotechnol. 2010;88:1161–7.
    Article  CAS  Google Scholar 

    72.
    Zhu XY, Zhou Y, Wang Y, Wu TT, Li XZ, Li DP, et al. Production of high-concentration n-caproic acid from lactate through fermentation using a newly isolated Ruminococcaceae bacterium CPB6. Biotechnol Biofuels. 2017;10:102.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    73.
    Robertson WJ, Bowman JP, Franzmann PD, Mee BJ. Desulfosporosinus meridiei sp nov., a spore-forming sulfate-reducing bacterium isolated from gasolene-contaminated groundwater. Int J Syst Evol Microbiol. 2001;51:133–40.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    74.
    Lee YJ, Romanek CS, Wiegel J. Desulfosporosinus youngiae sp nov., a spore-forming, sulfate-reducing bacterium isolated from a constructed wetland treating acid mine drainage. Int J Syst Evol Microbiol. 2009;59:2743–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar  More

  • in

    Spatial patterns in phage-Rhizobium coevolutionary interactions across regions of common bean domestication

    1.
    Breitbart M, Rohwer F. Here a virus, there a virus, everywhere the same virus? Trends Microbiol. 2005;13:278–84.
    CAS  PubMed  Article  PubMed Central  Google Scholar 
    2.
    Hatfull GF. Dark matter of the biosphere: the amazing world of bacteriophage diversity. J Virol. 2015;89:8107–10.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    3.
    Bouvier T, Del Giorgio PA. Key role of selective viral-induced mortality in determining marine bacterial community composition. Environ Microbiol. 2007;9:287–97.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    4.
    Canchaya C, Fournous G, Chibani-Chennoufi S, Dillmann ML, Brüssow H. Phage as agents of lateral gene transfer. Curr Opin Microbiol. 2003;6:417–24.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Howard-Varona C, Hargreaves KR, Solonenko NE, Markillie LM, White RA, Brewer HM, et al. Multiple mechanisms drive phage infection efficiency in nearly identical hosts. ISME J. 2018;12:1605–18.
    PubMed  PubMed Central  Article  Google Scholar 

    6.
    Weinbauer MG, Rassoulzadegan F. Are viruses driving microbial diversification and diversity? Environ Microbiol. 2004;6:1–11.
    PubMed  Article  PubMed Central  Google Scholar 

    7.
    Thurber RV. Current insights into phage biodiversity and biogeography. Curr Opin Microbiol. 2009;12:582–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    8.
    Chow C-ET, Suttle CA. Biogeography of viruses in the sea. Annu Rev Virol. 2015;2:41–66.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Roux S, Brum JR, Dutilh BE, Sunagawa S, Duhaime MB, Loy A, et al. Ecogenomics and potential biogeochemical impacts of globally abundant ocean viruses. Nature. 2016;537:689–93.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    10.
    Shkoporov AN, Khokhlova EV, Fitzgerald CB, Stockdale SR, Draper LA, Ross RP, et al. ΦCrAss001 represents the most abundant bacteriophage family in the human gut and infects Bacteroides intestinalis. Nat Commun. 2018;9:4781.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    11.
    Breitbart M, Miyake JH, Rohwer F. Global distribution of nearly identical phage-encoded DNA sequences. FEMS Microbiol Lett. 2004;236:249–56.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    12.
    Dutilh BE, Cassman N, McNair K, Sanchez SE, Silva GGZ, Boling L, et al. A highly abundant bacteriophage discovered in the unknown sequences of human faecal metagenomes. Nat Commun. 2014;5:4498.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    13.
    Jameson E, Mann NH, Joint I, Sambles C, Mühling M. The diversity of cyanomyovirus populations along a North-South Atlantic Ocean transect. ISME J. 2011;5:1713–21.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    14.
    Delong EF, Preston CM, Mincer T, Rich V, Hallam SJ, Frigaard N, et al. Community genomics among stratified microbial assemblages in the ocean’s interior. Science. 2006;311:496–503.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    15.
    Finke JF, Suttle CA. The environment and cyanophage diversity: insights from environmental sequencing of DNA polymerase. Front Microbiol. 2019;10:167.
    PubMed  PubMed Central  Article  Google Scholar 

    16.
    Hanson CA, Marston MF, Martiny JB. Biogeographic variation in host range phenotypes and taxonomic composition of marine cyanophage isolates. Front Microbiol. 2016;7:983.
    PubMed  PubMed Central  Article  Google Scholar 

    17.
    Huang S, Zhang S, Jiao N, Chen F. Marine cyanophages demonstrate biogeographic patterns throughout the global ocean. Appl Environ Microbiol. 2015;81:441–52.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    18.
    Marston MF, Taylor S, Sme N, Parsons RJ, Noyes TJE, Martiny JBH. Marine cyanophages exhibit local and regional biogeography. Environ Microbiol. 2013;15:1452–63.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    19.
    Paez-Espino D, Eloe-Fadrosh EA, Pavlopoulos GA, Thomas AD, Huntemann M, Mikhailova N, et al. Uncovering Earth’s virome. Nature. 2016;536:425–30.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    20.
    Winter C, Matthews B, Suttle CA. Effects of environmental variation and spatial distance on bacteria, archaea and viruses in sub-polar and arctic waters. ISME J. 2013;7:1507–18.
    PubMed  PubMed Central  Article  Google Scholar 

    21.
    Luo E, Aylward FO, Mende DR, Delong EF. Bacteriophage distributions and temporal variability in the ocean’s interior. mBio 2017;8:e01903–17.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    22.
    Brum JR, Ignacio-espinoza JC, Roux S, Doulcier G, Acinas SG, Alberti A, et al. Patterns and ecological drivers of ocean viral communities. Science. 2015;348:1261498.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    23.
    Dennehy JJ. What ecologists can tell virologists. Annu Rev Microbiol. 2014;68:117–35.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    24.
    Held NL, Whitaker RJ. Viral biogeography revealed by signatures in Sulfolobus islandicus genomes. Environ Microbiol. 2009;11:457–66.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    25.
    Ashby B, Boots M. Multi-mode fluctuating selection in host–parasite coevolution. Ecol Lett. 2017;20:357–65.
    PubMed  Article  PubMed Central  Google Scholar 

    26.
    Koskella B, Brockhurst MA. Bacteria-phage coevolution as a driver of ecological and evolutionary processes in microbial communities. FEMS Microbiol Rev. 2014;38:916–31.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    27.
    Vos M, Birkett PJ, Birch E, Griffiths RI, Buckling A. Local adaptation of bacteriophages to their bacterial hosts in soil. Science 2009;325:833.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    28.
    Gomez P, Buckling A. Coevolution with phages does not influence the evolution of bacterial mutation rates in soil. ISME J. 2013;7:2242–4.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    29.
    Kraemer SA, Boynton PJ. Evidence for microbial local adaptation in nature. Mol Ecol. 2017;26:1860–76.
    PubMed  Article  PubMed Central  Google Scholar 

    30.
    Kawecki T, Ebert D. Conceptual issues in local adaptation. Ecol Lett. 2004;7:1225–41.
    Article  Google Scholar 

    31.
    Lenormand T. Gene flow and the limits to natural selection. Trends Ecol Evol. 2002;17:183–9.
    Article  Google Scholar 

    32.
    Nosil P, Egan SP, Funk DJ. Heterogeneous genomic differentiation between walking-stick ecotypes: “isolation by adaptation” and multiple roles for divergent selection. Evolution. 2008;62:316–36.
    PubMed  Article  Google Scholar 

    33.
    Orsini L, Vanoverbeke J, Swillen I, Mergeay J, De Meester L. Drivers of population genetic differentiation in the wild: Isolation by dispersal limitation, isolation by adaptation and isolation by colonization. Mol Ecol. 2013;22:5983–99.
    PubMed  Article  Google Scholar 

    34.
    Zhang Q-G, Buckling A. Migration highways and migration barriers created by host–parasite interactions. Ecol Lett. 2016;19:1479–85.
    PubMed  Article  Google Scholar 

    35.
    Wang IJ, Bradburd GS. Isolation by environment. Mol Ecol. 2014;23:5649–62.
    PubMed  Article  Google Scholar 

    36.
    Buckling A, Rainey PB. Antagonistic coevolution between a bacterium and a bacteriophage. Proc Biol Sci. 2002;269:931–6.
    PubMed  PubMed Central  Article  Google Scholar 

    37.
    Kunin V, He S, Warnecke F, Peterson SB, Garcia Martin H, Haynes M, et al. A bacterial metapopulation adapts locally to phage predation despite global dispersal. Genome Res. 2008;18:293–7.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    38.
    Lopez Pascua L, Gandon S, Buckling A. Abiotic heterogeneity drives parasite local adaptation in coevolving bacteria and phages. J Evol Biol. 2012;25:187–95.
    CAS  PubMed  Article  Google Scholar 

    39.
    Baumann P. Biology of endosymbionts of plant sap-sucking insects. Annu Rev Microbiol. 2005;59:155–89.
    CAS  PubMed  Article  Google Scholar 

    40.
    Levy A, Gonzalez IS, Mittelviefhaus M, Clingenpeel S, Paredes SH, Miao J, et al. Genomic features of bacterial adaptation to plants. Nat Genet. 2018;50:138–50.
    CAS  Article  Google Scholar 

    41.
    Bäckhed F, Ley RE, Sonnenburg JL, Peterson DA, Gordon JI. Host-bacterial mutualism in the human intestine. Science 2005;307:1915–20.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    42.
    Heath KD, Tiffin P. Context dependence in the coevolution of plant and rhizobial mutualists. Proc Biol Sci. 2007;274:1905–12.
    PubMed  PubMed Central  Google Scholar 

    43.
    Koch M, Delmotte N, Rehrauer H, Vorholt JA, Pessi G, Hennecke H. Rhizobial adaptation to hosts, a new facet in the legume root-nodule symbiosis. Mol Plant Microbe Interact. 2010;23:784–90.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    44.
    Aguilar OM, Riva O, Peltzer E. Analysis of Rhizobium etli and of its symbiosis with wild Phaseolus vulgaris supports coevolution in centers of host diversification. Proc Natl Acad Sci. 2004;101:13548–53.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    45.
    Bitocchi E, Bellucci E, Giardini A, Rau D, Rodriguez M, Biagetti E, et al. Molecular analysis of the parallel domestication of the common bean (Phaseolus vulgaris) in Mesoamerica and the Andes. N Phytol. 2013;197:300–13.
    CAS  Article  Google Scholar 

    46.
    Koenig R, Gepts P. Allozyme diversity in wild Phaseolus vulgaris: further evidence for two major centers of genetic diversity. Theor Appl Genet. 1989;78:809–17.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    47.
    Melkonian R, Moulin L, Béna G, Tisseyre P, Chaintreuil C, Heulin K, et al. The geographical patterns of symbiont diversity in the invasive legume Mimosa pudica can be explained by the competitiveness of its symbionts and by the host genotype. Environ Microbiol. 2014;16:2099–111.
    PubMed  Article  PubMed Central  Google Scholar 

    48.
    Tian CF, Young JPW, Wang ET, Tamimi SM, Chen WX. Population mixing of Rhizobium leguminosarum bv. viciae nodulating Vicia faba: the role of recombination and lateral gene transfer. FEMS Microbiol Ecol. 2010;73:563–76.
    CAS  PubMed  PubMed Central  Google Scholar 

    49.
    Burdon JJ, Thrall PH. Spatial and temporal patterns in coevolving plant and pathogen associations. Am Nat. 1999;153:S15–S33.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    50.
    Van Cauwenberghe J, Visch W, Michiels J, Honnay O. Selection mosaics differentiate Rhizobium-host plant interactions across nitrogen environments. Oikos 2016;125:1755–61.
    Article  Google Scholar 

    51.
    Guimarães PR, Pires MM, Jordano P, Bascompte J, Thompson JN. Indirect effects drive coevolution in mutualistic networks. Nature 2017;550:511–4.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    52.
    Heath KD, Lau JA. Herbivores alter the fitness benefits of a plant–rhizobium mutualism. Acta Oecol. 2011;37:87–92.
    Article  Google Scholar 

    53.
    Rogers HS, Buhle ER, HilleRisLambers J, Fricke EC, Miller RH, Tewksbury JJ. Effects of an invasive predator cascade to plants via mutualism disruption. Nat Commun. 2017;8:6–13.
    Article  CAS  Google Scholar 

    54.
    Delmas E, Besson M, Brice MH, Burkle LA, Dalla Riva GV, Fortin MJ, et al. Analysing ecological networks of species interactions. Biol Rev. 2019;94:16–36.
    Article  Google Scholar 

    55.
    Gaiarsa MP, Guimarães PR. Interaction strength promotes robustness against cascading effects in mutualistic networks. Sci Rep. 2019;9:1–7.
    CAS  Article  Google Scholar 

    56.
    Sih A, Crowley P, McPeek M, Petranka J, Strohmeier K. Predation, competition, and prey communities: a review of field experiments. Annu Rev Ecol Syst. 1985;16:269–311.
    Article  Google Scholar 

    57.
    Parratt SR, Barrès B, Penczykowski RM, Laine AL. Local adaptation at higher trophic levels: contrasting hyperparasite–pathogen infection dynamics in the field and laboratory. Mol Ecol. 2017;26:1964–79.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    58.
    Hatcher MJ, Dick JTA, Dunn AM. How parasites affect interactions between competitors and predators. Ecol Lett. 2006;9:1253–71.
    PubMed  Article  PubMed Central  Google Scholar 

    59.
    Hutchinson MC, Bramon Mora B, Pilosof S, Barner AK, Kéfi S, Thébault E, et al. Seeing the forest for the trees: putting multilayer networks to work for community ecology. Funct Ecol. 2019;33:206–17.
    Article  Google Scholar 

    60.
    Koskella B, Taylor TB. Multifaceted impacts of bacteriophages in the plant microbiome. Annu Rev Phytopathol. 2018;56:361–80.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    61.
    Labrie SJ, Samson JE, Moineau S. Bacteriophage resistance mechanisms. Nat Rev Microbiol. 2010;8:317–27.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    62.
    Evans TJ, Ind A, Komitopoulou E, Salmond GPC. Phage-selected lipopolysaccharide mutants of Pectobacterium atrosepticum exhibit different impacts on virulence. J Appl Microbiol. 2010;109:505–14.
    CAS  PubMed  PubMed Central  Google Scholar 

    63.
    Perez Carrascal OM, Vaninsberghe D, Juárez S, Polz MF. Population genomics of the symbiotic plasmids of sympatric nitrogen-fixing Rhizobium species associated with Phaseolus vulgaris. Environ Microbiol. 2016;18:2660–76.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    64.
    Santamaría RI, Bustos P, Sepúlveda-Robles O, Lozano L, Rodríguez C, Fernández JL, et al. Narrow-host-range bacteriophages that infect Rhizobium etli associate with distinct genomic types. Appl Environ Microbiol. 2014;80:446–54.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    65.
    Carlson K. Working with bacteriophages: common techniques and methodological approaches. In: Kutter E, Sulakvelidze A (eds). Bacteriophages: biology and applications. Boca Raton, FL: CRC Press; 2005). p. 437–94.

    66.
    Werle E, Schneider C, Renner M, Völker M, Fiehn W. Convenient single-step, one tube purification of PCR products for direct sequencing. Nucleic Acids Res. 1994;22:4354–5.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    67.
    Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004;32:1792–7.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    68.
    Bolger AM, Lohse M, Usadel B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014;30:2114–20.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    69.
    Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its applications to single-Cell sequencing. J Comput Biol. 2012;19:455–77.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    70.
    Zerbino DR, Birney E. Velvet: Algorithms for de novo short read assembly using de Bruijn graphs. Genome Res. 2008;18:821–9.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    71.
    Gordon D, Green P. Consed: a graphical editor for next-generation sequencing. Bioinformatics 2013;29:2936–7.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    72.
    Chaudhari NM, Gupta VK, Dutta C. BPGA- an ultra-fast pan-genome analysis pipeline. Sci Rep. 2016;6:24373.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    73.
    Untergasser A, Cutcutache I, Koressaar T, Ye J, Faircloth BC, Remm M, et al. Primer3 — new capabilities and interfaces. Nucleic Acids Res. 2012;40:e115.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    74.
    Richter M, Rosselló-Móra R. Shifting the genomic gold standard for the prokaryotic species definition. Proc Natl Acad Sci. 2009;106:19126–31.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    75.
    Pritchard L, Glover RH, Humphris S, Elphinstone JG, Toth IK. Genomics and taxonomy in diagnostics for food security: soft-rotting enterobacterial plant pathogens. Anal Methods. 2016;8:12–14.
    Article  Google Scholar 

    76.
    Lopes A, Tavares P, Petit M, Guérois R, Zinn-justin S. Automated classification of tailed bacteriophages according to their neck organization. BMC Genom. 2014;15:1027.
    Article  CAS  Google Scholar 

    77.
    Hyman P, Abedon ST. Phage host range and efficiency of plating. In: Clokie MRJ, Kropinski AM (eds). Bacteriophages, methods and protocols. Vol. I: Isolation, characterization, and interactions. Totowa, NJ: Humana Press; 2009. p. 175–202.

    78.
    Hyman P, Abedon ST. Bacteriophage host range and bacterial resistance. Adv Appl Microbiol. 2010;70:217–48.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    79.
    Holmfeldt K, Solonenko N, Howard-Varona C, Moreno M, Malmstrom RR, Blow MJ, et al. Large-scale maps of variable infection efficiencies in aquatic Bacteroidetes phage-host model systems. Environ Microbiol. 2016;18:3949–61.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    80.
    Ishizawa H, Kuroda M, Morikawa M, Ike M. Evaluation of environmental bacterial communities as a factor affecting the growth of duckweed Lemna minor. Biotechnol Biofuels. 2017;10:1–10.
    Article  CAS  Google Scholar 

    81.
    Cenens W, Makumi A, Mebrhatu MT, Lavigne R, Aertsen A. Phage–host interactions during pseudolysogeny. Bacteriophage 2013;3:e25029.
    PubMed  PubMed Central  Article  Google Scholar 

    82.
    Kauffman KM, Hussain FA, Yang J, Arevalo P, Brown JM, Chang WK, et al. A major lineage of non-tailed dsDNA viruses as unrecognized killers of marine bacteria. Nature. 2018;554:118–22.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    83.
    Oksanen J, Blanchet FG, Friendly M, Kindt R, Legendre P, Glinn D, et al. Community Ecology Package. https://cran.r-project.org, https://github.com/vegandevs/vegan. 2019.

    84.
    Flores CO, Poisot T, Valverde S, Weitz JS. BiMat: a MATLAB package to facilitate the analysis of bipartite networks. Methods Ecol Evol. 2016;7:127–32.
    Article  Google Scholar 

    85.
    Consul PC. A simple urn model dependent on predetermined strategy. Sankhyā Indian J Stat Ser B. 1974;36:391–9.
    Google Scholar 

    86.
    Borcard D, Legendre P. All-scale spatial analysis of ecological data by means of principal coordinates of neighbour matrices. Ecol Modell. 2002;153:51–68.
    Article  Google Scholar 

    87.
    Flores CO, Valverde S, Weitz JS. Multi-scale structure and geographic drivers of cross-infection within marine bacteria and phages. ISME J. 2013;7:520–32.
    PubMed  Article  PubMed Central  Google Scholar 

    88.
    Porter SS, Chang PL, Conow CA, Dunham JP, Friesen ML. Association mapping reveals novel serpentine adaptation gene clusters in a population of symbiotic Mesorhizobium. ISME J. 2016;11:248–62.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    89.
    Greenlon A, Chang PL, Damtew ZM, Muleta A, Carrasquilla-Garcia N, Kim D, et al. Global-level population genomics reveals differential effects of geography and phylogeny on horizontal gene transfer in soil bacteria. Proc Natl Acad Sci. 2019;116:15200–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    90.
    Scola V, Ramond JB, Frossard A, Zablocki O, Adriaenssens EM, Johnson RM, et al. Namib desert soil microbial community diversity, assembly, and function along a natural xeric gradient. Micro Ecol. 2018;75:193–203.
    CAS  Article  Google Scholar 

    91.
    Short CM, Suttle CA. Nearly identical bacteriophage structural gene sequences are widely distributed in both marine and freshwater environments. Appl Environ Microbiol. 2005;71:480–6.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    92.
    Edwards RA, Vega AA, Norman HM, Ohaeri M, Levi K, Dinsdale EA, et al. Global phylogeography and ancient evolution of the widespread human gut virus crAssphage. Nat Microbiol. 2019;4:1727–36.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    93.
    Culley AI, Steward GF. New genera of RNA viruses in subtropical seawater, inferred from polymerase gene sequences. Appl Environ Microbiol. 2007;73:5937–44.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    94.
    Miranda-Sánchez F, Rivera J, Vinuesa P. Diversity patterns of Rhizobiaceae communities inhabiting soils, root surfaces and nodules reveal a strong selection of rhizobial partners by legumes. Environ Microbiol. 2016;18:2375–91.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    95.
    Bontemps C, Rogel MA, Wiechmann A, Mussabekova A, Moody S, Simon MF, et al. Endemic Mimosa species from Mexico prefer alphaproteobacterial rhizobial symbionts. N Phytol. 2016;209:319–33.
    CAS  Article  Google Scholar 

    96.
    Van Cauwenberghe J, Lemaire B, Stefan A, Efrose R, Michiels J, Honnay O. Symbiont abundance is more important than pre-infection partner choice in a Rhizobium – legume mutualism. Syst Appl Microbiol. 2016;39:345–9.
    PubMed  Article  PubMed Central  Google Scholar 

    97.
    Van Cauwenberghe J, Michiels J, Honnay O. Effects of local environmental variables and geographical location on the genetic diversity and composition of Rhizobium leguminosarum nodulating Vicia cracca populations. Soil Biol Biochem. 2015;90:71–9.
    Article  CAS  Google Scholar 

    98.
    Van Cauwenberghe J, Verstraete B, Lemaire B, Lievens B, Michiels J, Honnay O. Population structure of root nodulating Rhizobium leguminosarum in Vicia cracca populations at local to regional geographic scales. Syst Appl Microbiol. 2014;37:613–21.
    PubMed  Article  PubMed Central  Google Scholar 

    99.
    Hurwitz BL, Brum JR, Sullivan MB. Depth-stratified functional and taxonomic niche specialization in the ‘core’ and ‘flexible’ Pacific Ocean Virome. ISME J. 2015;9:472–84.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    100.
    Mühling M, Fuller NJ, Millard A, Somerfield PJ, Marie D, Wilson WH, et al. Genetic diversity of marine Synechococcus and co-occurring cyanophage communities: evidence for viral control of phytoplankton. Environ Microbiol. 2005;7:499–508.
    PubMed  Article  PubMed Central  Google Scholar 

    101.
    Sun Y, Zhang S, Long L, Dong J, Chen F, Huang S. Genetic diversity and cooccurrence patterns of marine cyanopodoviruses and picocyanobacteria. Appl Environ Microbiol. 2018;84:e00591–18.
    CAS  PubMed  PubMed Central  Google Scholar 

    102.
    Chase AB, Arevalo P, Brodie EL, Polz MF, Karaoz U, Martiny JBH. Maintenance of sympatric and allopatric populations in free-living terrestrial bacteria. mBio. 2019;10:e02361–19.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    103.
    Flores CO, Meyer JR, Valverde S, Farr L, Weitz JS. Statistical structure of host – phage interactions. Proc Natl Acad Sci. 2011;108:E288.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    104.
    Koskella B, Thompson JN, Preston GM, Buckling A. Local biotic environment shapes the spatial scale of bacteriophage adaptation to bacteria. Am Nat. 2011;177:440–51.
    PubMed  Article  PubMed Central  Google Scholar 

    105.
    Koskella B, Parr N. The evolution of bacterial resistance against bacteriophages in the horse chestnut phyllosphere is general across both space and time. Philos Trans R Soc B Biol Sci. 2015;370:20140297.
    Article  Google Scholar 

    106.
    Morgan AD, Gandon S, Buckling A. The effect of migration on local adaptation in a coevolving host-parasite system. Nature 2005;437:253–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    107.
    Gómez P, Paterson S, De Meester L, Liu X, Lenzi L, Sharma MD, et al. Local adaptation of a bacterium is as important as its presence in structuring a natural microbial community. Nat Commun. 2016;7:12453.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    108.
    Zhang Q-G, Buckling A. Resource-dependent antagonistic coevolution leads to a new paradox of enrichment. Ecology 2016;97:1319–28.
    PubMed  Article  PubMed Central  Google Scholar 

    109.
    Lopez-Pascua LDC, Buckling A. Increasing productivity accelerates host-parasite coevolution. J Evol Biol. 2008;21:853–60.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    110.
    Gurney J, Aldakak L, Betts A, Gougat-Barbera C, Poisot T, Kaltz O, et al. Network structure and local adaptation in co-evolving bacteria–phage interactions. Mol Ecol. 2017;26:1764–77.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    111.
    Thompson JN. The geographic mosaic of coevolution. Chicago, IL: Uni. Chicago Press; 2005. More

  • in

    Metabolomic signatures of coral bleaching history

    1.
    LaJeunesse, T. C. et al. Systematic revision of Symbiodiniaceae highlights the antiquity and diversity of coral endosymbionts. Curr. Biol. 28, 2570–2580 (2018).
    CAS  PubMed  Google Scholar 
    2.
    Muscatine, L. & Porter, J. W. Reef corals: mutualistic symbioses adapted to nutrient-poor environments. BioScience 27, 454–460 (1977).
    Google Scholar 

    3.
    van Hooidonk, R., Maynard, J. A. & Planes, S. Temporary refugia for coral reefs in a warming world. Nat. Clim. Change 3, 508–511 (2013).
    Google Scholar 

    4.
    National Academies of Sciences, Engineering, and Medicine A Research Review of Interventions to Increase the Persistence and Resilience of Coral Reefs (The National Academies Press, 2019); https://doi.org/10.17226/25279

    5.
    Barshis, D. J. et al. Genomic basis for coral resilience to climate change. Proc. Natl Acad. Sci. USA 110, 1387–1392 (2013).
    CAS  PubMed  Google Scholar 

    6.
    Palumbi, S. R., Barshis, D. J., Traylor-Knowles, N. & Bay, R. A. Mechanisms of reef coral resistance to future climate change. Science 344, 895–898 (2014).
    CAS  PubMed  Google Scholar 

    7.
    Bay, R. & Palumbi, S. Rapid acclimation ability mediated by transcriptome changes in reef-building corals. Genome Biol. Evol. 7, 1602–1612 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    8.
    Grottoli, A. G. et al. Coral physiology and microbiome dynamics under combined warming and ocean acidification. PLoS ONE 13, e0191156 (2018).
    PubMed  PubMed Central  Google Scholar 

    9.
    Ziegler, M., Seneca, F., Yum, L. & P, S.-N. Bacterial community dynamics are linked to patterns of coral heat tolerance. Nat. Commun. 8, 14213 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    10.
    Hillyer, K. E. et al. 13C metabolomics reveals widespread change in carbon fate during coral bleaching. Metabolomics 14, 12 (2018).
    Google Scholar 

    11.
    Hillyer, K. E. et al. Metabolite profiling of symbiont and host during thermal stress and bleaching in the coral Acropora aspera. Coral Reefs 36, 105–118 (2017).
    Google Scholar 

    12.
    Sogin, E. M., Putnam, H., Gates, R. D., Putnam, H. M. & Anderson, P. E. Metabolomic signatures of increases in temperature and ocean acidification from the reef-building coral Pocillopora damicornis. Metablomics 12, 71 (2016).
    Google Scholar 

    13.
    Hillyer, K. E., Tumanov, S., Villas-Bô As, S. & Davy, S. K. Metabolite profiling of symbiont and host during thermal stress and bleaching in a model cnidarian-dinoflagellate symbiosis. J. Exp. Biol. https://doi.org/10.1242/jeb.128660 (2016).

    14.
    Fisch, J., Drury, C., Towle, E. K., Winter, R. N. & Miller, M. W. Physiological and reproductive repercussions of consecutive summer bleaching events of the threatened Caribbean coral Orbicella faveolata. Coral Reefs 38, 863–876 (2019).
    Google Scholar 

    15.
    Pinzón, J. H. et al. Whole transcriptome analysis reveals changes in expression of immune-related genes during and after bleaching in a reef-building coral. R. Soc. Open Sci. 2, 140214 (2015).
    PubMed  PubMed Central  Google Scholar 

    16.
    Thomas, L. & Palumbi, S. R. The genomics of recovery from coral bleaching. Proc. R. Soc. B 284, 20171790 (2017).
    PubMed  Google Scholar 

    17.
    Wall, C. B. et al. Shifting baselines: repeat bleaching drives coral physiotypes through environmental legacy and cellular memory. Preprint at bioRxiv https://doi.org/10.1101/2020.04.23.056457 (2020).

    18.
    Matsuda, S. et al. Coral bleaching susceptibility is predictive of subsequent mortality within but not between coral species. Front. Ecol. Evol. 8, 178 (2020).
    Google Scholar 

    19.
    Howells, E. J., Abrego, D., Meyer, E., Kirk, N. L. & Burt, J. A. Host adaptation and unexpected symbiont partners enable reef-building corals to tolerate extreme temperatures. Glob. Change Biol. 22, 2702–2714 (2016).
    Google Scholar 

    20.
    van Oppen, M. J. H. et al. Shifting paradigms in restoration of the world’s coral reefs. Glob. Change Biol. 23, 3437–3448 (2017).
    Google Scholar 

    21.
    Anthony, K. R. N. et al. Operationalizing resilience for adaptive coral reef management under global environmental change. Glob. Change Biol. 21, 48–61 (2015).
    Google Scholar 

    22.
    da Silva, R. R., Lopes, N. P. & Silva, D. B. in Mass Spectrometry in Chemical Biology: Evolving Applications (eds da Silva, R. R. & Lopes, N. P.) 57–81 (Royal Society of Chemistry, 2017).

    23.
    Cunning, R., Ritson-Williams, R. & Gates, R. Patterns of bleaching and recovery of Montipora capitata in Kāne’ohe Bay, Hawai’i, USA. Mar. Ecol. Prog. Ser. 551, 131–139 (2016).
    CAS  Google Scholar 

    24.
    Sumner, L. W. et al. Proposed minimum reporting standards for chemical analysis: Chemical Analysis Working Group (CAWG) Metabolomics Standards Initiative (MSI). Metabolomics 3, 211–221 (2007).
    CAS  PubMed  PubMed Central  Google Scholar 

    25.
    Rosset, S. et al. Lipidome analysis of Symbiodiniaceae reveals possible mechanisms of heat stress tolerance in reef coral symbionts. Coral Reefs 38, 1241–1253 (2019).
    Google Scholar 

    26.
    Li, Y. et al. Simultaneous structural identification of diacylglyceryl-N-trimethylhomoserine (DGTS) and diacylglycerylhydroxymethyl-N,N,N-trimethyl-β-alanine (DGTA) in microalgae using dual Li+/H+ adduct ion mode by ultra-performance liquid chromatography/quadrupole time‐of‐flight mass spectrometry. Rapid Commun. Mass Spectrom. 31, 457–468 (2017).
    CAS  PubMed  Google Scholar 

    27.
    Matthews, J. L. et al. Optimal nutrient exchange and immune responses operate in partner specificity in the cnidarian–dinoflagellate symbiosis. Proc. Natl Acad. Sci. USA 114, 13194–13199 (2017).
    CAS  PubMed  Google Scholar 

    28.
    Weis, V. M. Cellular mechanisms of cnidarian bleaching: stress causes the collapse of symbiosis. J. Exp. Biol. 211, 3059–3066 (2008).
    CAS  PubMed  Google Scholar 

    29.
    Mansour, J. S., Pollock, F. J., Díaz-Almeyda, E., Iglesias-Prieto, R. & Medina, M. Intra- and interspecific variation and phenotypic plasticity in thylakoid membrane properties across two Symbiodinium clades. Coral Reefs 37, 841–850 (2018).
    Google Scholar 

    30.
    Roach, T. N. F. et al. A multiomic analysis of in situ coral–turf algal interactions. Proc. Natl Acad. Sci. USA 117, 13588–13595 (2020).
    CAS  PubMed  Google Scholar 

    31.
    Quinn, R. A. et al. Metabolomics of reef benthic interactions reveals a bioactive lipid involved in coral defence. Proc. R. Soc. B 283, 20160469 (2016).
    PubMed  Google Scholar 

    32.
    Rosset, S., Wiedenmann, J., Reed, A. J. & D’Angelo, C. Phosphate deficiency promotes coral bleaching and is reflected by the ultrastructure of symbiotic dinoflagellates. Mar. Pollut. Bull. 118, 180–187 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    33.
    Galtier d’Auriac, I. et al. Before platelets: the production of platelet-activating factor during growth and stress in a basal marine organism. Proc. R. Soc. B 285, 20181307 (2018).
    PubMed  Google Scholar 

    34.
    Quistad, S. D. et al. Evolution of TNF-induced apoptosis reveals 550 My of functional conservation. Proc. Natl Acad. Sci. USA 111, 9567–9572 (2014).
    CAS  PubMed  Google Scholar 

    35.
    Williams, A. et al. Metabolomic shifts associated with heat stress in coral holobionts. Sci. Adv. 7, eabd4210 (2021).
    PubMed Central  Google Scholar 

    36.
    Takahashi, N. Chemistry of Plant Hormones (CRC, 1986).

    37.
    Reyes, F., Martín, R. & Fernández, R. Granulatamides A and B, cytotoxic tryptamine derivatives from the soft coral Eunicella granulata. J. Nat. Prod. 69, 668–670 (2006).
    CAS  PubMed  Google Scholar 

    38.
    Hill, R., Larkum, A. W. & Kramer, D. Light-induced dissociation of antenna complexes in the symbionts of scleractinian corals correlates with sensitivity to coral bleaching. Coral Reefs 31, 963–975 (2012).
    Google Scholar 

    39.
    Venn, A. A., Wilson, M. A., Trapido-Rosenthal, H. G., Keely, B. J. & Douglas, A. E. The impact of coral bleaching on the pigment profile of the symbiotic alga, Symbiodinium. Plant Cell Environ. 29, 2133–2142 (2006).
    CAS  PubMed  Google Scholar 

    40.
    Martin, F. J. et al. A top-down systems biology view of microbiome–mammalian metabolic interactions in a mouse model. Mol. Syst. Biol. 3, 112 (2007).
    PubMed  PubMed Central  Google Scholar 

    41.
    Quinn, R. A. et al. Global chemical effects of the microbiome include new bile-acid conjugations. Nature 579, 123–129 (2020).
    CAS  PubMed  PubMed Central  Google Scholar 

    42.
    Wikoff, W. R. et al. Metabolomics analysis reveals large effects of gut microflora on mammalian blood metabolites. Proc. Natl Acad. Sci. USA 106, 3698–3703 (2009).
    CAS  PubMed  Google Scholar 

    43.
    Dixon, G., Abbott, E. & Matz, M. Meta-analysis of the coral environmental stress response: Acropora corals show opposing responses depending on stress intensity. Mol. Ecol. https://doi.org/10.1111/mec.15535 (2020).

    44.
    Boström-Einarsson, L. et al. Coral restoration – a systematic review of current methods, successes, failures and future directions. PLoS ONE 15, e0226631 (2020).
    PubMed  PubMed Central  Google Scholar 

    45.
    Van Oppen, M. J. H., Oliver, J. K., Putnam, H. M. & Gates, R. D. Building coral reef resilience through assisted evolution. Proc. Natl Acad. Sci. USA 112, 2307–2313 (2015).
    PubMed  Google Scholar 

    46.
    Baums, I. B. et al. Considerations for maximizing the adaptive potential of restored coral populations in the western Atlantic. Ecol. Appl. 29, e01978 (2019).
    PubMed  PubMed Central  Google Scholar 

    47.
    Bay, R., Rose, N., Logan, C. & Palumbi, S. Genomic models predict successful coral adaptation if future ocean warming rates are reduced. Sci. Adv. 3, e1701413 (2017).
    PubMed  PubMed Central  Google Scholar 

    48.
    Dührkop, K. et al. SIRIUS 4: a rapid tool for turning tandem mass spectra into metabolite structure information. Nat. Methods 16, 299–302 (2019).
    PubMed  Google Scholar 

    49.
    Pluskal, T., Castillo, S., Villar-Briones, A. & Orešič, M. MZmine 2: modular framework for processing, visualizing, and analyzing mass spectrometry-based molecular profile data. BMC Bioinform. 11, 395 (2010).
    Google Scholar 

    50.
    Wang, M. et al. Sharing and community curation of mass spectrometry data with Global Natural Products Social Molecular Networking. Nat. Biotechnol. 34, 828–837 (2016).
    CAS  PubMed  PubMed Central  Google Scholar 

    51.
    Nothias, L.-F. et al. Feature-based molecular networking in the GNPS analysis environment. Nat. Methods 17, 905–908 (2020).
    CAS  PubMed  Google Scholar 

    52.
    Martin, C. et al. Viscosin-like lipopeptides from frog skin bacteria inhibit Aspergillus fumigatus and Batrachochytrium dendrobatidis detected by imaging mass spectrometry. Sci. Rep. 9, 3019 (2019).
    Google Scholar 

    53.
    Cunning, R., Gillette, P., Capo, T., Galvez, K. & Baker, A. C. Growth tradeoffs associated with thermotolerant symbionts in the coral Pocillopora damicornis are lost in warmer oceans. Coral Reefs 34, 155–160 (2015).
    Google Scholar 

    54.
    Cunning, R. & Baker, A. C. Excess algal symbionts increase the susceptibility of reef corals to bleaching. Nat. Clim. Change 3, 259–262 (2013).
    Google Scholar  More

  • in

    Adaptive evolution in a conifer hybrid zone is driven by a mosaic of recently introgressed and background genetic variants

    1.
    Abbott, R. D. et al. Hybridization and speciation. J. Evol. Bio. 26, 229–246 (2013).
    CAS  Article  Google Scholar 
    2.
    de Lafontaine, G. & Bousquet, J. Asymmetry matters: a genomic assessment of directional biases in gene flow between hybridizing spruces. Ecol. Evol. 7, 3883–3893 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    3.
    Todesco, M. et al. Hybridization and extinction. Evol. Appl. 9, 892–908 (2016).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    4.
    Anderson, E. & Stebbins, G. L. Hybridization as an evolutionary stimulus. Evolution 8, 378–388 (1954).
    Article  Google Scholar 

    5.
    De La Torre, A. R., Li, Z., Van de Peer, Y. & Ingvarsson, P. K. Contrasting rates of molecular evolution and patterns of selection among gymnosperms and flowering plants. Mol. Bio. Evol. 34, 1363–1377 (2017).
    Article  CAS  Google Scholar 

    6.
    Critchfield, W. B. Hybridization and classification of the white pines (Pinus section Strobus). Taxon 35, 647–656 (1986).
    Article  Google Scholar 

    7.
    Nystedt, B. et al. The Norway spruce genome sequence and conifer genome evolution. Nature 497, 579–584 (2013).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    8.
    Bouille, M. & Bousquet, J. Trans-species shared polymorphisms at orthologous nuclear gene loci among distant species in the conifer Picea (Pinaceae): implications for long term maintenance of genetic diversity in trees. Am. J. Bot. 92, 63–73 (2005).
    PubMed  Article  PubMed Central  Google Scholar 

    9.
    Hamilton, J. A., Lexer, C. & Aitken, S. N. Genomic and phenotypic architecture of a spruce hybrid zone (Picea sitchensis × P. glauca). Mol. Ecol. 22, 827–841 (2013).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    10.
    Hamilton, J. & Miller, J. Adaptive introgression as a resource for management and genetic conservation in a changing climate. Conserv. Biol. 30, 33–41 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    11.
    Jagoda, E. et al. Disentangling immediate adaptive introgression from selection on standing introgressed variation in humans. Mol. Biol. Evol. 35, 623–630 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    12.
    Bresadola, L. et al. Admixture mapping in interspecific Populus hybrids identifies classes of genomic architectures for phytochemical, morphological and growth traits. N. Phytol. 223, 2076–2089 (2019).
    CAS  Article  Google Scholar 

    13.
    Suarez-Gonzalez, A. et al. Genomic and functional approaches reveal a case of adaptive introgression from Populus balsamifera (balsam poplar) in P. trichocarpa (black cottonwood). Mol. Ecol. 25, 2427–2442 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    14.
    Suarez-Gonzalez, A., Hefer, C. A., Lexer, C., Cronk, Q. C. & Douglas, C. J. Scale and direction of adaptive introgression between black cottonwood (Populus trichocarpa) and balsam poplar (P. balsamifera). Mol. Ecol. 27, 1667–1680 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    15.
    Leroy, T. et al. Adaptive introgression as a driver of local adaptation to climate in European white oaks. New. Phytol. 226, 1171–1182 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    16.
    Hufford, M.B. et al. Genomic signature of crop-wild introgression in Maize. PLoS Genet. 9, e100347 (2013).
    Article  Google Scholar 

    17.
    Ma, Y. et al. Ancient introgression drives adaptation to cooler and drier mountain habitats in a cypress species complex. Commun. Biol. 18, 210–213 (2019).
    Google Scholar 

    18.
    Pyhäjärvi, T., Hufford, M. B., Mezmouk, S. & Ross-Ibarra, J. Complex patterns of local adaptation in Teosinte. Genome Biol. Evol. 5, 1594–1609 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    19.
    Mei, W., Stetter, M. G. & Stitzer, M. C. Adaptation in plant genomes: bigger is different. Am. J. Bot. 105, 16–19 (2019).
    Article  Google Scholar 

    20.
    Syring, J. et al. Widespread genealogical non-monophyly in species of the Pinus subgenus. Strobus. Syst. Biol. 56, 163–181 (2007).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    21.
    Menon, M. et al. The role of hybridization during ecological divergence of southwestern white pine (Pinus strobiformis) and limber pine (P. flexilis). Mol. Ecol. 27, 1245–1260 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    22.
    Looney, C. E. & Waring, K. M. Pinus strobiformis (southwestern white pine) stand dynamics, regeneration, and disturbance ecology: a review. For. Ecol. Manag. 287, 90–102 (2013).
    Article  Google Scholar 

    23.
    Schoettle, A. W. & Rochelle, S. G. Morphological variation of Pinus flexilis (Pinaceae), a bird-dispersed pine, across a range of elevations. Am. J. Bot. 87, 1797–1806 (2000).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    24.
    Frankis, M. P. The high altitude white pines (Pinus L. subgenus Strobus Lemmon, Pinaceae) of Mexico and the adjacent SW USA. Int. Dendrol. Soc. Yearb. 2008, 64–72 (2009).
    Google Scholar 

    25.
    Tomback, D. F. et al. Seed dispersal in limber and southwestern white pine: comparing core and peripheral populations. In The Future of High Elevation, Five-Needle White Pines in Western North America: Proceedings of the High Five Symposium. Proceedings RMRS-P- 63 69–71 (US Department of Agriculture, Forest Service, Rocky Mountain Research Station, Fort Collins, CO, 2011).

    26.
    Moreno-Letelier, A., Ortíz-Medrano, A. & Piñero, D. Niche divergence versus neutral processes: combined environmental and genetic analyses identify contrasting patterns of differentiation in recently diverged pine species. PLoS ONE 8, e78228 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    27.
    Moreno-Letelier, A. & Barraclough, T. G. Mosaic genetic differentiation along environmental and geographic gradients indicate divergent selection in a white pine species complex. Evol. Ecol. 29, 733–748 (2015).
    Article  Google Scholar 

    28.
    Little, E. L. Jr. Atlas of United States Trees. Vol. 5, 22 (Florida. Misc. Publ. 1361, U.S. Department of Agriculture, Forest Service, 1978).

    29.
    Bisbee, J. Cone morphology of the Pinus ayacahuite-flexilis complex of the southwestern United States and Mexico. Bull. Cupressus Conserv. Proj. 3, 3–33 (2014).
    Google Scholar 

    30.
    Borgman, E. M., Schoettle, A. W. & Angert, A. L. Assessing the potential for maladaptation during active management of limber pine populations: a common garden study detects genetic differentiation in response to soil moisture in the Southern Rocky Mountains. Can. J. For. Res. 45, 496–505 (2015).
    CAS  Article  Google Scholar 

    31.
    Neale, D. B. & Kremer, A. Forest tree genomics: growing resources and applications. Nat. Rev. Genet. 12, 111–122 (2011).
    CAS  PubMed  Article  Google Scholar 

    32.
    Mitton, J., Kreiser, B. R. & Latta, R. G. Glacial refugia of limber pine (Pinus flexilis James) inferred from the population structure of mitochondrial DNA. Mol. Ecol. 9, 91–97 (2000).
    CAS  PubMed  Article  Google Scholar 

    33.
    Jorgensen, S., Hamrick, J. L. & Wells, P. V. Regional patterns of genetic diversity in Pinus flexilis (Pinaceae) reveal complex species history. Am. J. Bot. 89, 792–800 (2002).
    PubMed  Article  Google Scholar 

    34.
    Goodrich, B. A., Waring, K. M. & Kolb, T. E. Genetic variation in Pinus strobiformis growth and drought tolerance from southwestern US populations. Tree Physiol. 36, 1219–1235 (2016).
    CAS  PubMed  Article  Google Scholar 

    35.
    DaBell, J. Pinus Strobiformis Response to an Elevational Gradient and Correlation with Source Climate. Master’s thesis, Northern Arizona University (2017).

    36.
    Francis, J. A. & Vavrus, S. J. Evidence for a wavier jet stream in response to rapid Arctic warming. Environ. Res. Lett. 10, 014005 (2015).
    Article  Google Scholar 

    37.
    Rellstab, C. et al. A practical guide to environmental association analysis in landscape genomics. Mol. Ecol. 24, 4348–4370 (2015).
    PubMed  Article  PubMed Central  Google Scholar 

    38.
    Coop, G., Witonsky, D., Di Rienzo, A. & Pritchard, J. K. Using environmental correlations to identify loci underlying local adaptation. Genetics 185, 1411–1423 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    39.
    Levitt J. Responses of Plants to Environmental Stress. Chilling, Freezing, and High Temperature Stresses 2nd edn (Academic Press, 1980).

    40.
    Bierne, N., Welch, J., Loire, E., Bonhomme, F. & David, P. The coupling hypothesis: why genome scans may fail to map local adaptation genes. Mol. Ecol. 20, 2044–2072 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    41.
    Zuur, A. F., Ieno, E. N. & Elphick, C. S. A protocol for data exploration to avoid common statistical problems. Methods Ecol. 1, 3–14 (2010).
    Article  Google Scholar 

    42.
    Harrison, K. A. et al. Signatures of polygenic adaptation associated with climate across the range of a threatened fish species with high genetic connectivity. Mol. Ecol. 26, 6253–6269 (2017).
    Article  Google Scholar 

    43.
    Lind, B. M. et al. Water availability drives signatures of local adaptation in whitebark pine (Pinus albicaulis Engelm.) across fine spatial scales of the Lake Tahoe Basin, USA. Mol. Ecol. 26, 3168–3185 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    44.
    Csillery, K. et al. Detecting short spatial scale local adaptation and epistatic selection in climate‐related candidate genes in European beech (Fagus sylvatica) populations. Mol. Ecol. 23, 4696–4708 (2014).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    45.
    Schumer, M. & Brandvain, Y. Determining epistatic selection in admixed populations. Mol. Ecol. 25, 2577–2591 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    46.
    Menon, M. et al. Tracing the footprints of a moving hybrid zone under a demographic history of speciation with gene flow. Evol. Appl. 13, 195–209 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    47.
    Whitney, K. D. et al. Quantitative trait locus mapping identifies candidate alleles involved in adaptive introgression and range expansion in a wild sunflower. Mol. Ecol. 24, 2194–2211 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    48.
    Chhatre, V. E., Evan, L. M., DiFazio, S. P. & Keller, S. R. Adaptive introgression and maintenance of a trispecies hybrid complex in range‐edge populations of Populus. Mol. Ecol. 27, 4820–4838 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    49.
    Aitken, S. A. et al. Adaptation, migration or extirpation: climate change outcomes for tree populations. Evol. Appl. 1, 95–111 (2008).
    PubMed  PubMed Central  Article  Google Scholar 

    50.
    Alberto, F. J. et al. Potential for evolutionary responses to climate change—evidence from tree populations. Glob. Chn. Bio. 19, 1645–1661 (2013).
    Article  Google Scholar 

    51.
    Kirkpatrick, M. & Barton, N. H. Evolution of a species’ range. Am. Nat. 150, 1–23 (1997).
    CAS  PubMed  Article  Google Scholar 

    52.
    Stebbins, G. L. The role of hybridization in evolution. Proc. Am. Philos. Soc. 103, 231–251 (1959).
    Google Scholar 

    53.
    Petit, R. J. & Excoffier, L. Gene flow and species delimitation. Trends Ecol. Evol. 24, 386–393 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    54.
    Barton, N. H. & Hewitt, G. M. Analysis of hybrid zones. Annu. Rev. Ecol. Evol. S 16, 113–148 (1985).
    Article  Google Scholar 

    55.
    Mimura, M., Mishima, M., Lascoux, M. & Yahara, T. Range shift and introgression of the rear and leading populations in two ecologically distinct Rubus species. BMC Evol. Biol. 2014, 209 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    56.
    De La Torre, A. R., Wang, T., Jaquish, B. & Aitken, S. N. Adaptation and exogenous selection in a Picea glauca × Picea engelmannii hybrid zone: implications for forest management under climate change. N. Phytol. 201, 687–699 (2014).
    Article  CAS  Google Scholar 

    57.
    Hamilton, J. R., De La Torre, A. R. & Aitken, S. N. Fine-scale environmental variation contributes to introgression in a three-species spruce hybrid complex. Tree Genet. Genomes 11, 1–14 (2015).
    Article  Google Scholar 

    58.
    Fraïsse, C. K., Belkhir, J., Welch, J. & Bierne, N. Local interspecies introgression is the main cause of extreme levels of intraspecific differentiation in mussels. Mol. Ecol. 25, 269–770 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    59.
    Wu, D. D. et al. Pervasive introgression facilitated domestication and adaptation in the Bos species complex. Nat. Ecol. Evol. 2, 1139–1145 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    60.
    Kremer, A. & Le Corre, V. Decoupling of differentiation between traits and their underlying genes in response to divergent selection. Heredity 10, 375–385 (2012).
    Article  Google Scholar 

    61.
    Eckert, A. J. et al. Local adaptation at fine spatial scales: an example from sugar pine (Pinus lambertiana, Pinaceae). Tree Genet. Genomes 11, 42 (2015).
    Article  Google Scholar 

    62.
    Hornoy, B., Pavy, N., Gérardi, S., Beaulieu, J. & Bousquet, J. Genetic adaptation to climate in white spruce involves small to moderate allele frequency shifts in functionally diverse genes. Genome Biol. Evol. 7, 3269–3285 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    63.
    Rieseberg, L. H. et al. Hybridization and the colonization of novel habitats by annual sunflowers. Genetica 129, 149–165 (2007).
    PubMed  Article  PubMed Central  Google Scholar 

    64.
    Lewontin, R. C. & Birch, L. C. Hybridization as a source of variation for adaptation to new environments. Evolution 20, 315–336 (1966).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    65.
    Pavy, N. et al. The heterogeneous levels of linkage disequilibrium in white spruce genes and comparative analysis with other conifers. Heredity 108, 273–284 (2011).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    66.
    Kim, B. Y., Huber, C. D. & Lohmueller, K. Deleterious variation shapes the genomic landscape of introgression. PLoS Genet. 14, e1007741 (2018).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    67.
    Eyre-Walker, A., Woolfit, M. & Phelps, T. The distribution of fitness effects of new deleterious amino acid mutations in humans. Genetics 173, 891–900 (2006).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    68.
    Christe, C. et al. Adaptive evolution and segregating load contribute to the genomic landscape of divergence in two tree species connected by episodic gene flow. Mol. Ecol. 26, 59–76 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    69.
    Lu, M., Hodgins, K. A., Degner, J. C. & Yeaman, S. Purifying selection does not drive signatures of convergent local adaptation of lodgepole pine and interior spruce. BMC Evol. Biol. 19, 110 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    70.
    Whitlock, M. C. Temporal fluctuations in demographic parameters and the genetic variance among populations. Evolution 46, 608–615 (1992).
    PubMed  Article  Google Scholar 

    71.
    Lexer, C. et al. Genomic admixture analysis in European Populus spp. reveals unexpected patterns of reproductive isolation and mating. Genetics 186, 699–712 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    72.
    Lowry, D. et al. Breaking RAD: an evaluation of the utility of restriction site-associated DNA sequencing for genome scans of adaptation. Mol. Ecol. Resour. 17, 142–152 (2017).
    CAS  PubMed  Article  Google Scholar 

    73.
    Parchman, T. L. et al. RADseq approaches and applications for forest tree genetics. Tree Genet. Genomes 14, 39 (2018).
    Article  Google Scholar 

    74.
    Gossmann, T. I., Keightley, P. D. & Eyre-Walker, A. The effect of variation in the effective population size on the rate of adaptive molecular evolution in eukaryotes. Genome Biol. Evol. 4, 658–667 (2012).
    PubMed  PubMed Central  Article  Google Scholar 

    75.
    Lexer, C. & Widmer, A. The genic view of plant speciation: recent progress and emerging questions. Philos. Trans. R. Soc. Lond. B Biol. Sci. 363, 3023–3036 (2008).
    PubMed  PubMed Central  Article  Google Scholar 

    76.
    Bucholz, E. Early Growth, Water Relations and Growth: Common Garden Studies of Pinus Strobiformis under Climate Change. PhD dissertation, Northern Arizona University (2020).

    77.
    Lotterhos, K. & Whitlock, M. The relative power of genome scans to detect local adaptation depends on sampling design and statistical method. Mol. Ecol. 24, 1031–1046 (2015).
    PubMed  Article  Google Scholar 

    78.
    Skotte, L., Korneliussen, T. S. & Albrechtsen, A. Estimating individual admixture proportions from next generation sequencing data. Genetics 195, 693–702 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    79.
    Goudet, J. hierfstat, a package for R to compute and test hierarchical F-statistics. Mol. Ecol. Notes 5, 184–186 (2005).
    Article  Google Scholar 

    80.
    R Core Team. R v.3.3.2: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2017).

    81.
    Parchman, T. L. et al. Genome -wide association genetics of an adaptive trait in lodgepole pine: association mapping of serotiny. Mol. Ecol. 21, 2991–3005 (2012).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    82.
    Puritz, J. B., Hollenbeck, C. M. & Gold, J. R. dDocent: a RADseq, variant -calling pipeline designed for population genomics of non -model organisms. PeerJ 2, e431 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    83.
    Wang, T., Hamann, A., Spittlehouse, D. L. & Carroll, C. Locally downscaled and spatially customizable climate data for historical and future periods for North America. PLoS ONE 11, e0156720 (2016).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    84.
    Hengl, T. et al. SoilGrids1km—global soil information based on automated mapping. PLoS ONE 9, e105992 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    85.
    Günther, T. & Coop, G. Robust identification of local adaptation from allele frequencies. Genetics 195, 205–220 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    86.
    Brooks, S. P. & Gelman, A. General methods for monitoring convergence of iterative simulations. J. Comput. Graph. Stat. 7, 434–455 (1998).
    Google Scholar 

    87.
    Camacho et al. BLAST+: architecture and applications. BMC Bioinfo. 10, 421 (2009).
    Article  CAS  Google Scholar 

    88.
    Warnes, G., Gorjanc, G., Leisch, F. & Man, M. genetics: Population Genetics. R package version 1.3.8.1 (2013).

    89.
    Oksanen, J. et al. vegan: Community Ecology Package. R package version 2.5-2 (2013).

    90.
    Legendre, P. & Legendre, L. Numerical Ecology 2nd English edn (Elsevier, 1998).

    91.
    Montgomery, D. C. & Peck, E. A. Introduction to Linear Regression Analysis 2nd edn (John Wiley & Sons, 1992).

    92.
    Liu, Q. Variation partitioning by partial redundancy analysis (RDA). Environmetrics 8, 75–85 (1997).
    CAS  Article  Google Scholar 

    93.
    Kemppainen, P. et al. Linkage disequilibrium network analysis (LDna) gives a global view of chromosomal inversions, local adaptation and geographic structure. Mol. Ecol. Resour. 15, 1031–1045 (2015).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    94.
    Ohta, T. Linkage disequilibrium with the island model. Genetics 101, 139–155 (1982).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    95.
    Beissinger, T. M. et al. Using the variability of linkage disequilibrium between subpopulations to infer sweeps and epistatic selection in a diverse panel of chickens. Heredity 116, 58–166 (2015).
    Google Scholar 

    96.
    Hijmans, R. J. geosphere: Spherical trigonometry. R package version 1.5‐7 (2017).

    97.
    Adamack, A. T. & Gruber, B. PopGenReport: simplifying basic population genetic analyses in R. Methods Ecol. Evol. 5, 384–387 (2014).
    Article  Google Scholar 

    98.
    Gompert, Z. & Buerkle, A. C. introgress: methods for analyzing introgression between divergent lineages. R package version 1.2.3 (2012).

    99.
    Gompert, Z. & Buerkle, C. A. A powerful regression-based method for admixture mapping of isolation across the genome of hybrids. Mol. Ecol. 18, 1207–1224 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    100.
    Janoušek, V. et al. Genome‐wide architecture of reproductive isolation in a naturally occurring hybrid zone between Mus musculus musculus and M. m. domesticus. Mol. Ecol. 21, 3032–3047 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    101.
    Hancock, A. M. et al. Adaptation to climate across the Arabidopsis thaliana genome. Science 334, 83–86 (2011).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    102.
    Menon, M. et al. Data from: adaptive evolution in a confier hybrid zone is driven by a mosaic of recently introgressed and background genetic variants. Figshare, Dataset https://doi.org/10.6084/m9.figshare.c.5130104 (2020).

    103.
    Shirk, A. J. et al. Southwestern white pine (Pinus strobiformis) species distribution models predict large range shift and contraction due to climate change. For. Ecol. Manag. 411, 176–186 (2018).
    Article  Google Scholar 

    104.
    Little, E. L., Jr. Atlas of United States Trees, Vol. 1., Conifers and important hardwoods. Misc. Publ. 1146, 320 (U.S. Department of Agriculture, Forest Service, 1971).

    105.
    Menon, M. et al. Code from: adaptive evolution in a confier hybrid zone is driven by a mosaic of recently introgressed and background genetic variants. Zenodo, Dataset https://doi.org/10.5281/zenodo.4054085 (2020). More

  • in

    C-STABILITY an innovative modeling framework to leverage the continuous representation of organic matter

    C-STABILITY description
    Organic matter representation
    The description of SOM in C-STABILITY consists of several subdivisions. First, organic matter is separated in two main pools, one for living microbes (noted Cmic) and one for the substrate (noted Csub) (Fig. 1a). Several groups of living microbes can be considered simultaneously (e.g., bacteria, fungi, etc.) and C-STABILITY classes them into functional communities. Second, SOM is also separated between several biochemical classes, e.g., cellulose (or plant sugar), lignin, lipid, protein, and microbial sugar in this study (Fig. 1a). Third for each biochemical class, substrate accessible to its enzymes (noted ac) is separated from substrate which is inaccessible (noted in) due to specific physicochemical conditions, e.g., interaction between different molecules, inclusion in aggregates, sorption on mineral surfaces, etc.
    Polymerization is a driver of interactions between substrate and living microbes and a continuous description of the degree of organic matter polymerization (noted p) is provided for each of these pools, as a distribution (Fig. 1b). The polymerization axis is oriented from the lowest to the highest degree of polymerization. A right-sided distribution corresponds to a highly polymerized substrate whereas a left-sided distribution corresponds to monomer or small oligomer forms. For each biochemical class ∗ (∗ = cellulose, lignin, lipid, etc.), the polymerization range is identical for both accessible and inaccessible pools. The total amount of C (in gC) in the accessible and the inaccessible pools of any biochemical class is as follows:

    $${C}_{* }^{rm{ac}}=int_{{p}_{* }^{rm{min}}}^{{p}_{* }^{rm{max}}}{chi }_{* }^{rm{ac}}(p)dp,$$
    (1)

    $${C}_{* }^{rm{in}}=int_{{p}_{* }^{rm{min}}}^{{p}_{* }^{rm{max}}}{chi }_{* }^{rm{in}}(p)dp,$$
    (2)

    where ({p}_{* }^{rm{min}}) and ({p}_{* }^{rm{max}}) are the minimum and maximum degrees of polymerization of the biochemical class ∗ and ({chi }_{* }^{rm{ac}}), ({chi }_{* }^{rm{in}}) (gC.p−1) are the polymerization distributions. Finally, the total substrate C pool is defined as the sum of all biochemical pools,

    $${C}_{rm{sub}}=sum _{* }left({C}_{* }^{rm{in}}+{C}_{* }^{rm{ac}}right).$$
    (3)

    Accessibility to microbe uptake is described by the interval (also called domain) ({{mathcal{D}}}_{u}), which corresponds to small substrate compounds, monomers, dimers or trimers smaller than 600 Daltons1, that microbes are able to take up (in red in Fig. 1b). Besides, accessibility to enzymes occurs in the ({{mathcal{D}}}_{rm{enz}}) domain (in blue in Fig. 1b). Over time the substrate accessible to enzymes is depolymerized and its distribution shifts toward ({{mathcal{D}}}_{u}) where it eventually becomes accessible to microbe uptake.
    The numerical rules chosen to represent polymerization are as simple as possible in the context of theoretical simulations. Each pool is associated with a polymerization interval ([{p}_{* }^{rm{min}},{p}_{* }^{rm{max}}]) of length two. Initial substrate distributions are represented by Gaussian distributions centered at a relative distance of 25% from ({p}_{* }^{rm{max}}) (here 0.5), with the standard deviation set at 5% of polymerization interval length (here 0.1). In the accessible pool, the microbial uptake domains ({{mathcal{D}}}_{u}) are positioned at the left of the interval with a relative length of 20% (here 0.4), and enzymatic domains ({{mathcal{D}}}_{rm{enz}}) overlap the entire polymerization intervals.
    Organic matter dynamics
    As described in Fig. 1, three processes drive OM dynamics: (i) enzymatic activity, (ii) microbial uptake, biotransformation, and mortality, and (iii) changes in local physicochemical conditions. First, enzymes have a depolymerization role, which enables the transformation of highly polymerized substrate into fragments accessible to microbes. Second, microbial uptake of substrate is only possible for molecules having a very small degree of polymerization. When C is taken up, a fraction is respired and the remaining is metabolized, and biotransformed into microbial molecules that return to the substrate upon microbe death. Each microbial group has a specific signature that describes its composition in terms of biochemistry and polymerization. Third, changes in local substrate conditions drive exchanges between substrate accessible and inaccessible to enzymes (e.g., aggregate formation and break). All of these processes are considered with a daily time step (noted d).
    Enzymatic activity Enzymes are specific to biochemical classes. They are not individually reported, but rather as a family of enzymes contributing to the depolymerization of a biochemical substrate (e.g., combined action of endoglucanase, exoglucanase, betaglucosidase, etc., on cellulose will be reported as cellulolytic action). Figure 2 describes how substrate polymerization distributions are impacted by enzymes. The overall functioning of each enzyme family (noted enz) is described by two parameters: a depolymerization rate ({tau }_{rm{enz}}^{0}) providing the number of broken bonds per time unit and a factor accounting for the type of substrate cleavage αenz. The term ({F}_{rm{enz}}^{rm{act}}) (gC.p−1.d−1) represents the change in polymerization of ({chi }_{* }^{rm{ac}}) due to enzyme activity for all (pin {{mathcal{D}}}_{rm{enz}}),

    $${F}_{rm{enz}}^{rm{act}}({chi }_{* }^{rm{ac}},p,t)= -{tau }_{rm{enz}}(t){chi }_{* }^{rm{ac}}(p,t)\ +int_{{{mathcal{D}}}_{rm{enz}}}{{mathcal{K}}}_{rm{enz}}(p,p^{prime} ){tau }_{rm{enz}}(t){chi }_{* }^{rm{ac}}(p^{prime} ,t)dp^{prime}.$$
    (4)

    The depolymerization rate, τenz (d−1), is expressed as a linear function of microbial C biomass Cmic (gC),

    $${tau }_{rm{enz}}(t)={tau }_{rm{enz}}^{0}{C}_{rm{mic}}(t),$$
    (5)

    where ({tau }_{rm{enz}}^{0}) (g({,}_{C}^{-1}).d−1) is the action rate of a given enzyme per amount of microbial C. If several microbial communities are associated with the same enzyme family, we replace the Cmic term by a weighted sum of the C mass of all communities involved in Eq. (5). The ({{mathcal{K}}}_{rm{enz}}) (p−1) kernel provides the polymerization change from (p^{prime}) to p,

    $${{mathcal{K}}}_{rm{enz}}(p,p^{prime} )={{mathbb{1}}}_{ple p^{prime} }({alpha }_{rm{enz}}+1)frac{{(p-{p}_{* }^{rm{min}})}^{{alpha }_{rm{enz}}}}{{(p^{prime} -{p}_{* }^{rm{min}})}^{{alpha }_{rm{enz}}+1}},$$
    (6)

    where ({{mathbb{1}}}_{ple p^{prime} }) equals 1 if (ple p^{prime}) and 0 otherwise. The αenz cleavage factor denotes the enzyme efficiency to generate a large amount of small fragments and to shift the substrate polymerization distribution toward the microbe uptake domain ({{mathcal{D}}}_{u}) (Fig. 2). αenz = 1 is typical of the action of endo-cleaving enzymes, which randomly disrupts any bond of its polymeric substrate and generates oligomers. The shift toward ({{mathcal{D}}}_{u}) is slower if αenz increases. This is characteristic of exo-cleaving enzymes, which attack the end-members of their polymeric substrate, generate small fragments, and preserve highly polymerized compounds. To satisfy the mass balance, the kernel verifies (int {{mathcal{K}}}_{rm{enz}}(p,p^{prime} )dp=1). Then ({F}_{rm{enz}}^{rm{act}}) does not change the total C mass but only the polymerization distribution (i.e., (int {F}_{rm{enz}}^{rm{act}}(chi ,p,t)dp=0)).
    Microbial biotransformation Each microbial group (denoted mic) produces new organic compounds from the assimilated C. After death, the composition of the necromass returning to each biochemical pool ∗ of SOM is assumed to be constant, accessible and is depicted with a set of distributions smic,∗, named signature. Each distribution smic,* (p–1) describes the polymerization of the dead microbial compounds returning to the pool ∗. The signature is normalized and unitless to ensure mass conservation, i.e., if we note that,

    $${S}_{rm{mic},* }=mathop{int}nolimits_{{p}_{* }^{rm{min}}}^{{p}_{* }^{rm{max}}}{s}_{rm{mic},* }(p)dp,$$
    (7)

    then we have ∑*Smic,* = 1.
    For each accessible pool of substrate, the term ({F}_{rm{mic},* }^{rm{upt}}) describes how the microbes utilize the substrate available in the microbial uptake ({{mathcal{D}}}_{u}) domain (Fig. 1b). For all (pin {{mathcal{D}}}_{u}),

    $${F}_{rm{mic},* }^{rm{upt}}({chi }^{rm{ac}},p,t)={u}_{rm{mic},* }^{0}{C}_{rm{mic}}(t){chi }_{* }^{rm{ac}}(p,t),$$
    (8)

    where ({u}_{rm{mic},* }^{0}) (g({,}_{C}^{-1}).d−1) is the uptake rate per amount of microbe C. The substrate uptake rate linearly depends on the microbial C quantity.
    Depending on a carbon use efficiency parameter ({e}_{rm{mic},* }^{0}) (ratio between microbe assimilated C and taken up C), taken up C is respired or assimilated and biotransformed into microbial metabolites. This induces a change in the biochemistry and polymerization (Fig. 1a).
    Finally, microbial necromass returns to the substrate pools with a specific mortality, which linearly depends on the microbial C quantity,

    $${F}_{rm{mic},* }^{rm{nec}}(p,t)={m}_{rm{mic}}^{0}{C}_{rm{mic}}(t){s}_{rm{mic},* }(p),$$
    (9)

    where ({m}_{rm{mic}}^{0}) (d−1) is the mortality rate of the microbe (Fig. 1a).
    Change in local physicochemical conditions The polymerization of a substrate inaccessible to its enzymes remains unchanged over time. A specific event changing the accessibility to enzymes (e.g., aggregate disruption or desorption from mineral surfaces) is modeled with a flux from the inaccessible to the accessible pool. Transfer between these pools is described by the ({F}_{rm{ac},* }^{rm{loc}}) term for each biochemistry ∗,

    $${F}_{rm{ac},* }^{rm{loc}}(p,t)={tau }_{rm{tr}}^{rm{ac}}{chi }_{* }^{rm{in}}(p,t)$$
    (10)

    where ({tau }_{rm{tr},* }^{rm{ac}}) (d−1) is rate of local condition change toward accessibility.
    Transfer in the opposite way (e.g., aggregate formation, association with mineral surfaces) is described by the ({F}_{rm{in},* }^{rm{loc}}) term,

    $${F}_{rm{in},* }^{rm{loc}}(p,t)={tau }_{rm{tr}}^{rm{in}}{chi }_{* }^{rm{ac}}(p,t)$$
    (11)

    where ({tau }_{rm{tr},* }^{rm{in}}) (d−1) is rate of local condition change toward inaccessibility.
    Organic matter input We defined time dependent distributions for carbon input fluxes. There are denoted ({i}_{* }^{rm{ac}}) and ({i}_{* }^{rm{in}}) (gC.p−1.d−1) for both accessible and inaccessible pools of biochemical classes ∗. The total carbon input flux, expressed in gC.d−1, is:

    $$I(t)=sum _{* }int_{{p}_{* }^{rm{min}}}^{{p}_{* }^{rm{max}}}left({i}_{* }^{rm{in}}(p,t)+{i}_{* }^{rm{ac}}(p,t)right)dp.$$
    (12)

    General dynamics equations The distribution dynamics for each biochemical class * is obtained from Eqs. (4)–(6) and (8)–(11),

    $$frac{partial {chi }_{* }^{rm{ac}}}{partial t}(p,t)= , {F}_{rm{ac},* }^{loc}(p,t)-{F}_{rm{in},* }^{rm{loc}}(p,t)\ +{F}_{rm{enz}}^{rm{act}}({chi }_{* }^{rm{ac}},p,t) +sum _{{rm{mic}}}left({F}_{rm{mic},* }^{rm{nec}}(p,t)-{F}_{rm{mic},* }^{rm{upt}}({chi }^{rm{ac}},p,t)right) +{i}_{* }^{rm{ac}}(p,t),$$
    (13)

    $$frac{partial {chi }_{* }^{rm{in}}}{partial t}(p,t)={F}_{rm{in},* }^{rm{loc}}(p,t)-{F}_{rm{ac},* }^{rm{loc}}(p,t) +{i}_{* }^{rm{in}}(p,t).$$
    (14)

    Then, the expended equations are,

    $$frac{partial {chi }_{* }^{rm{ac}}}{partial t}(p,t)= , {tau }_{rm{tr},* }^{rm{ac}}{chi }_{* }^{rm{in}}(p,t)-{tau }_{rm{tr},* }^{rm{in}}{chi }_{* }^{rm{ac}}(p,t)\ -{tau }_{rm{enz}}^{0}sum _{,text{mic},}{C}_{rm{mic}}(t){chi }_{* }^{rm{ac}}(p,t)\ +{tau }_{rm{enz}}^{0}sum _{,text{mic},}{C}_{rm{mic}}(t)({alpha }_{rm{enz}}+1)\ int_{p}^{{p}_{* }^{rm{max}}}frac{{(p-{p}_{* }^{rm{min}})}^{{alpha }_{rm{enz}}}}{{(p^{prime} -{p}_{* }^{rm{min}})}^{{alpha }_{rm{enz}}+1}}{chi }_{* }^{rm{ac}}(p^{prime} ,t)dp^{prime} \ +sum _{,text{mic},}{C}_{rm{mic}}(t){m}_{rm{mic}}^{0}{s}_{rm{mic},* }(p)\ -sum _{,text{mic},}{C}_{rm{mic}}(t){{mathbb{1}}}_{{{mathcal{D}}}_{u}}(p){u}_{rm{mic},* }^{0}{chi }_{* }^{rm{ac}}(p,t)\ +{i}_{* }^{rm{ac}}(p,t),$$
    (15)

    $$frac{partial {chi }_{* }^{rm{in}}}{partial t}(p,t)={tau }_{rm{tr},* }^{rm{in}}{chi }_{* }^{rm{ac}}(p,t)-{tau }_{rm{tr},* }^{rm{ac}}{chi }_{* }^{rm{in}}(p,t)+{i}_{* }^{rm{in}}(p,t).$$
    (16)

    where ({{mathbb{1}}}_{{{mathcal{D}}}_{u}}(p)) equals 1 if (pin {{mathcal{D}}}_{u}) and 0 otherwise.
    The dynamics of the total substrate is ruled by,

    $$frac{d{C}_{rm{sub}}(t)}{dt}=sum _{* }int_{{p}_{* }^{rm{min}}}^{{p}_{* }^{rm{max}}}left(frac{partial {chi }_{* }^{rm{ac}}}{partial t}(p,t)+frac{partial {chi }_{* }^{rm{in}}}{partial t}(p,t)right)dp.$$
    (17)

    The dynamics of microbial Cmic is obtained by,

    $$frac{d{C}_{rm{mic}}}{dt}(t)=-{m}_{rm{mic}}^{0}{C}_{rm{mic}}(t) +sum _{* }{u}_{rm{mic},* }^{0}{e}_{rm{mic},* }^{0}{C}_{rm{mic}}(t)int_{{{mathcal{D}}}_{u}}{chi }_{* }^{rm{ac}}(p,t)dp,$$
    (18)

    and the CO2 flux (gC.d−1) produced by the microbes is given by,

    $${F}_{rm{CO}_{2}}(t)={C}_{rm{mic}}(t)sum _{* }{u}_{rm{mic},* }^{0}left(1-{e}_{rm{mic},* }^{0}right){int}_{{{mathcal{D}}}_{u}}{chi }_{* }^{rm{ac}}(p,t)dp.$$
    (19)

    Model implementation
    The model was implemented in the Julia© language63,64. An explicit finite difference scheme approximates the solutions of integro-differential equations with a Δt = 0.1d time step and a Δp = 0.01p polymerization step. Differential equations were solved with a Runge–Kutta method.
    Scenarios
    Scenario 1: cellulose decomposition kinetics and model sensitivity
    A first simulation was run to depict cellulose depolymerization and uptake by a decomposer community over one year (see parameters in Table 1). A global sensitivity analysis focusing on the residual cellulose variable was made to determine (i) the relative influence of parameters, and (ii) how parameters influence varies over time (Fig. 3c). A specific attention was given on enzymatic parameters (especially α) to verify the pertinence of their introduction in the model.
    We considered a specific method defined by Sobol for calculating sensitivity indices65. It provides the relative contribution of the model parameters to the total model variance, here at different times of the simulation. The method relies on the same principle as the analysis of variance. It was designed to decompose the variance of a model output according to the various degrees of interaction between the n uncertain parameters ({({x}_{i})}_{iin {1,n}}). Formally, by assuming that the parameter uncertainties are independent, the model output, denoted y, could be expressed as a sum of functions that take parameter interactions into account,

    $$y={f}_{0}+mathop{sum }limits_{i=1}^{n}{f}_{i}({x}_{i})+mathop{sum }limits_{{i,j=1}atop {ine j}}^{n}{f}_{i,j}({x}_{i},{x}_{j})+…+{f}_{1, , …, , n}({x}_{1},…,{x}_{n}).$$
    (20)

    Under independence assumptions between models parameters variability, model variance is:

    $${{mathbb{V}}ar}(y)= , mathop{sum }limits_{i=1}^{n}{{mathbb{V}}ar}({f}_{i}({x}_{i}))\ +mathop{sum }limits_{{i,j=1}atop {ine j}}^{n}{{mathbb{V}}ar}({f}_{i,j}({x}_{i},{x}_{j}))\ +…+{{mathbb{V}}ar}({f}_{1, , …, , n}({x}_{1},…,{x}_{n})).$$
    (21)

    This variance decomposition leads to the definition of several sensitivity indices. The first-order Sobol’s index of each parameter is,

    $${S}_{i}=frac{{{mathbb{V}}ar}({f}_{i}({x}_{i}))}{{{mathbb{V}}ar}(y)},$$
    (22)

    and higher order indices are defined by:

    $${S}_{i,j}=frac{{{mathbb{V}}ar}({f}_{i,j}({x}_{i},{x}_{j}))}{{{mathbb{V}}ar}(y)},$$
    (23)

    and so on. These indices are unique, with a value of 0–1 and their sum equals 1. Here we focused on Sobol’s first-order indices as they are usually sufficient to give a straightforward interpretation of the actual influence of different parameters66,67. We computed the sensitivity of the model outputs at several times of the simulation to highlight the role of model’s parameters at different phases. Figure 3c shows the normalized Sobol’s first-order indices to illustrate the relative influence of the model parameters on residual cellulose-C amount. Sobol’s indices were estimated using a Monte Carlo estimator of the variance68. This was performed for a small variation in parameter values (±5% uniform variability), by running 12,000 model simulations for the Monte Carlo sampling.
    Scenario 2: effect of substrate inaccessibility to enzyme on litter decomposition kinetics
    A simulation of lignocellulose (76% cellulose, 24% lignin) degradation was performed by taking into account peroxidases, which deconstruct the lignin polymer, and cellulases, which hydrolyze cellulose. The cellulose was initially embedded in lignin and inaccessible to cellulase. The lignolytic activity (peroxidases) induces a disentanglement of the cellulose from the lignocellulosic complex. Therefore, the action of peroxidases was seen as a change of cellulose physicochemical local conditions resulting in a progressive transfer to the accessible pool. This transfer was assumed to be linearly related to the activity of lignolytic enzymes in Eq. (10),

    $${tau }_{rm{tr,cell.}}^{rm{ac}}={tau }_{rm{tr,cell.}}^{rm{ac},0}int_{{{mathcal{D}}}_{rm{lig.}}}{tau }_{rm{lig.}}^{0}{C}_{rm{mic}}(t){chi }_{rm{lig.}}^{rm{ac}}(p,t)dp,$$
    (24)

    where ({{mathcal{D}}}_{rm{lig.}}) is the domain of lignolytic activity and where the ({tau }_{rm{tr,cell.}}^{ac,0}) coefficient is set at 13 g({,}_{C}^{-1}) for the current illustration.
    The simulation was performed over one year. Enzymatic and microbial parameters given in Table 1 were chosen to be closely in line with the litter decomposition and enzyme action observation16,37,69.
    Scenario 3: effect of community succession on C fluxes and substrate biochemistry
    We simulated the succession of two microbial functional communities, on the same previous lignocellulose, considering microbial residue recycling. The parameters (Table 1) were chosen according to the microbial community succession observations43,45,70. The first microbial community was specialized in plant substrate degradation, the second was specialized in the degradation of microbial residues. We referred to them as plant decomposers and microbial residue decomposers. Microbial residue decomposers were more competitive than plant decomposers because of their higher carbon use efficiency and lower mortality rate (Table 1). Both communities had the same biochemical signature, i.e., 50% polysaccharides, 30% lipids, and 20% proteins. We tested the impact of cheating as follows. Either uptake was impossible, i.e., u0 equaled 0 for the community not involved in enzyme production, or uptake was possible but at a lower rate than the enzyme producers because the substrate fragments were released in the vicinity of the enzyme producers (Table 1).
    Scenario 4: soil organic matter composition at steady state
    We resolved the analytic formulation of the C stock and chemistry at steady state under several assumptions. We only considered one microbial community, a continuous constant plant input I at a rate of 2.74.10−4 gC.cm−2.d−1 and microbial recycling47. To be able to explicitly calculate the steady state, we only considered accessible pools, then cellulose substrate was not embedded in lignin but directly accessible to cellulolytic enzymes (Fig. 6). Finally, we considered that C use efficiency (({e}_{rm{mic}}^{0})) and uptake (({u}_{rm{mic}}^{0})) parameters were identical for all biochemical classes. A full mathematical proof is given in Supplementary Note 3.
    At steady state, the amount of microbial carbon is,

    $${C}_{rm{mic}}=frac{{e}_{rm{mic}}^{0}I}{(1-{e}_{rm{mic}}^{0}){m}_{rm{mic}}^{0}}.$$
    (25)

    For each biochemical class ∗, we define ({p}_{* }^{u}) which verifies ({p}_{* }^{rm{min}} More

  • in

    Development of a robust protocol for the characterization of the pulmonary microbiota

    Many precautions should be taken to limit the modification of the commensal communities studied and the increase of interindividual variation not attributable to the experimental variables. The following factors can influence the human microbiota and should be considered when designing studies targeting the lung microbiota: the administration of antibiotics or neoadjuvant25,26,27,28, the size of the lesion, the type of surgical procedure, the type of pulmonary pathology under study, and living habits of patients (e.g., smoking status, physical exercise, buccal hygiene, alcohol consumption)29,30,31,32,33,34.
    A more exhaustive list of concomitant factors was pointed out by Carney et al.35. However, as the different fields of microbiota studies expand, it is likely that additional variables that can alter its composition will be uncovered. The molecular tools currently used to analyze the human microbiota do not have the power to discriminate the impact of that many factors over the microbial profiles. Whenever possible, patients selected for lung microbiota studies should be extensively screened so that they can be as similar as possible. Longitudinal studies could also minimize the impact of those variables, as the same patient, with similar concomitant factors through the study, would be compared to himself overtime.
    Tissue management steps should consider the contamination possibilities. In addition to the selection of a less contamination-prone procedure, such as thoracoscopic lobectomy, the manipulations and the instrument used in subsampling the excised organ should be taken into account. A combination of bleach and humid heat was chosen to sterilize the instruments used to sample the cancerous and healthy tissue as it was considered the most easily accessible method. The use of humid heat itself (autoclave) lacks the power to completely neutralize bacterial genomic DNA in solutions and on surfaces36. On the other hand, the utilization of bleach, or a chlorinated detergent, leads to the complete degradation of contaminating DNA on surfaces, such as benches and instruments37,38, but requires rinsing to avoid corrosion. Hence, combining both methods, soaking the instruments in bleach 1.6% for 10 min before rinsing with distilled water and autoclaving in a sterilization pouches, ensures a minimal amount of DNA has to be degraded by moist heat. The rest of the single-use equipment used was commercially sterilized with ionizing radiation.
    Healthy lung tissue was subsampled from the pulmonary lobe containing the tumor to ensure that the developed method could be used on a variety of lung tissue samples. It could also act as a control of non-pathologic microbiota to allow comparisons of cancerous and non-cancerous samples within the same subject, hence minimizing the impact on inter-individual microbiota variations. In fact, Riquelme et al. found that the gut microbiota has the capacity to specifically colonize pancreatic tissue8. Correspondingly, the use of adjacent pulmonary tissue to the tumor could help get better insights at a specific colonization of the tumor by lung bacteria. A 5 cm distance between the tumor and the healthy sample was ensured to minimize the potential effect of increased inflammation surrounding the tumor. Furthermore, the lung microbiota composition seems to vary dependently on the position and depth of the respiratory tract, even inside a same lobe39. The healthy tissue was collected in the same tierce of pulmonary depth (Supplementary Fig. 4) in an attempt to sample a microbial community that it would be as representative of non-pathologic microbiota in the tumoral region as possible.
    The homogenization of frozen and thawed pulmonary tissues was attempted and was unsuccessful, both with the use of only a 2.8 mm tungsten bead in the Retsch – MM301 mixer mill (30 beats/s, 20 min) or of the Fisherbrand 150 homogenizer with plastic probes (Fisher scientific, Pittsburg). The elasticity of the tissue or its frozen state make the mass nearly unbreakable. The use of the Liberase™ TM enzymatic cocktail (collagenase I & II, thermolysin) prior to the mechanical homogenization proved successful and a homogeneous suspension was obtained using the two-step homogenization protocol (Supplementary Fig. 3). Multiple ratios of liquid to mass of tissue were tested and 3 mL/g was found optimal, as it facilitates the homogenization without overly diluting to sample. A similar ratio of liquid to tissue was used in breast tissue microbiota study40. The samples were first thawed at 4 °C to reduce potential growth or degradation of microorganisms. The digestion was performed directly in the 50 mL collection tube to limit the tissue manipulation and ensure possible contaminant tracking.
    Our team was also unable to replicate the results obtain by Yu et al. on larger tissue samples using 0.2 mg/mL of Proteinase K for 24 h13. The samples remained firm and turned brown. Using the Liberase™ cocktail enabled a much faster digestion (75 min) and broke down specifically the lung component responsible for its elasticity, the collagen.
    Three commercially available DNA extraction kits were tested. They were selected for their previous successful use in the study of pulmonary or gut microbiota and their intended application as described by the manufacturer. The extraction kits were first tested on homogenized lung tissue spiked with whole-cell bacterial community to assess the efficiency of DNA extraction and recuperation of the commercial kits. The three kits were able to recover more than 88% of the genera added to the samples. All the genera that were not detected by the Microbial and Powersoil (Cutibacterium acnes, Bacteroides vulgatus, Bifidobacterium adolescentis, D. radiodurans, Clostridium beijerinckii, L. gasseri), with the exception of H. pylori, were Gram-positive bacteria. This type of bacteria has been reported to require more aggressive extraction methods to break their tougher cell walls19. However, the bacterial community did not go through the enzymatic and physical homogenization that usually takes place before DNA extraction since we needed to obtain a homogenous tissue sample that could be processed with or without spiked bacteria. These hard to lyse Gram-positive bacteria could have been fragilized by these processes, rendering them easier to break down during the extraction protocol. Furthermore, the detection of the artificially incorporated bacteria does not account for the natural physical association that may occur between the human tissue and microbial cells. Nonetheless, these high percentages of recovery were promising and lead us to continue with the characterization of the extraction kits in a real-life context, meaning the analysis of the base-level microbiota in pulmonary samples collected and processed through the entire pipeline.
    Every measurement of the efficiency of extraction, including DNA yield (Supplementary Fig. 5), DNA purity (Supplementary Figs. 6 and 7), and alpha diversity (Fig. 1), pointed in the same direction. In fact, they all showed that the Blood extraction kit was the best option out of the three kits. Therefore, using the Blood kit is recommended as one of the pieces of a complete study design. Additionally, the presence of a high concentration of host DNA in tissue samples might tend to saturate the purification column, which could reduce to amount of bacterial DNA recovered. The superior DNA binding capacity of the affinity column of the Blood kit compared to the two others could explain its better performance and its higher yield in most cases. The samples extracted with the Blood kit were also associated with higher alpha diversity (Shannon index). Therefore, this extraction method was able to recover a higher number of different bacterial organisms (richness) and proportionality in the different OTUs (evenness). The absence of PCR inhibitors and a higher recuperation rate of bacterial DNA in the Blood extracted samples could have led to a more proper amplification in the sequencing process and to the recuperation of very low abundance bacterial DNA in the extraction eluate. For further research, it is advised to take the additional precaution of working under a biosafety cabinet or in the sterile field when analyzing the microbiota of lung tissues to reduce the risk of incorporation of airborne contaminants.
    The Illumina Miseq sequencing platform with the use of dual-index strategy has become the dominant technology used in microbial ecology studies for its cost efficiency, low error rate, and user-friendliness41,42,43. Most studies interested in the pulmonary microbiota have also used this technology11,13,14. The sequencing of the 16S rRNA gene amplicon was favored over a shotgun sequencing method because of the overwhelming quantity of human DNA joining bacterial genomes in the pulmonary tissue. The 16S rRNA gene is the most used marker of bacterial identification. No consensus has been reached on the selection of the 16S rRNA gene variable region (V) to sequence for human microbiota18,44. However, it should be kept consistent across studies to allow comparisons. Targeting the V3–V4 regions was suggested using the universal primers developed by Klindworth et al.45. Several microbiota studies, including lung microbiota, have also used these regions7,13,46,47,48.
    In the context of this study, genomic mock-community was spiked in DNA extracted from the pulmonary tissue at a biological meaningful concentration. Every genus added to the samples was successfully detected. Consequently, the high ratio of human DNA to bacterial DNA did not interfere with the amplification and detection steps of the sequencing procedure. The sequencing method in place seems adequate for its application in the characterization of pulmonary microbiota.
    Contaminating bacteria or DNA can have an important impact of the microbial profile observed in very low biomass samples such as pulmonary tissue23. Consequentially, in addition to proper protocol selection, methodological design that attempts to follow, detect, and account for contamination was proposed. Its main features include the incorporation of a single negative control that monitors the incorporation of contaminants at every step of the experimental method (Supplementary Fig. 3). Since every step of the protocol prior to the extraction is meant to be executed in a single tube and only by the addition of reagents, it is possible to carry and detect the contaminants introduced throughout the procedure. On the contrary, microbiota study methodologies usually dictate for the incorporation of multiple controls at every step of the procedure (e.g. DNA extraction kit, PCR controls, etc.)18. Although more informative as to which step leads to contamination, it makes data analysis harder since the presence of contamination in the multiple controls cannot by added.
    No bioinformatics standard operating procedure is available and what should be done with controls sequencing data is still under debate18. Some research groups tried to use a neutral community model49, additional qPCR data50, amplicon DNA yield, or prevalence algorithms51 to assess the influence of methodological contaminants. The removal of every bacterial OTU found in controls from the samples is often not appropriate as these OTUs might also be naturally present in the samples22. We propose using relative abundance ratio between samples and controls to remove contaminating OTUs. Since controls have much lower richness than extracted lung samples and that the total number of reads (sequencing depth) is distributed across every OTU, the relative abundance of reads for each OTU tend to be much higher in the control than the same OTU in samples. Therefore, if the relative abundance of an OTU is greatly superior in the sample than in the control, it is reasonable to think that the same OTU was also in the sample in a substantial quantity. To ensure that OTUs that were present in very low absolute abundance (e.g., from only 1–2 reads) do not lead to the removal of the highly abundant corresponding OTU in samples, only the OTUs with a ratio of 1000 (relative abundance of sample/relative abundance of sample) were kept. The rest of the OTUs found in controls were completely removed from the related samples, since the influence of contaminating DNA could not be differentiated from the pulmonary microbiota. This method would theoretically tolerate no more than 20 reads (0.1%) before removing the entire OTU from the sample if only one OTU was present in the samples (20,000 reads, 100%). The use of relative abundance helps reduce the absolute abundance bias induced by the divergence in sequencing depth. The OTUs were removed from both tissues at the same time or not at all to avoid adding artificial intraindividual variation. The authors acknowledge that the proposed contaminant management method does not have the in-dept validation of other methods, such as described by Davis et al. with the decontam package51. However, it does not share its limitations regarding the lack of consideration for OTU abundance and need of high number of controls to ensure sensitivity while using prevalence-based detection. Further research focused on the development of statistical methods to detect contaminant OTUs in the cases of lung microbiota is needed. This work is to be a starting point toward methodological standardization and its modular nature makes the bioinformatic contaminant management method proposed here interchangeable once a more robust one is uncovered.
    Pearson’s correlation tests were performed on the number of reads per OTU between the samples and their respective controls. Although these values were not normally distributed (Shapiro-Wilk, p  More

  • in

    Rats show direct reciprocity when interacting with multiple partners

    1.
    Lehmann, L. & Keller, L. The evolution of cooperation and altruism—a general framework and a classification of models. J. Evol. Biol. 19, 1365–1376 (2006).
    CAS  PubMed  Article  PubMed Central  Google Scholar 
    2.
    Taborsky, M., Frommen, J. G. & Riehl, C. Correlated pay-offs are key to cooperation. Philos. Trans. R. Soc. B Biol. Sci. 371, 20150084 (2016).
    Article  Google Scholar 

    3.
    Trivers, R. L. The evolution of altruism. Q. Rev. Biol. 46, 35–57 (1971).
    Article  Google Scholar 

    4.
    Axelrod, R. & Hamilton, W. D. The evolution of cooperation. Science (80-. ) 211, 1390–1396 (1981).
    ADS  MathSciNet  CAS  MATH  Article  Google Scholar 

    5.
    Pfeiffer, T., Rutte, C., Killingback, T., Taborsky, M. & Bonhoeffer, S. Evolution of cooperation by generalized reciprocity. Proc. R. Soc. B Biol. Sci. 272, 1115–1120 (2005).
    Article  Google Scholar 

    6.
    Rankin, D. J. & Taborsky, M. Assortment and the evolution of generalized reciprocity. Evolution (N. Y). 63, 1913–1922 (2009).

    7.
    Alexander, R. D. The Biology of Moral Systems (Aldine Gruyter, New York, 1987).
    Google Scholar 

    8.
    Nowak, M. A. & Sigmund, K. Evolution of indirect reciprocity by image scoring. Nature 393, 573 (1998).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Winkler, I., Jonas, K. & Rudolph, U. On the usefulness of memory skills in social interactions: Modifying the iterated Prisoner’s Dilemma. J. Conflict Resolut. 52, 375–384 (2008).
    Article  Google Scholar 

    10.
    Stevens, J. R., Volstorf, J., Schooler, L. J. & Rieskamp, J. Forgetting constrains the emergence of cooperative decision strategies. Front. Psychol. 1, 1–12 (2011).
    Article  Google Scholar 

    11.
    Milinski, M. & Wedekind, C. Working memory constrains human cooperation in the Prisoner’s Dilemma. Proc. Natl. Acad. Sci. 95, 13755–13758 (1998).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    12.
    Volstorf, J., Rieskamp, J. & Stevens, J. R. The good, the bad, and the rare: Memory for partners in social interactions. PLoS One 6, e18945 (2011).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    13.
    Barta, Z. et al. Optimal moult strategies in migratory birds. Philos. Trans. R. Soc. B Biol. Sci. 363, 211–229 (2008).
    Article  Google Scholar 

    14.
    Houston, A. I. & McNamara, J. M. Models of Adaptive Behaviour: An Approach Based on State (Cambridge University Press, Cambridge, 1999).
    Google Scholar 

    15.
    Tinbergen, N. The Study of Instinct (Clarendon Press, Oxford, 1951).
    Google Scholar 

    16.
    McNamara, J. M. & Houston, A. I. Integrating function and mechanism. Trends Ecol. Evol. 24, 670–675 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    17.
    Gigerenzer, G., Todd, P. M. & ABC Research Group. Simple heuristics that make us smart. (Oxford University Press, 1999).

    18.
    Hurley, S. Social heuristics that make us smarter. Philos. Psychol. 18, 585–612 (2005).
    Article  Google Scholar 

    19.
    Isen, A. M. Positive affect, cognitive processes, and social behavior. Adv. Exp. Soc. Psychol. 20, 203–253 (1987).
    Google Scholar 

    20.
    Bartlett, M. Y. & DeSteno, D. Gratitude and prosocial behavior: Helping when it costs you. Psychol. Sci. 17, 319–325 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    21.
    Stanca, L. Measuring indirect reciprocity: Whose back do we scratch? J. Econ. Psychol. 30, 190–202 (2009).
    Article  Google Scholar 

    22.
    Leimgruber, K. L. et al. Give what you get: Capuchin monkeys (Cebus apella) and 4-year-old children pay forward positive and negative outcomes to conspecifics. PLoS One 9, e87035 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    23.
    Claidière, N. et al. Selective and contagious prosocial resource donation in capuchin monkeys, chimpanzees and humans. Sci. Rep. 5, 7631 (2015).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    24.
    Gfrerer, N. & Taborsky, M. Working dogs cooperate among one another by generalised reciprocity. Sci. Rep. 7, 43867 (2017).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    25.
    Rutte, C. & Taborsky, M. Generalized reciprocity in rats. PLoS Biol. 5, e196 (2007).
    CAS  Article  Google Scholar 

    26.
    Rutte, C. & Taborsky, M. The influence of social experience on cooperative behaviour of rats (Rattus norvegicus): Direct vs generalised reciprocity. Behav. Ecol. Sociobiol. 62, 499–505 (2008).
    Article  Google Scholar 

    27.
    Schneeberger, K., Dietz, M. & Taborsky, M. Reciprocal cooperation between unrelated rats depends on cost to donor and benefit to recipient. BMC Evol. Biol. 12, 41 (2012).
    PubMed  PubMed Central  Article  Google Scholar 

    28.
    Schweinfurth, M. K., Aeschbacher, J., Santi, M. & Taborsky, M. Male Norway rats cooperate according to direct but not generalized reciprocity rules. Anim. Behav. 152, 93–101 (2019).
    Article  Google Scholar 

    29.
    Dolivo, V. & Taborsky, M. Cooperation among Norway rats: The importance of visual cues for reciprocal cooperation, and the role of coercion. Ethology 121, 1071–1080 (2015).
    Article  Google Scholar 

    30.
    Dolivo, V. & Taborsky, M. Norway rats reciprocate help according to the quality of help they received. Biol. Lett. 11, 20140959 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    31.
    Wood, R. I., Kim, J. Y. & Li, G. R. Cooperation in rats playing the iterated Prisoner’s Dilemma game. Anim. Behav. 114, 27–35 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    32.
    Schweinfurth, M. K. & Taborsky, M. The transfer of alternative tasks in reciprocal cooperation. Anim. Behav. 131, 35–41 (2017).
    Article  Google Scholar 

    33.
    Schweinfurth, M. K. & Taborsky, M. Reciprocal trading of different commodities in Norway rats. Curr. Biol. 28, 594–599 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    34.
    Stieger, B., Schweinfurth, M. K. & Taborsky, M. Reciprocal allogrooming among unrelated Norway rats (Rattus norvegicus) is affected by previously received cooperative, affiliative and aggressive behaviours. Behav. Ecol. Sociobiol. 71, 182 (2017).
    Article  Google Scholar 

    35.
    Delmas, G. E., Lew, S. E. & Zanutto, S. B. High mutual cooperation rates in rats learning reciprocal altruism: The role of payoff matrix. PLoS ONE 14, 1–14 (2019).
    Article  CAS  Google Scholar 

    36.
    Barnett, S. A. & Spencer, M. M. Feeding social behaviour and interspecific competition in wild rats. Behaviour 3, 229–242 (1951).
    Google Scholar 

    37.
    Norton, S., Culver, B. & Mullenix, P. Development of nocturnal behavior in albino rats. Behav. Biol. 15, 317–331 (1975).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    38.
    Schweinfurth, M. K. & Taborsky, M. Rats play tit-for-tat instead of integrating cooperative experiences over multiple interactions. Proc. R. Soc. 287, 20192423 (2020).
    Google Scholar 

    39.
    Engqvist, L. The mistreatment of covariate interaction terms in linear model analyses of behavioural and evolutionary ecology studies. Anim. Behav. 70, 967–971 (2005).
    Article  Google Scholar 

    40.
    Kilkenny, C., Browne, W. J., Cuthill, I. C., Emerson, M. & Altman, D. G. Improving bioscience research reporting: The arrive guidelines for reporting animal research. PLoS Biol. 8, 6–10 (2010).
    Article  CAS  Google Scholar 

    41.
    Schweinfurth, M. K. et al. Do female Norway rats form social bonds? Behav. Ecol. Sociobiol. 71, 98 (2017).
    Article  Google Scholar 

    42.
    Davis, D. E. The characteristics of rat populations. Q. Rev. Biol. 28, 373–401 (1953).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    43.
    McGuire, B., Pizzuto, T., Bemis, W. E. & Getz, L. L. General ecology of a rural population of Norway rats (Rattus norvegicus) based on intensive live trapping. Am. Midl. Nat. 155, 221–236 (2006).
    Article  Google Scholar 

    44.
    Dunbar, R. I. Neocortex size as a constraint on group size in primates. J. Hum. Evol. 22, 469–493 (1992).
    Article  Google Scholar 

    45.
    Panoz-Brown, D. et al. Replay of episodic memories in the rat. Curr. Biol. 28, 1628–1634 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    46.
    Crystal, J. D. Prospective memory. Curr. Biol. 23, 750–751 (2013).
    Article  CAS  Google Scholar 

    47.
    Telle, H. Beitrag zur Erkenntnis der Verhaltensweise von Ratten, vergleichend dargestellt bei Rattus norvegicus und Rattus rattus. Zeitschrift für Angew. Zool. 53, 129–196 (1966).
    Google Scholar 

    48.
    Calhoun, J. B. The ecology and sociobiology of the Norway rat. (U.S. Dept. of Health, Education, and Welfare, Public Health Service, 1979).

    49.
    Mogil, J. S. Mice are people too: Increasing evidence for cognitive, emotional and social capabilities in laboratory rodents. Can. Psychol. 60, 14–20 (2019).
    Article  Google Scholar 

    50.
    Dolivo, V., Rutte, C. & Taborsky, M. Ultimate and proximate mechanisms of reciprocal altruism in rats. Learn. Behav. 44, 223–226 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    51.
    Müller, J. J. A., Massen, J. J. M., Bugnyar, T. & Osvath, M. Ravens remember the nature of a single reciprocal interaction sequence over 2 days and even after a month. Anim. Behav. 128, 69–78 (2017).
    Article  Google Scholar 

    52.
    Stevens, J. R. & Hauser, M. D. Why be nice? Psychological constraints on the evolution of cooperation. Trends Cogn. Sci. 8, 60–65 (2004).
    PubMed  Article  PubMed Central  Google Scholar  More

  • in

    Honey bee hives decrease wild bee abundance, species richness, and fruit count on farms regardless of wildflower strips

    1.
    Steffan-Dewenter, I., Potts, S. G. & Packer, L. Pollinator diversity and crop pollination services are at risk. Trends Ecol. Evol. 20, 651–652 (2005).
    PubMed  Article  PubMed Central  Google Scholar 
    2.
    Aizen, M. A., Garibaldi, L. A., Cunningham, S. A. & Klein, A. M. Long-term global trends in crop yield and production reveal no current pollination shortage but increasing pollinator dependency. Curr. Biol. 18, 1572–1575 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    3.
    Garibaldi, L. A., Aizen, M. A., Klein, A. M., Cunningham, S. A. & Harder, L. D. Global growth and stability of agricultural yield decrease with pollinator dependence. Proc. Natl. Acad. Sci. 108, 5909–5914 (2011).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    4.
    Goulson, D. Effects of introduced bees on native ecosystems. Annu. Rev. Ecol. Evol. Syst. 34, 1–26 (2003).
    Article  Google Scholar 

    5.
    Paini, D. Impact of the introduced honey bee (Apis mellifera) (Hymenoptera: Apidae) on native bees: A review. Aust. Ecol. 29, 399–407 (2004).
    Article  Google Scholar 

    6.
    Aslan, C. E., Liang, C. T., Galindo, B., Kimberly, H. & Topete, W. The role of honey bees as pollinators in natural areas. Nat. Areas J. 36, 478–489 (2016).
    Article  Google Scholar 

    7.
    Mallinger, R. E., Gaines-Day, H. R. & Gratton, C. Do managed bees have negative effects on wild bees? A systematic review of the literature. PLoS ONE 12, e0189268 (2017).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    8.
    Wignall, V. R. et al. Seasonal variation in exploitative competition between honeybees and bumblebees. Oecologia 192, 351–361 (2020).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Thomson, D. M. Detecting the effects of introduced species: A case study of competition between Apis and Bombus. Oikos 114, 407–418 (2006).
    Article  Google Scholar 

    10.
    Franco, E. L., Aguiar, C. M. & Ferreiraz, V. S. Plant use and niche overlap between the introduced honey bee (Apis mellifera) and the native bumblebee (Bombus atratus) (Hymenoptera: Apidae) in an area of tropical mountain vegetation in northeastern Brazil. Sociobiology 53, 141–150 (2009).
    Google Scholar 

    11.
    Herbertsson, L., Lindström, S. A., Rundlöf, M., Bommarco, R. & Smith, H. G. Competition between managed honeybees and wild bumblebees depends on landscape context. Basic Appl. Ecol. 17, 609–616 (2016).
    Article  Google Scholar 

    12.
    Thomson, D. M. Local bumble bee decline linked to recovery of honey bees, drought effects on floral resources. Ecol. Lett. 19, 1247–1255 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    13.
    Greenleaf, S. S. & Kremen, C. Wild bees enhance honey bees’ pollination of hybrid sunflower. Proc. Natl. Acad. Sci. 103, 13890–13895 (2006).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    14.
    Badano, E. I. & Vergara, C. H. Potential negative effects of exotic honey bees on the diversity of native pollinators and yield of highland coffee plantations. Agric. For. Entomol. 13, 365–372 (2011).
    Article  Google Scholar 

    15.
    Brittain, C., Williams, N., Kremen, C. & Klein, A.-M. Synergistic effects of non-Apis bees and honey bees for pollination services. Proc. R. Soc. B Biol. Sci. 280, 20122767 (2013).
    Article  Google Scholar 

    16.
    Müller, H. T. Interaction Between Bombus terrestris and Honeybees in Red Clover Fields Reduces Abundance of Other Bumblebees and Red Clover Yield and Honeybees in Red Clover Fields Reduces Abundance of Other Bumblebees and Red Clover Yield (Norwegian University of Life Sciences, Ås, 2016).
    Google Scholar 

    17.
    Grass, I. et al. Pollination limitation despite managed honeybees in South African macadamia orchards. Agric. Ecosyst. Environ. 260, 11–18 (2018).
    Article  Google Scholar 

    18.
    hUallacháin, D. Ó. (United Nations Convention to Combat Desertification, Bonn, Germany, 2017).

    19.
    Vaughan, M. & Skinner, M. Using 2014 farm bill programs for pollinator conservation. USDA Biol. Tech. Note 78, 2nd Ed. (2015).

    20.
    Vaughan, M. & Skinner, M. Using Farm Bill programs for pollinator conservation. USDA-NRCS National Plant Data Center, USDA Biol. Tech. Note 78 (2008).

    21.
    FSA. CP42 pollinator habitat: Establishing and supporting diverse pollinator-friendly habitat. (Farm Service Agency, U.S. Department of Agriculture, Washington, D.C., 2013).

    22.
    Venturini, E. M., Drummond, F. A., Hoshide, A. K., Dibble, A. C. & Stack, L. B. Pollination reservoirs for wild bee habitat enhancement in cropping systems: a review. Agroecol. Sustain. Food Syst. 41, 101–142 (2017).
    Article  Google Scholar 

    23.
    Wood, T. J., Holland, J. M., Hughes, W. O. & Goulson, D. Targeted agri-environment schemes significantly improve the population size of common farmland bumblebee species. Mol. Ecol. 24, 1668–1680 (2015).
    PubMed  Article  PubMed Central  Google Scholar 

    24.
    Haaland, C. & Gyllin, M. Butterflies and bumblebees in greenways and sown wildflower strips in southern Sweden. J. Insect Conserv. 14, 125–132 (2010).
    Article  Google Scholar 

    25.
    Ponisio, L. C., M’Gonigle, L. K. & Kremen, C. On-farm habitat restoration counters biotic homogenization in intensively managed agriculture. Glob. Change Biol. 22, 704–715 (2016).
    ADS  Article  Google Scholar 

    26.
    Dolezal, A. G., Clair, A. L. S., Zhang, G., Toth, A. L. & O’Neal, M. E. Native habitat mitigates feast–famine conditions faced by honey bees in an agricultural landscape. Proc. Natl. Acad. Sci. 116, 25147–25155 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    27.
    Venturini, E., Drummond, F., Hoshide, A., Dibble, A. & Stack, L. B. Pollination reservoirs in lowbush blueberry (Ericales: Ericaceae). J. Econ. Entomol. 110, 333–346 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    28.
    Morandin, L. A. & Kremen, C. Hedgerow restoration promotes pollinator populations and exports native bees to adjacent fields. Ecol. Appl. 23, 829–839 (2013).
    PubMed  Article  PubMed Central  Google Scholar 

    29.
    Blaauw, B. R. & Isaacs, R. Flower plantings increase wild bee abundance and the pollination services provided to a pollination-dependent crop. J. Appl. Ecol. 51, 890–898 (2014).
    Article  Google Scholar 

    30.
    Feltham, H., Park, K., Minderman, J. & Goulson, D. Experimental evidence of the benefit of wild flower strips to crop pollination. Ecol. Evolut. 5, 3523–3530 (2015).
    Article  Google Scholar 

    31.
    Gross, C. & Mackay, D. Honeybees reduce fitness in the pioneer shrub Melastoma affine (Melastomataceae). Biol. Cons. 86, 169–178 (1998).
    Article  Google Scholar 

    32.
    do Carmo, R. M., Franceschinelli, E. V. & da Silveira, F. A. Introduced honeybees (Apis mellifera) reduce pollination success without affecting the floral resource taken by native pollinators. Biotropica 36, 371–376 (2004).
    Google Scholar 

    33.
    Bruckman, D. & Campbell, D. R. Floral neighborhood influences pollinator assemblages and effective pollination in a native plant. Oecologia 176, 465–476 (2014).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    34.
    Garibaldi, L. A. et al. Wild pollinators enhance fruit set of crops regardless of honey bee abundance. Science 339, 1608–1611 (2013).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    35.
    Carvalheiro, L. G. et al. Natural and within-farmland biodiversity enhances crop productivity. Ecol. Lett. 14, 251–259 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    36.
    Jönsson, A. M. et al. Modelling the potential impact of global warming on Ips typographus voltinism and reproductive diapause. Clim. Change 109, 695–718 (2011).
    ADS  Article  Google Scholar 

    37.
    Scheper, J. et al. Local and landscape-level floral resources explain effects of wildflower strips on wild bees across four European countries. J. Appl. Ecol. 52, 1165–1175 (2015).
    Article  Google Scholar 

    38.
    Krimmer, E., Martin, E. A., Krauss, J., Holzschuh, A. & Steffan-Dewenter, I. Size, age and surrounding semi-natural habitats modulate the effectiveness of flower-rich agri-environment schemes to promote pollinator visitation in crop fields. Agric. Ecosyst. Environ. 284, 106590 (2019).
    Article  Google Scholar 

    39.
    Klein, A. M. et al. Wild pollination services to California almond rely on semi-natural habitat. J. Appl. Ecol. 49, 723–732 (2012).
    Google Scholar 

    40.
    Grab, H., Poveda, K., Danforth, B. & Loeb, G. Landscape context shifts the balance of costs and benefits from wildflower borders on multiple ecosystem services. Proc. R. Soc. B Biol. Sci. 285, 20181102 (2018).
    Article  Google Scholar 

    41.
    Prendergast, K. S., Menz, M. H., Dixon, K. W. & Bateman, P. W. The relative performance of sampling methods for native bees: An empirical test and review of the literature. Ecosphere 11, e03076 (2020).
    Article  Google Scholar 

    42.
    Cane, J. H., Minckley, R. L. & Kervin, L. J. Sampling bees (Hymenoptera: Apiformes) for pollinator community studies: pitfalls of pan-trapping. J. Kansas Entomol. Soc. 73, 225–231 (2000).
    Google Scholar 

    43.
    O’Connor, R. S. et al. Monitoring insect pollinators and flower visitation: The effectiveness and feasibility of different survey methods. Methods Ecol. Evol. 10, 2129–2140. https://doi.org/10.1111/2041-210x.13292 (2019).
    Article  Google Scholar 

    44.
    Graystock, P., Blane, E. J., McFrederick, Q. S., Goulson, D. & Hughes, W. O. Do managed bees drive parasite spread and emergence in wild bees?. Int. J. Parasitol. Parasites Wildlife 5, 64–75 (2016).
    Article  Google Scholar 

    45.
    Alger, S. A., Burnham, P. A., Boncristiani, H. F. & Brody, A. K. RNA virus spillover from managed honeybees (Apis mellifera) to wild bumblebees (Bombus spp.). PloS One 14, e0217822 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    46.
    Schaffer, W. M. et al. Competition, foraging energetics, and the cost of sociality in three species of bees. Ecology 60, 976–987 (1979).
    Article  Google Scholar 

    47.
    Pleasants, J. M. Bumblebee response to variation in nectar availability. Ecology 62, 1648–1661 (1981).
    Article  Google Scholar 

    48.
    Ginsberg, H. S. Foraging ecology of bees in an old field. Ecology 64, 165–175 (1983).
    Article  Google Scholar 

    49.
    Schaffer, W. M. et al. Competition for nectar between introduced honey bees and native North American bees and ants. Ecology 64, 564–577 (1983).
    Article  Google Scholar 

    50.
    Gross, C. L. The effect of introduced honeybees on native bee visitation and fruit-set in Dillwynia juniperina (Fabaceae) in a fragmented ecosystem. Biol. Cons. 102, 89–95 (2001).
    Article  Google Scholar 

    51.
    Hudewenz, A. & Klein, A.-M. Competition between honey bees and wild bees and the role of nesting resources in a nature reserve. J. Insect Conserv. 17, 1275–1283 (2013).
    Article  Google Scholar 

    52.
    Johnson, L. K. & Hubbell, S. P. Aggression and competition among stingless bees: Field studies. Ecology 55, 120–127 (1974).
    Article  Google Scholar 

    53.
    Winfree, R., Fox, J. W., Williams, N. M., Reilly, J. R. & Cariveau, D. P. Abundance of common species, not species richness, drives delivery of a real-world ecosystem service. Ecol. Lett. 18, 626–635 (2015).
    PubMed  Article  PubMed Central  Google Scholar 

    54.
    Woodcock, B. A. et al. Meta-analysis reveals that pollinator functional diversity and abundance enhance crop pollination and yield. Nat. Commun. 10, 1–10 (2019).
    ADS  CAS  Article  Google Scholar 

    55.
    Garibaldi, L. A. et al. From research to action: enhancing crop yield through wild pollinators. Front. Ecol. Environ. 12, 439–447 (2014).
    Article  Google Scholar 

    56.
    Connelly, H., Poveda, K. & Loeb, G. Landscape simplification decreases wild bee pollination services to strawberry. Agric. Ecosyst. Environ. 211, 51–56 (2015).
    Article  Google Scholar 

    57.
    MacInnis, G. & Forrest, J. R. K. Pollination by wild bees yields larger strawberries than pollination by honey bees. J. Appl. Ecol. 56, 824–832. https://doi.org/10.1111/1365-2664.13344 (2019).
    Article  Google Scholar 

    58.
    Seeley, T. D. Social foraging by honeybees: how colonies allocate foragers among patches of flowers. Behav. Ecol. Sociobiol. 19, 343–354 (1986).
    Article  Google Scholar 

    59.
    Bänsch, S., Tscharntke, T., Gabriel, D. & Westphal, C. Crop pollination services: complementary resource use by social vs solitary bees facing crops with contrasting flower supply. J. Appl. Ecol. https://doi.org/10.1111/1365-2664.13777 (2020).

    60.
    Nye, W. P. & Anderson, J. L. Insect pollinators frequenting strawberry blossoms and the effect of honey bees on yield and fruit quality. J. Am. Soc. Horticult. Sci. 99, 40 (1974).
    Google Scholar 

    61.
    De Oliveira, D., Savoie, L. & Vincent, C. in VI International Symposium on Pollination 288, 420–424 (1990).

    62.
    Chagnon, M., Gingras, J. & DeOliveira, D. Complementary aspects of strawberry pollination by honey and indigenous bees (Hymenoptera). J. Econ. Entomol. 86, 416–420 (1993).
    Article  Google Scholar 

    63.
    Horth, L. & Campbell, L. A. Supplementing small farms with native mason bees increases strawberry size and growth rate. J. Appl. Ecol. 55, 591–599 (2018).
    Article  Google Scholar 

    64.
    Pfister, S. C. et al. Dominance of cropland reduces the pollen deposition from bumble bees. Sci. Rep. 8, 13873 (2018).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    65.
    Artz, D. R. & Nault, B. A. Performance of Apis mellifera, Bombus impatiens, and Peponapis pruinosa (Hymenoptera: Apidae) as pollinators of pumpkin. J. Econ. Entomol. 104, 1153–1161 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    66.
    Petersen, J., Huseth, A. & Nault, B. Evaluating pollination deficits in pumpkin production in New York. Environ. Entomol. 43, 1247–1253 (2014).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    67.
    McGrady, C., Troyer, R. & Fleischer, S. Wild bee visitation rates exceed pollination thresholds in commercial cucurbita agroecosystems. J. Econ. Entomol. 113, 562–574 (2020).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    68.
    Geslin, B. et al. Advances in Ecological Research Vol. 57, 147–199 (Elsevier, San Diego, 2017).
    Google Scholar 

    69.
    Steffan-Dewenter, I. & Tscharntke, T. Resource overlap and possible competition between honey bees and wild bees in central Europe. Oecologia 122, 288–296 (2000).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    70.
    Torné-Noguera, A., Rodrigo, A., Osorio, S. & Bosch, J. Collateral effects of beekeeping: Impacts on pollen-nectar resources and wild bee communities. Basic Appl. Ecol. 17, 199–209 (2016).
    Article  Google Scholar 

    71.
    Free, J. B. Insect Pollination of Crops (Academic Press, London, 1970).
    Google Scholar 

    72.
    Delaplane, K. S., Mayer, D. R. & Mayer, D. F. Crop pollination by bees. (CABI, 2000).

    73.
    Phillips, B. Current honey bee and bumble bee stocking information. Michigan State University, MSU Extension: Pollination (2019). https://www.canr.msu.edu/news/current_honey_bee_stocking_information_and_an_introduction_to_commercial_bu.

    74.
    Angelella, G. M. & O’Rourke, M. E. Pollinator habitat establishment after organic and no-till seedbed preparation methods. HortScience 52, 1349–1355 (2017).
    CAS  Article  Google Scholar 

    75.
    Blaauw, B. R. & Isaacs, R. Larger patches of diverse floral resources increase insect pollinator density, diversity, and their pollination of native wildflowers. Basic Appl. Ecol. 15, 701–711 (2014).
    Article  Google Scholar 

    76.
    Klatt, B. K. et al. Bee pollination improves crop quality, shelf life and commercial value. Proc. R. Soc. B Biol. Sci. 281, 20132440 (2014).
    Article  Google Scholar 

    77.
    King, S. R., Davis, A. R. & Wehner, T. C. Classical genetics and traditional breeding. In Genetics, Genomics, and Breeding of Cucurbits (eds. Wang, Y.-H. et al.) 61–92 (CRC Press, 2012).

    78.
    Kronenberg, H. G. Poor fruit setting in strawberries. I. Euphytica 8, 47–57 (1959).
    Article  Google Scholar 

    79.
    Kronenberg, H. G., Braak, J. & Zeilinga, A. Poor fruit setting in strawberries. II. Euphytica 8, 245–251 (1959).
    Article  Google Scholar 

    80.
    Robinson, R. W. & Decker-Walters, D. S. Cucurbits (CAB Intl., New York, 1997).
    Google Scholar 

    81.
    R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, Vienna, Austria, 2019).

    82.
    Magnusson, A. et al. Package ‘glmmTMB’. R Package Version 0.2. 0 (2017).

    83.
    Bates, D., Sarkar, D., Bates, M. D. & Matrix, L. The lme4 package. R Package Version 2, 74 (2007).
    Google Scholar 

    84.
    Lenth, R., Singmann, H., Love, J., Buerkner, P. & Herve, M. Emmeans: Estimated marginal means, aka least-squares means. R Package Version 1, 3 (2018).
    Google Scholar 

    85.
    Wien, H., Stapleton, S., Maynard, D., McClurg, C. & Riggs, D. Flowering, sex expression, and fruiting of pumpkin (Cucurbita sp.) cultivars under various temperatures in greenhouse and distant field trials. HortScience 39, 239–242 (2004).
    Article  Google Scholar  More