tech_banner
Applied and Environmental Microbiology Journal Homepage
Can\'t sign in? Forgot your username? Enter your email address below and we will send you your username If the address matches an existing account you will receive an email with instructions to retrieve your usernameCan\'t sign in? Forgot your password?Enter your email address below and we will send you the reset instructions If the address matches an existing account you will receive an email with instructions to reset your password. Close Dr. Gemma Reguera Editor in Chief (2026) | Michigan State University Gemma Reguera is a professor in the Department of Microbiology and Molecular Genetics at Michigan State University. Her work investigates energy conversion reactions catalyzed by microbes in natural and anthropogenic systems. Editorial Board Applied and Environmental MicrobiologyArticleJuly 2016Symbiotic Nitrogen Fixation and the Challenges to Its Extension to Nonlegumes Florence Mus, Matthew B. Crook, Kevin Garcia, Amaya Garcia Costas, Barney A. Geddes, Evangelia D. Kouri, Ponraj Paramasivan, Min-Hyung Ryu, Giles E. D. Oldroyd, Philip S. Poole, Michael K. Udvardi, Christopher A. Voigt, Jean-Michel Anéand John W. PetersSymbiotic Nitrogen Fixation and the Challenges to Its Extension to NonlegumesAuthors: Florence Mus, Matthew B. Crook, Kevin Garcia, Amaya Garcia Costas, Barney A. Geddes, Evangelia D. Kouri, Ponraj Paramasivan, … Show All … , Min-Hyung Ryu, Giles E. D. Oldroyd, Philip S. Poole, Michael K. Udvardi, Christopher A. Voigt, Jean-Michel Ané, and John W. Peters Show FewerDOI: https://doi.org/10.1128/AEM.01055-16Volume 82, Number 131 July 2016ABSTRACTREFERENCESABSTRACTAccess to fixed or available forms of nitrogen limits the productivity of crop plants and thus food production. Nitrogenous fertilizer production currently represents a significant expense for the efficient growth of various crops in the developed world. There are significant potential gains to be had from reducing dependence on nitrogenous fertilizers in agriculture in the developed world and in developing countries, and there is significant interest in research on biological nitrogen fixation and prospects for increasing its importance in an agricultural setting. Biological nitrogen fixation is the conversion of atmospheric N2 to NH3, a form that can be used by plants. However, the process is restricted to bacteria and archaea and does not occur in eukaryotes. Symbiotic nitrogen fixation is part of a mutualistic relationship in which plants provide a niche and fixed carbon to bacteria in exchange for fixed nitrogen. This process is restricted mainly to legumes in agricultural systems, and there is considerable interest in exploring whether similar symbioses can be developed in nonlegumes, which produce the bulk of human food. We are at a juncture at which the fundamental understanding of biological nitrogen fixation has matured to a level that we can think about engineering symbiotic relationships using synthetic biology approaches. This minireview highlights the fundamental advances in our understanding of biological nitrogen fixation in the context of a blueprint for expanding symbiotic nitrogen fixation to a greater diversity of crop plants through synthetic biology.REFERENCES1.Boyd ES, Peters JW. 2013. New insights into the evolutionary history of biological nitrogen fixation. Front Microbiol 4:201.CrossrefPubMedGoogle Scholar2.Hardoim PR, van Overbeek LS, Berg G, Pirttilä AM, Compant S, Campisano A, Döring M, Sessitsch A. 2015. The hidden world within plants: ecological and evolutionary considerations for defining functioning of microbial endophytes. Microbiol Mol Biol Rev 79:293–320.CrossrefPubMedGoogle Scholar3.Santi C, Bogusz D, Franche C. 2013. Biological nitrogen fixation in non-legume plants. Ann Bot 111:743–767.CrossrefPubMedGoogle Scholar4.Compant S, Clément C, Sessitsch A. 2010. Plant growth-promoting bacteria in the rhizo- and endosphere of plants: their role, colonization, mechanisms involved and prospects for utilization. Soil Biol Biochem 42:669–678.CrossrefGoogle Scholar5.Schmid M, Hartmann A. 2007. Molecular phylogeny and ecology of root associated diazotrophic α- and β-proteobacteria, p 21–40. In Elmerich C, Newton W (ed), Associative and endophytic nitrogen-fixing bacteria and cyanobacterial associations. Springer-Verlag New York, Inc., New York, NY.Google Scholar6.Ahemad M, Kibret M. 2014. Mechanisms and applications of plant growth promoting rhizobacteria: current perspective. J King Saud Univ Sci 26:1–20.CrossrefGoogle Scholar7.Steenhoudt O, Vanderleyden J. 2000. Azospirillum, a free-living nitrogen-fixing bacterium closely associated with grasses: genetic, biochemical and ecological aspects. FEMS Microbiol Rev 24:487–506.CrossrefPubMedGoogle Scholar8.Peters GA, Meeks JC. 1989. The Azolla-Anabaena symbiosis: basic biology. Annu Rev Plant Physiol Plant Mol Biol 40:193–210.CrossrefGoogle Scholar9.Pedraza RO. 2008. Recent advances in nitrogen-fixing acetic acid bacteria. Int J Food Microbiol 125:25–35.CrossrefPubMedGoogle Scholar10.Nair DN, Padmavathy S. 2014. Impact of endophytic microorganisms on plants, environment and humans. Sci World J 2014:250693.CrossrefGoogle Scholar11.Eskin N, Vessey K, Tian L. 2014. Research progress and perspectives of nitrogen fixing bacterium, Gluconacetobacter diazotrophicus, in monocot plants. Int J Agron 2014:1–13.CrossrefGoogle Scholar12.Adams DG, Duggan PS. 2008. Cyanobacteria-bryophyte symbioses. J Exp Bot 59:1047–1058.CrossrefPubMedGoogle Scholar13.Costa JL, Lindblad P. 2002. Cyanobacteria in symbiosis with cycads, p 195–205. In Rai AN, Bergman B, Rasmussen U (ed), Cyanobacteria in symbiosis. Springer, Dordrecht, The Netherlands.Google Scholar14.Bergman B, Osborne B. 2002. The Gunnera:Nostoc symbiosis. Biol Environ 102B:35–39.Google Scholar15.Long SR. 1996. Rhizobium symbiosis: Nod factors in perspective. Plant Cell 8:1885–1898.CrossrefPubMedGoogle Scholar16.Oldroyd GED, Downie JA. 2008. Coordinating nodule morphogenesis with rhizobial infection in legumes. Annu Rev Plant Biol 59:519–546.CrossrefPubMedGoogle Scholar17.Davis EO, Evans IJ, Johnston AW. 1988. Identification of nodX, a gene that allows Rhizobium leguminosarum biovar viciae strain TOM to nodulate Afghanistan peas. Mol Gen Genet 212:531–535.CrossrefPubMedGoogle Scholar18.Devine TE, Kuykendall LD, Breithaupt BH. 1980. Nodulation of soybeans carrying the nodulation-restrictive gene, rj1, by an incompatible Rhizobium japonicum strain upon mixed inoculation with a compatible strain. Can J Microbiol 26:179–182.CrossrefPubMedGoogle Scholar19.Radutoiu S, Madsen LH, Madsen EB, Jurkiewicz A, Fukai E, Quistgaard EM, Albrektsen AS, James EK, Thirup S, Stougaard J. 2007. LysM domains mediate lipochitin-oligosaccharide recognition and Nfr genes extend the symbiotic host range. EMBO J 26:3923–3935.CrossrefPubMedGoogle Scholar20.Sprent JI, James EK. 2007. Legume evolution: where do nodules and mycorrhizas fit in? Plant Physiol 144:575–581.CrossrefPubMedGoogle Scholar21.Hirsch AM. 1992. Developmental biology of legume nodulation. New Phytol 122:211–237.CrossrefPubMedGoogle Scholar22.Svistoonoff S, Hocher V, Gherbi H. 2014. Actinorhizal root nodule symbioses: what is signalling telling on the origins of nodulation? Curr Opin Plant Biol 20:11–18.CrossrefPubMedGoogle Scholar23.Sytsma KJ, Morawetz J, Pires JC, Nepokroeff M, Conti E, Zjhra M, Hall JC, Chase MW. 2002. Urticalean rosids: circumscription, rosid ancestry, and phylogenetics based on rbcL, trnL-F, and ndhF sequences. Am J Bot 89:1531–1546.CrossrefPubMedGoogle Scholar24.Behm JE, Geurts R, Kiers ET. Parasponia: a novel system for studying mutualism stability. Trends Plant Sci 19:757–763.CrossrefPubMedGoogle Scholar25.Cárdenas L, Domínguez J, Quinto C, López-Lara IM, Lugtenberg BJ, Spaink HP, Rademaker GJ, Haverkamp J, Thomas-Oates JE. 1995. Isolation, chemical structures and biological activity of the lipo-chitin oligosaccharide nodulation signals from Rhizobium etli. Plant Mol Biol 29:453–464.CrossrefPubMedGoogle Scholar26.Perret X, Staehelin C, Broughton WJ. 2000. Molecular basis of symbiotic promiscuity. Microbiol Mol Biol Rev 64:180–201.CrossrefPubMedGoogle Scholar27.Finan TM, Hirsch AM, Leigh JA, Johansen E, Kuldau GA, Deegan S, Walker GC, Signer ER. 1985. Symbiotic mutants of Rhizobium meliloti that uncouple plant from bacterial differentiation. Cell 40:869–877.CrossrefPubMedGoogle Scholar28.Leigh JA, Signer ER, Walker GC. 1985. Exopolysaccharide-deficient mutants of Rhizobium meliloti that form ineffective nodules. Proc Natl Acad Sci U S A 82:6231–6235.CrossrefPubMedGoogle Scholar29.Cheng HP, Walker GC. 1998. Succinoglycan is required for initiation and elongation of infection threads during nodulation of alfalfa by Rhizobium meliloti. J Bacteriol 180:5183–5191.CrossrefPubMedGoogle Scholar30.Dylan T, Ielpi L, Stanfield S, Kashyap L, Douglas C, Yanofsky M, Nester E, Helinski DR, Ditta G. 1986. Rhizobium meliloti genes required for nodule development are related to chromosomal virulence genes in Agrobacterium tumefaciens. Proc Natl Acad Sci U S A 83:4403–4407.CrossrefPubMedGoogle Scholar31.Mithöfer A. 2002. Suppression of plant defence in rhizobia-legume symbiosis. Trends Plant Sci 7:440–444.CrossrefPubMedGoogle Scholar32.Kawaharada Y, Kelly S, Nielsen MW, Hjuler CT, Gysel K, Muszynski A, Carlson RW, Thygesen MB, Sandal N, Asmussen MH, Vinther M, Andersen SU, Krusell L, Thirup S, Jensen KJ, Ronson CW, Blaise M, Radutoiu S, Stougaard J. 2015. Receptor-mediated exopolysaccharide perception controls bacterial infection. Nature 523:308–312.CrossrefPubMedGoogle Scholar33.Persson T, Battenberg K, Demina IV, Vigil-Stenman T, Vanden Heuvel B, Pujic P, Facciotti MT, Wilbanks EG, O\'Brien A, Fournier P, Cruz Hernandez MA, Mendoza Herrera A, Médigue C, Normand P, Pawlowski K, Berry AM. Candidatus Frankia datiscae Dg1, the actinobacterial microsymbiont of Datisca glomerata, expresses the canonical nod genes nodABC in symbiosis with its host plant. PLoS One 10:e0127630.CrossrefPubMedGoogle Scholar34.Cérémonie H, Debellé F, Fernandez MP. 1999. Structural and functional comparison of Frankia root hair deforming factor and rhizobia Nod factor. Can J Bot 77:1293–1301.CrossrefGoogle Scholar35.Wagner GM. 1997. Azolla: a review of its biology and utilization. Bot Rev 63:1–26.CrossrefGoogle Scholar36.Jones DL, Nguyen C, Finlay RD. 2009. Carbon flow in the rhizosphere: carbon trading at the soil-root interface. Plant Soil 321:5–33.CrossrefGoogle Scholar37.Günter N, Volker R. 2007. The release of root exudates as affected by the plant physiological status, p 23–72. In Pinton R, Varanini Z, Nannipieri P (ed), The rhizophere: biochemistry and organic substances at the soil-plant interface. CRC Press, Boca Raton, FL.Google Scholar38.Turner TR, James EK, Poole PS. 2013. The plant microbiome. Genome Biol 14:209.CrossrefPubMedGoogle Scholar39.Turner TR, Ramakrishnan K, Walshaw J, Heavens D, Alston M, Swarbreck D, Osbourn A, Grant A, Poole PS. 2013. Comparative metatranscriptomics reveals kingdom level changes in the rhizosphere microbiome of plants. ISME J 7:2248–2258.CrossrefPubMedGoogle Scholar40.Kamilova F, Validov S, Azarova T, Mulders I, Lugtenberg B. 2005. Enrichment for enhanced competitive plant root tip colonizers selects for a new class of biocontrol bacteria. Environ Microbiol 7:1809–1817.CrossrefPubMedGoogle Scholar41.Kamilova F, Kravchenko LV, Shaposhnikov AI, Azarova T, Makarova N, Lugtenberg B. 2006. Organic acids, sugars, and l-tryptophane in exudates of vegetables growing on stonewool and their effects on activities of rhizosphere bacteria. Mol Plant Microbe Interact 19:250–256.CrossrefPubMedGoogle Scholar42.van Egeraat AWSM. 1975. The possible role of homoserine in the development of Rhizobium leguminosarum in the rhizosphere of pea seedlings. Plant Soil 42:381–386.CrossrefGoogle Scholar43.Vanderlinde EM, Hynes MF, Yost CK. 2014. Homoserine catabolism by Rhizobium leguminosarum bv. viciae 3841 requires a plasmid-borne gene cluster that also affects competitiveness for nodulation. Environ Microbiol 16:205–217.CrossrefPubMedGoogle Scholar44.Ramachandran VK, East AK, Karunakaran R, Downie JA, Poole PS. 2011. Adaptation of Rhizobium leguminosarum to pea, alfalfa and sugar beet rhizospheres investigated by comparative transcriptomics. Genome Biol 12:R106.CrossrefPubMedGoogle Scholar45.Baetz U, Martinoia E. 2014. Root exudates: the hidden part of plant defense. Trends Plant Sci 19:90–98.CrossrefPubMedGoogle Scholar46.Neal AL, Ahmad S, Gordon-Weeks R, Ton J. 2012. Benzoxazinoids in root exudates of maize attract Pseudomonas putida to the rhizosphere. PLoS One 7:e35498.CrossrefPubMedGoogle Scholar47.Fan J, Crooks C, Creissen G, Hill L, Fairhurst S, Doerner P, Lamb C. 2011. Pseudomonas sax genes overcome aliphatic isothiocyanate-mediated non-host resistance in Arabidopsis. Science 331:1185–1188.CrossrefPubMedGoogle Scholar48.Soedarjo M, Borthakur D. 1997. Mimosine produced by the tree-legume Leucaena provides growth advantages to some Rhizobium strains that utilize it as a source of carbon and nitrogen, p 87–92. In Elkan GH, Upchurch RG (ed), Current issues in symbiotic nitrogen fixation. Springer, Dordrecht, The Netherlands.Google Scholar49.Cai T, Cai W, Zhang J, Zheng H, Tsou AM, Xiao L, Zhong Z, Zhu J. 2009. Host legume-exuded antimetabolites optimize the symbiotic rhizosphere. Mol Microbiol 73:507–517.CrossrefPubMedGoogle Scholar50.Savka MA, Dessaux Y, Oger P, Rossbach S. 2002. Engineering bacterial competitiveness and persistence in the phytosphere. Mol Plant Microbe Interact 15:866–874.CrossrefPubMedGoogle Scholar51.Murphy PJ, Wexler W, Grzemski W, Rao JP, Gordon D. 1995. Rhizopines—their role in symbiosis and competition. Soil Biol Biochem 27:525–529.CrossrefGoogle Scholar52.Gordon DM, Ryder MH, Heinrich K, Murphy PJ. 1996. An experimental test of the rhizopine concept in Rhizobium meliloti. Appl Environ Microbiol 62:3991–3996.CrossrefPubMedGoogle Scholar53.Oger P, Petit A, Dessaux Y. 1997. Genetically engineered plants producing opines alter their biological environment. Nat Biotechnol 15:369–372.CrossrefPubMedGoogle Scholar54.Mondy S, Lenglet A, Beury-Cirou A, Libanga C, Ratet P, Faure D, Dessaux Y. 2014. An increasing opine carbon bias in artificial exudation systems and genetically modified plant rhizospheres leads to an increasing reshaping of bacterial populations. Mol Ecol 23:4846–4861.CrossrefPubMedGoogle Scholar55.Oger P, Mansouri H, Dessaux Y. 2000. Effect of crop rotation and soil cover on alteration of the soil microflora generated by the culture of transgenic plants producing opines. Mol Ecol 9:881–890.CrossrefPubMedGoogle Scholar56.Savka MA, Farrand SK. 1997. Modification of rhizobacterial populations by engineering bacterium utilization of a novel plant-produced resource. Nat Biotechnol 15:363–368.CrossrefPubMedGoogle Scholar57.Kiers ET, Rousseau RA, West SA, Denison RF. 2003. Host sanctions and the legume-Rhizobium mutualism. Nature 425:78–81.CrossrefPubMedGoogle Scholar58.Kiers ET, Duhamel M, Beesetty Y, Mensah JA, Franken O, Verbruggen E, Fellbaum CR, Kowalchuk GA, Hart MM, Bago A, Palmer TM, West SA, Vandenkoornhuyse P, Jansa J, Bücking H. 2011. Reciprocal rewards stabilize cooperation in the mycorrhizal symbiosis. Science 333:880–882.CrossrefPubMedGoogle Scholar59.Maróti G, Downie JA, Kondorosi É. 2015. Plant cysteine-rich peptides that inhibit pathogen growth and control rhizobial differentiation in legume nodules. Curr Opin Plant Biol 26:57–63.CrossrefPubMedGoogle Scholar60.Czernic P, Gully D, Cartieaux F, Moulin L, Guefrachi I, Patrel D, Pierre O, Fardoux J, Chaintreuil C, Nguyen P, Gressent F, Da Silva C, Poulain J, Wincker P, Rofidal V, Hem S, Barrière Q, Arrighi J-F, Mergaert P, Giraud E. 2015. Convergent evolution of endosymbiont differentiation in Dalbergioid and inverted repeat-lacking clade legumes mediated by nodule-specific cysteine-rich peptides. Plant Physiol 169:1254–1265.CrossrefPubMedGoogle Scholar61.Carro L, Pujic P, Trujillo M, Normand P. 2013. Micromonospora is a normal occupant of actinorhizal nodules. J Biosci 38:685–693.CrossrefPubMedGoogle Scholar62.Meeks JC, Elhai J. 2002. Regulation of cellular differentiation in filamentous cyanobacteria in free-living and plant-associated symbiotic growth states. Microbiol Mol Biol Rev 66:94–121.CrossrefPubMedGoogle Scholar63.Rai AN, Bergman B, Rasmussen U (ed). 2002. Cyanobacteria in symbiosis. Springer, Dordrecht, The Netherlands.CrossrefGoogle Scholar64.Valverde C, Huss-Danell K. 2008. Carbon and nitrogen metabolism in actinorhizal nodules, p 167–198. In Pawlowski K, Newton WE (ed), Nitrogen-fixing actinorhizal symbioses. Springer, Dordrecht, The Netherlands.Google Scholar65.Colebatch G, Desbrosses G, Ott T, Krusell L, Montanari O, Kloska S, Kopka J, Udvardi MK. 2004. Global changes in transcription orchestrate metabolic differentiation during symbiotic nitrogen fixation in Lotus japonicus. Plant J 39:487–512.CrossrefPubMedGoogle Scholar66.Benedito VA, Li H, Dai X, Wandrey M, He J, Kaundal R, Torres-Jerez I, Gomez SK, Harrison MJ, Tang Y, Zhao PX, Udvardi MK. 2010. Genomic inventory and transcriptional analysis of Medicago truncatula transporters. Plant Physiol 152:1716–1730.CrossrefPubMedGoogle Scholar67.Kouchi H, Yoneyama T. 1984. Dynamics of carbon photosynthetically assimilated in nodulated soya bean plants under steady-state conditions 2. The incorporation of 13C into carbohydrates, organic acids, amino acids and some storage compounds. Ann Bot 53:883–896.CrossrefGoogle Scholar68.Craig J, Barratt P, Tatge H, Déjardin A, Handley L, Gardner CD, Barber L, Wang T, Hedley C, Martin C, Smith AM. 1999. Mutations at the rug4 locus alter the carbon and nitrogen metabolism of pea plants through an effect on sucrose synthase. Plant J 17:353–362.CrossrefGoogle Scholar69.Horst I, Welham T, Kelly S, Kaneko T, Sato S, Tabata S, Parniske M, Wang TL. 2007. TILLING mutants of Lotus japonicus reveal that nitrogen assimilation and fixation can occur in the absence of nodule-enhanced sucrose synthase. Plant Physiol 144:806–820.CrossrefPubMedGoogle Scholar70.Udvardi M, Poole PS. 2013. Transport and metabolism in legume-rhizobia symbioses. Annu Rev Plant Biol 64:781–805.CrossrefPubMedGoogle Scholar71.Limpens E, Moling S, Hooiveld G, Pereira PA, Bisseling T, Becker JD, Küster H. 2013. Cell- and tissue-specific transcriptome analyses of Medicago truncatula root nodules. PLoS One 8:e64377.CrossrefPubMedGoogle Scholar72.Udvardi MK, Price GD, Gresshoff PM, Day DA. 1988. A dicarboxylate transporter on the peribacteroid membrane of soybean nodules. FEBS Lett 231:36–40.CrossrefGoogle Scholar73.Yurgel SN, Kahn ML. 2004. Dicarboxylate transport by rhizobia. FEMS Microbiol Rev 28:489–501.CrossrefPubMedGoogle Scholar74.Finan TM, Oresnik I, Bottacin A. 1988. Mutants of Rhizobium meliloti defective in succinate metabolism. J Bacteriol 170:3396–3403.CrossrefPubMedGoogle Scholar75.Finan TM, McWhinne E, Driscoll B, Watson RJ. 1991. Complex symbiotic phenotypes result from gluconeogenic mutations in Rhizobium meliloti. Mol Plant Microbe Interact 4:386–392.CrossrefGoogle Scholar76.Lodwig EM, Hosie AH, Bourdès A, Findlay K, Allaway D, Karunakaran R, Downie J, Poole PS. 2003. Amino-acid cycling drives nitrogen fixation in the legume-Rhizobium symbiosis. Nature 422:722–726.CrossrefPubMedGoogle Scholar77.Mulley G, White JP, Karunakaran R, Prell J, Bourdes A, Bunnewell S, Hill L, Poole PS. 2011. Mutation of GOGAT prevents pea bacteroid formation and N2 fixation by globally downregulating transport of organic nitrogen sources. Mol Microbiol 80:149–167.CrossrefPubMedGoogle Scholar78.Udvardi MK, Lister DL, Day DA. 1992. Isolation and characterization of a ntrC mutant of Bradyrhizobium (Parasponia) sp. ANU289. Microbiology 138:1019–1025.Google Scholar79.Patriarca EJ, Tatè R, Iaccarino M. 2002. Key role of bacterial NH4+ metabolism in Rhizobium-plant symbiosis. Microbiol Mol Biol Rev 66:203–222.CrossrefPubMedGoogle Scholar80.Day DA, Kaiser BN, Thomson R, Udvardi MK, Moreau S, Puppo A. 2001. Nutrient transport across symbiotic membranes from legume nodules. Aust J Plant Physiol 28:669–676.CrossrefGoogle Scholar81.Niemietz CM, Tyerman SD. 2000. Channel-mediated permeation of ammonia gas through the peribacteroid membrane of soybean nodules. FEBS Lett 465:110–114.CrossrefPubMedGoogle Scholar82.Tyerman SD, Whitehead LF, Day DA. 1995. A channel-like transporter for NH4+ on the symbiotic interface of N2-fixing plants. Nature 378:629–632.CrossrefGoogle Scholar83.Hwang JH, Ellingson SR, Roberts DM. 2010. Ammonia permeability of the soybean nodulin 26 channel. FEBS Lett 584:4339–4343.CrossrefPubMedGoogle Scholar84.Masalkar P, Wallace IS, Hwang JH, Roberts DM. 2010. Interaction of cytosolic glutamine synthetase of soybean root nodules with the C-terminal domain of the symbiosome membrane nodulin 26 aquaglyceroporin. J Biol Chem 285:23880–23888.CrossrefPubMedGoogle Scholar85.Eckardt NA. 2005. Insights into plant cellular mechanisms: of phosphate transporters and arbuscular mycorrhizal infection. Plant Cell 17:3213–3216.CrossrefGoogle Scholar86.Kaiser BN, Moreau S, Castelli J, Thomson R, Lambert A, Bogliolo S, Puppo A, Day DA. 2003. The soybean NRAMP homologue, GmDMT1, is a symbiotic divalent metal transporter capable of ferrous iron transport. Plant J 35:295–304.CrossrefPubMedGoogle Scholar87.Krusell L, Krause K, Ott T, Desbrosses G, Krämer U, Sato S, Nakamura Y, Tabata S, James EK, Sandal N, Stougaard J, Kawaguchi M, Miyamoto A, Suganuma N, Udvardi MK. 2005. The sulfate transporters SST1 is crucial for symbiotic nitrogen fixation in Lotus japonicus root nodules. Plant Cell 17:1625–1636.CrossrefPubMedGoogle Scholar88.Bellenger JP, Wichard T, Kustka AB, Kraepiel AML. 2008. Uptake of molybdenum and vanadium by a nitrogen-fixing soil bacterium using siderophores. Nat Geosci 1:243–246.CrossrefGoogle Scholar89.Delgado MJ, Tresierra-Ayala A, Talbi C, Bedmar EJ. 2006. Functional characterization of the Bradyrhizobium japonicum modA and modB genes involved in molybdenum transport. Microbiology 152:199–207.CrossrefPubMedGoogle Scholar90.Campbell GRO, Taga ME, Mistry K, Lloret J, Anderson PJ, Roth JR, Walker GC. 2006. Sinorhizobium meliloti bluB is necessary for production of 5,6-dimethylbenzimidazole, the lower ligand of B12. Proc Natl Acad Sci U S A 103:4634–4639.CrossrefPubMedGoogle Scholar91.Black KG, Parsons R, Osborne BA. 2002. Uptake and metabolism of glucose in the Nostoc-Gunnera symbiosis. New Phytol 153:297–305.CrossrefGoogle Scholar92.Peters JW, Boyd ES, Hamilton TL, Rubio L. 2011. Biochemistry of Mo-nitrogenase, p 59–100. In Moir JWB (ed), Nitrogen cycling in bacteria: molecular analysis. Caister Academic Press, Norfolk, United Kingdom.Google Scholar93.Rubio LM, Ludden PW. 2008. Biosynthesis of the iron-molybdenum cofactor of nitrogenase. Annu Rev Microbiol 62:93–111.CrossrefPubMedGoogle Scholar94.Edgren T, Nordlund S. 2004. The fixABCX genes in Rhodospirillum rubrum encode a putative membrane complex participating in electron transfer to nitrogenase. J Bacteriol 186:2052–2060.CrossrefPubMedGoogle Scholar95.Boyd ES, Costas AM, Hamilton TL, Mus F, Peters JW. 2015. Evolution of molybdenum nitrogenase during the transition from anaerobic to aerobic metabolism. J Bacteriol 197:1690–1699.CrossrefPubMedGoogle Scholar96.Ott T, van Dongen JT, Gunther C, Krusell L, Desbrosses G, Vigeolas H, Bock V, Czechowski T, Geigenberger P, Udvardi MK. 2005. Symbiotic leghemoglobins are crucial for nitrogen fixation in legume root nodules but not for general plant growth and development. Curr Biol 15:531–535.CrossrefPubMedGoogle Scholar97.Dixon R, Kahn D. 2004. Genetic regulation of biological nitrogen fixation. Nat Rev Microbiol 2:621–631.CrossrefPubMedGoogle Scholar98.Pitcher RS, Watmough NJ. 2004. The bacterial cytochrome cbb3 oxidases. Biochim Biophys Acta 1655:388–399.CrossrefPubMedGoogle Scholar99.Preisig O, Anthamatten D, Hennecke H. 1993. Genes for a microaerobically induced oxidase complex in Bradyrhizobium japonicum are essential for a nitrogen-fixing endosymbiosis. Proc Natl Acad Sci U S A 90:3309–3313.CrossrefPubMedGoogle Scholar100.Fay P. 1992. Oxygen relations of nitrogen fixation in cyanobacteria. Microbiol Rev 56:340–373.CrossrefPubMedGoogle Scholar101.Murry MA, Horne AJ, Benemann JR. 1984. Physiological studies of oxygen protection mechanisms in the heterocysts of Anabaena cylindrica. Appl Environ Microbiol 47:449–454.CrossrefPubMedGoogle Scholar102.Stal LJ, Krumbien WE. 1985. Nitrogenase activity in the non-heterocystous cyanobacterium Oscillatoria sp. grown under alternating light-dark cycles. Arch Microbiol 143:67–71.CrossrefGoogle Scholar103.Poole RK, Hill S. 1997. Respiratory protection of nitrogenase activity in Azotobacter vinelandii—roles of the terminal oxidases. Biosci Rep 17:303–320.CrossrefPubMedGoogle Scholar104.Maier RJ, Moshiri F. 2000. Role of the Azotobacter vinelandii nitrogenase-protective shethna protein in preventing oxygen-mediated cell death. J Bacteriol 182:3854–3857.CrossrefPubMedGoogle Scholar105.Sabra W, Zeng A-P, Lünsdorf H, Deckwer W-D. 2000. Effect of oxygen on formation and structure of Azotobacter vinelandii alginate and its role protecting nitrogenase. Appl Environ Microbiol 66:4037–4044.CrossrefPubMedGoogle Scholar106.Berry AM, Harriott OT, Moreau RA, Osman SF, Benson DR, Jones AD. 1993. Hopanoid lipids compose the Frankia vesicle envelope, presumptive barrier of oxygen diffusion to nitrogenase. Proc Natl Acad Sci U S A 90:6091–6094.CrossrefPubMedGoogle Scholar107.Pawlowski K, Bisseling T. 1996. Rhizobial and actinorhizal symbioses: what are the shared features? Plant Cell 8:1899–1913.CrossrefPubMedGoogle Scholar108.Oldroyd GE. 2013. Speak, friend, and enter: signalling systems that promote beneficial symbiotic associations in plants. Nat Rev Microbiol 11:252–263.CrossrefPubMedGoogle Scholar109.Oldroyd GE, Murray JD, Poole PS, Downie JA. 2011. The rules of engagement in the legume-rhizobial symbiosis. Annu Rev Genet 45:119–144.CrossrefPubMedGoogle Scholar110.Curatti L, Rubio LM. 2014. Challenges to develop nitrogen-fixing cereals by direct nif-gene transfer. Plant Sci 225:130–137.CrossrefPubMedGoogle Scholar111.Hawkesford MJ. 2014. Reducing the reliance on nitrogen fertilizer for wheat production. J Cereal Sci 59:276–283.CrossrefPubMedGoogle Scholar112.Weber E, Engler C, Gruetzner R, Werner S, Marillonnet S. 2011. A modular cloning system for standardized assembly of multigene constructs. PLoS One 6:e16765.CrossrefPubMedGoogle Scholar113.Gibson DG. 2011. Enzymatic assembly of overlapping DNA fragments. Methods Enzymol 498:349–361.CrossrefPubMedGoogle Scholar114.Rogers C, Oldroyd GED. 2014. Synthetic biology approaches to engineering the nitrogen symbiosis in cereals. J Exp Bot 65:1939–1946.CrossrefPubMedGoogle Scholar115.Martinez-Argudo I, Little R, Shearer N, Johnson P, Dixon R. 2004. The NifL-NifA system: a multidomain transcriptional regulatory complex that integrates environmental signals. J Bacteriol 186:601–610.CrossrefPubMedGoogle Scholar116.Bode HB, Müller R. 2005. The impact of bacterial genomics on natural product research. Angew Chem Int Ed Engl 44:6828–6846.CrossrefPubMedGoogle Scholar117.Hertweck C. 2009. Hidden biosynthetic treasures brought to light. Nat Chem Biol 5:450–452.CrossrefPubMedGoogle Scholar118.Brakhage AA, Schroeckh V. 2011. Fungal secondary metabolites–strategies to activate silent gene clusters. Fungal Genet Biol 48:15–22.CrossrefPubMedGoogle Scholar119.Temme K, Zhao D, Voigt CA. 2012. Refactoring the nitrogen fixation gene cluster from Klebsiella oxytoca. Proc Natl Acad Sci U S A 109:7085–7090.CrossrefPubMedGoogle Scholar120.Fischbach M, Voigt CA. 2010. Prokaryotic gene clusters: a rich toolbox for synthetic biology. Biotechnol J 5:1277–1296.CrossrefPubMedGoogle Scholar121.Jaschke PR, Lieberman EK, Rodriguez J, Sierra A, Endy D. 2012. A fully decompressed synthetic bacteriophage øX174 genome assembled and archived in yeast. Virology 434:278–284.CrossrefPubMedGoogle Scholar122.Chan LY, Kosuri S, Endy D. 2005. Refactoring bacteriophage T7. Mol Syst Biol 1:2005.0018.CrossrefPubMedGoogle Scholar123.Smanski MJ, Bhatia S, Zhao D, Park Y, Woodruff LBA, Giannoukos G, Ciulla D, Busby M, Calderon J, Nicol R, Gordon DB, Densmore D, Voigt CA. 2014. Functional optimization of gene clusters by combinatorial design and assembly. Nat Biotechnol 32:1241–1249.CrossrefPubMedGoogle Scholar124.Wang X, Yang JG, Chen L, Wang JL, Cheng Q, Dixon R, Wang YP. 2013. Using synthetic biology to distinguish and overcome regulatory and functional barriers related to nitrogen fixation. PLoS One 8:e68677.CrossrefPubMedGoogle Scholar125.Kosuri S, Church GM. 2014. Large-scale de novo DNA synthesis: technologies and applications. Nat Methods 11:499–507.CrossrefPubMedGoogle Scholar126.Kodumal SJ, Patel KG, Reid R, Menzella HG, Welch M, Santi DV. 2004. Total synthesis of long DNA sequences: synthesis of a contiguous 32-kb polyketide synthase gene cluster. Proc Natl Acad Sci U S A 101:15573–15578.CrossrefPubMedGoogle Scholar127.Annaluru N, Muller H, Mitchell LA, Ramalingam S, Stracquadanio G, Richardson SM, Dymond JS, Kuang Z, Scheifele LZ, Cooper EM, Cai Y, Zeller K, Agmon N, Han JS, Hadjithomas M, Tullman J, Caravelli K, Cirelli K, Guo Z, London V, Yeluru A, Murugan S, Kandavelou K, Agier N, Fischer G, Yang K, Martin JA, Bilgel M, Bohutski P, Boulier KM, Capaldo BJ, Chang J, Charoen K, Choi WJ, Deng P, DiCarlo JE, Doong J, Dunn J, Feinberg JI, Fernandez C, Floria CE, Gladowski D, Hadidi P, Ishizuka I, Jabbari J, Lau CY, Lee PA, Li S, Lin D, Linder ME, et al. 2014. Total synthesis of a functional designer eukaryotic chromosome. Science 344:55–58.CrossrefPubMedGoogle Scholar128.Gibson DG, Glass JI, Lartigue C, Noskov VN, Chuang RY, Algire MA, Benders GA, Montague MG, Ma L, Moodie MM, Merryman C, Vashee S, Krishnakumar R, Assad-Garcia N, Andrews-Pfannkoch C, Denisova EA, Young L, Qi ZQ, Segall-Shapiro TH, Calvey CH, Parmar PP, Hutchison CA, III, Smith HO, Venter JC. 2010. Creation of a bacterial cell controlled by a chemically synthesized genome. Science 329:52–56.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 82 • Number 13 • 1 July 2016Pages: 3698 - 3710Editor: R. M. KellyNorth Carolina State UniversityHistoryPublished online: 15 April 2016Copyright© 2016 Mus et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsFlorence MusDepartment of Chemistry and Biochemistry, Montana State University, Bozeman, Montana, USAView all articles by this authorMatthew B. CrookDepartment of Bacteriology, University of Wisconsin—Madison, Madison, Wisconsin, USAView all articles by this authorKevin GarciaDepartment of Bacteriology, University of Wisconsin—Madison, Madison, Wisconsin, USAView all articles by this authorAmaya Garcia CostasDepartment of Chemistry and Biochemistry, Montana State University, Bozeman, Montana, USAView all articles by this authorBarney A. GeddesDepartment of Plant Sciences, University of Oxford, Oxford, United KingdomView all articles by this authorEvangelia D. KouriPlant Biology Division, The Samuel Roberts Noble Foundation, Ardmore, Oklahoma, USAView all articles by this authorPonraj ParamasivanJohn Innes Centre, Norwich Research Park, Norwich, United KingdomView all articles by this authorMin-Hyung RyuDepartment of Biological Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts, USAView all articles by this authorGiles E. D. OldroydJohn Innes Centre, Norwich Research Park, Norwich, United KingdomView all articles by this authorPhilip S. PooleDepartment of Plant Sciences, University of Oxford, Oxford, United KingdomView all articles by this authorMichael K. UdvardiPlant Biology Division, The Samuel Roberts Noble Foundation, Ardmore, Oklahoma, USAView all articles by this authorChristopher A. VoigtDepartment of Biological Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts, USAView all articles by this authorJean-Michel AnéDepartment of Bacteriology, University of Wisconsin—Madison, Madison, Wisconsin, USAView all articles by this authorJohn W. PetersDepartment of Chemistry and Biochemistry, Montana State University, Bozeman, Montana, USAView all articles by this authorEditorR. M. KellyEditorNorth Carolina State UniversityNotesAddress correspondence to Jean-Michel Ané, [email protected], or John W. Peters, [email protected].Metrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleFebruary 2013Why Orange Guaymas Basin Beggiatoa spp. Are Orange: Single-Filament-Genome-Enabled Identification of an Abundant Octaheme Cytochrome with Hydroxylamine Oxidase, Hydrazine Oxidase, and Nitrite Reductase Activities Barbara J. MacGregor, Jennifer F. Biddle, Jason R. Siebert, Eric Staunton, Eric L. Hegg, Ann G. Matthysseand Andreas TeskeWhy Orange Guaymas Basin Beggiatoa spp. Are Orange: Single-Filament-Genome-Enabled Identification of an Abundant Octaheme Cytochrome with Hydroxylamine Oxidase, Hydrazine Oxidase, and Nitrite Reductase ActivitiesAuthors: Barbara J. MacGregor, Jennifer F. Biddle, Jason R. Siebert, Eric Staunton, Eric L. Hegg, Ann G. Matthysse, and Andreas TeskeDOI: https://doi.org/10.1128/AEM.02538-12Volume 79, Number 415 February 2013ABSTRACTREFERENCESABSTRACTOrange, white, and yellow vacuolated Beggiatoaceae filaments are visually dominant members of microbial mats found near sea floor hydrothermal vents and cold seeps, with orange filaments typically concentrated toward the mat centers. No marine vacuolate Beggiatoaceae are yet in pure culture, but evidence to date suggests they are nitrate-reducing, sulfide-oxidizing bacteria. The nearly complete genome sequence of a single orange Beggiatoa (\"Candidatus Maribeggiatoa”) filament from a microbial mat sample collected in 2008 at a hydrothermal site in Guaymas Basin (Gulf of California, Mexico) was recently obtained. From this sequence, the gene encoding an abundant soluble orange-pigmented protein in Guaymas Basin mat samples (collected in 2009) was identified by microcapillary reverse-phase high-performance liquid chromatography (HPLC) nano-electrospray tandem mass spectrometry (μLC–MS-MS) of a pigmented band excised from a denaturing polyacrylamide gel. The predicted protein sequence is related to a large group of octaheme cytochromes whose few characterized representatives are hydroxylamine or hydrazine oxidases. The protein was partially purified and shown by in vitro assays to have hydroxylamine oxidase, hydrazine oxidase, and nitrite reductase activities. From what is known of Beggiatoaceae physiology, nitrite reduction is the most likely in vivo role of the octaheme protein, but future experiments are required to confirm this tentative conclusion. Thus, while present-day genomic and proteomic techniques have allowed precise identification of an abundant mat protein, and its potential activities could be assayed, proof of its physiological role remains elusive in the absence of a pure culture that can be genetically manipulated.REFERENCES1.Jannasch HW, Nelson DC, and Wirsen CO. 1989. Massive natural occurrence of unusually large bacteria (Beggiatoa sp.) at a hydrothermal deep-sea vent site.Nature 342:834–836.Google Scholar2.Albertin ML. 1989. Interpretations and analysis of Guaymas Basin multi-channel seismic reflection profiles: implications for tectonic history. M.A. thesis. University of Texas at Austin, Austin, TX.Google Scholar3.Aragón-Arreola M, Morandi M, Martín-Barajas A, Delgado-Argote L, and González-Fernández A. 2005. Structure of the rift basins in the central Gulf of California: kinematic implications for oblique rifting. Tectonophysics 409:19–38.Google Scholar4.McKay LJ, MacGregor BJ, Biddle JF, Albert DB, Mendlovitz HP, Hoer DR, Lipp JS, Lloyd KG, and Teske AP. 2012. Spatial heterogeneity and underlying geochemistry of phylogenetically diverse orange and white Beggiatoa mats in Guaymas Basin hydrothermal sediments. Deep Sea Res. 67:21–31.Google Scholar5.Hinck S, Mußmann M, Salman V, Neu TR, Lenk S, de Beer D, and Jonkers HM. 2011. Vacuolated Beggiatoa-like filaments from different hypersaline environments form a novel genus. Environ. Microbiol. 13:3194–3205.PubMedGoogle Scholar6.Otte S, Kuenen JG, Nielsen LP, Paerl HW, Zopfi J, Schulz HN, Teske A, Strotmann B, Gallardo VA, and Jørgensen BB. 1999. Nitrogen, carbon, and sulfur metabolism in natural Thioploca samples. Appl. Environ. Microbiol. 65:3148–3157.CrossrefPubMedGoogle Scholar7.McHatton SC, Barry JP, Jannasch HW, and Nelson DC. 1996. High nitrate concentrations in vacuolate, autotrophic marine Beggiatoa spp. Appl. Environ. Microbiol. 62:954–958.CrossrefPubMedGoogle Scholar8.de Albuquerque JP, Keim CN, and Lins U. 2010. Comparative analysis of Beggiatoa from hypersaline and marine environments. Micron 41:507–517.PubMedGoogle Scholar9.Hinck S, Neu TR, Lavik G, Mussmann M, De Beer D, and Jonkers HM. 2007. Physiological adaptation of a nitrate-storing Beggiatoa sp. to diel cycling in a phototrophic hypersaline mat. Appl. Environ. Microbiol. 73:7013–7022.CrossrefPubMedGoogle Scholar10.Nelson DC, Wirsen CO, and Jannasch HW. 1989. Characterization of large, autotrophic Beggiatoa spp. abundant at hydrothermal vents of the Guaymas Basin. Appl. Environ. Microbiol. 55:2909–2917.CrossrefPubMedGoogle Scholar11.Prince RC, Stokley KE, Haith CE, and Jannasch HW. 1988. The cytochromes of a marine Beggiatoa. Arch. Microbiol. 150:193–196.Google Scholar12.Nikolaus R, Ammerman JW, and MacDonald IR. 2003. Distinct pigmentation and trophic modes in Beggiatoa from hydrocarbon seeps in the Gulf of Mexico. Aquatic Microb. Ecol. 32:85–93.Google Scholar13.Markowitz VM, Mavromatis K, Ivanova NN, Chen I-MA, Chu K, and Kyrpides NC. 2009. IMG ER: a system for microbial genome annotation expert review and curation. Bioinformatics 25:2271–2278.CrossrefPubMedGoogle Scholar14.Aziz RK, Bartels D, Best AA, DeJongh M, Disz T, Edwards RA, Formsma K, Gerdes S, Glass EM, Kubal M, Meyer F, Olsen GJ, Olson R, Osterman AL, Overbeek RA, McNeil LK, Paarmann D, Paczian T, Parrello B, Pusch GD, Reich C, Stevens R, Vassieva O, Vonstein V, Wilke A, and Zagnitko O. 2008. The RAST server: rapid annotations using subsystems technology. BMC Genomics 9:75.PubMedGoogle Scholar15.Eng JK, McCormack AL, and Yates JR. 1994. An approach to correlate tandem mass spectral data of peptides with amino acid sequences in a protein database. J. Am. Soc. Mass Spectrom. 5:976–989.CrossrefPubMedGoogle Scholar16.Chittum HS, Lane WS, Carlson BA, Roller PP, Lung FD, Lee BJ, and Hatfield DL. 1998. Rabbit β-globin is extended beyond its UGA stop codon by multiple suppressions and translational reading gaps. Biochemistry 37:10866–10870.PubMedGoogle Scholar17.Letunic I, Doerks T, and Bork P. 2009. SMART 6: recent updates and new developments. Nucleic Acids Res. 37:D229–D232.CrossrefPubMedGoogle Scholar18.Schultz J, Milpetz F, Bork P, and Ponting CP. 1998. SMART, a simple modular architecture research tool: identification of signaling domains. Proc. Natl. Acad. Sci. U. S. A. 95:5857–5864.CrossrefPubMedGoogle Scholar19.Yu NY, Wagner JR, Laird MR, Melli G, Rey S, Lo R, Dao P, Sahinalp SC, Ester M, Foster LJ, and Brinkman FSL. 2010. PSORTb 3.0: improved protein subcellular localization prediction with refined localization subcategories and predictive capabilities for all prokaryotes. Bioinformatics 26:1608–1615.CrossrefPubMedGoogle Scholar20.Nejidat A, Shmuely H, and Abeliovich A. 1997. Effect of ammonia starvation on hydroxylamine oxidoreductase activity of Nitrosomonas europaea. J. Biochem. 121:957–960.PubMedGoogle Scholar21.Tong Y and Guo ML. 2007. Cloning and characterization of a novel periplasmic heme-transport protein from the human pathogen Pseudomonas aeruginosa. J. Biol. Inorg. Chem. 12:735–750.PubMedGoogle Scholar22.Schalk J, de Vries S, Kuenen JG, and Jetten MSM. 2000. Involvement of a novel hydroxylamine oxidoreductase in anaerobic ammonium oxidation. Biochemistry 39:5405–5412.PubMedGoogle Scholar23.Shimamura M, Nishiyama T, Shigetomo H, Toyomoto T, Kawahara Y, Furukawa K, and Fujii T. 2007. Isolation of a multiheme protein with features of a hydrazine-oxidizing enzyme from an anaerobic ammonium-oxidizing enrichment culture. Appl. Environ. Microbiol. 73:1065–1072.CrossrefPubMedGoogle Scholar24.Shimamura M, Nishiyama T, Shinya K, Kawahara Y, Furukawa K, and Fujii T. 2008. Another multiheme protein, hydroxylamine oxidoreductase, abundantly produced in an anammox bacterium besides the hydrazine-oxidizing enzyme. J. Biosci. Bioeng. 105:243–248.PubMedGoogle Scholar25.Kostera J, McGarry J, and Pacheco AA. 2010. Enzymatic interconversion of ammonia and nitrite: the right tool for the job. Biochemistry 49:8546–8553.PubMedGoogle Scholar26.Stothard P. 2000. The Sequence Manipulation Suite: JavaScript programs for analyzing and formatting protein and DNA sequences. Biotechniques 28:1102–1104.CrossrefPubMedGoogle Scholar27.National Center for Biotechnology Information. 30 March 2012 posting date. PubChem compound database. http://www.ncbi.nlm.nih.gov/pccompound.Google Scholar28.Walsh DA, Zaikova E, Howes CG, Song YC, Wright JJ, Tringe SG, Tortell PD, and Hallam SJ. 2009. Metagenome of a versatile chemolithoautotroph from expanding oceanic dead zones. Science 326:578–582.PubMedGoogle Scholar29.Klotz MG, Schmid MC, Strous M, den Camp HJMO, Jetten MSM, and Hooper AB. 2008. Evolution of an octahaem cytochrome c protein family that is key to aerobic and anaerobic ammonia oxidation by bacteria. Environ. Microbiol. 10:3150–3163.PubMedGoogle Scholar30.Friedrich T and Weiss H. 1997. Modular evolution of the respiratory NADH:ubiquinone oxidoreductase and the origin of its modules. J. Theor. Biol. 187:529–540.PubMedGoogle Scholar31.Moparthi VK and Hagerhall C. 2011. The evolution of respiratory chain complex I from a smaller last common ancestor consisting of 11 protein subunits. J. Mol. Evol. 72:484–497.PubMedGoogle Scholar32.de Almeida NM, Maalcke WJ, Keltjens JT, Jetten MSM, and Kartal B. 2011. Proteins and protein complexes involved in the biochemical reactions of anaerobic ammonium-oxidizing bacteria. Biochem. Soc. Trans. 39:303–308.PubMedGoogle Scholar33.Junier P, Molina V, Dorador C, Hadas O, Kim OS, Junier T, Witzel K-P, and Imhoff JF. 2010. Phylogenetic and functional Marker genes to study ammonia-oxidizing microorganisms (AOM) in the environment. Appl. Microbiol. Biotechnol. 85:425–440.PubMedGoogle Scholar34.Tikhonova TV, Slutsky A, Antipov AN, Boyko KM, Polyakov KM, Sorokin DY, Zvyagilskaya RA, and Popov AN. 2006. Molecular and catalytic properties of a novel cytochrome c nitrite reductase from nitrate-reducing haloalkaliphilic sulfur-oxidizing bacterium Thioalkalivibrio nitratireducens. Biochim. Biophys. Acta 1764:715–723.PubMedGoogle Scholar35.Atkinson SJ, Mowat CG, Reid GA, and Chapman SK. 2007. An octaheme c-type cytochrome from Shewanella oneidensis can reduce nitrite and hydroxylamine. FEBS Lett. 581:3805–3808.PubMedGoogle Scholar36.Mowat CG, Rothery E, Miles CS, McIver L, Doherty MK, Drewette K, Taylor P, Walkinshaw MD, Chapman SK, and Reid GA. 2004. Octaheme tetrathionate reductase is a respiratory enzyme with novel heme ligation. Nat. Struct. Mol. Biol. 11:1023–1024.PubMedGoogle Scholar37.Polerecky L, Bissett A, Al-Najjar M, Faerber P, Osmers H, Suci PA, Stoodley P, and de Beer D. 2009. Modular spectral imaging system for discrimination of pigments in cells and microbial communities. Appl. Environ. Microbiol. 75:758–771.CrossrefPubMedGoogle Scholar38.Jannasch HW, Wheat CG, Plant JN, Kastner M, and Stakes DS. 2004. Continuous chemical monitoring with osmotically pumped water samplers: OsmoSampler design and applications. Limnol. Oceanogr. Methods 2:102–113.Google Scholar39.Schobert M and Dieter J. 2002. Regulation of heme biosynthesis in non-phototrophic bacteria. J. Mol. Microbiol. Biotechnol. 4:287–294.PubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 79 • Number 4 • 15 February 2013Pages: 1183 - 1190HistoryReceived: 17 August 2012Accepted: 3 December 2012Published online: 4 February 2013Copyright© 2013 American Society for Microbiology.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsBarbara J. MacGregorDepartment of Marine Sciences, University of North Carolina—Chapel Hill, Chapel Hill, North Carolina, USAView all articles by this authorJennifer F. BiddleCollege of Earth, Ocean, and the Environment, University of Delaware, Lewes, Delaware, USAView all articles by this authorJason R. SiebertDepartment of Biochemistry Molecular Biology, Michigan State University, East Lansing, Michigan, USAView all articles by this authorEric StauntonDepartment of Environmental and Health Sciences, University of North Carolina—Chapel Hill, Chapel Hill, North Carolina, USAView all articles by this authorEric L. HeggDepartment of Biochemistry Molecular Biology, Michigan State University, East Lansing, Michigan, USAView all articles by this authorAnn G. MatthysseDepartment of Biology, University of North Carolina—Chapel Hill, Chapel Hill, North Carolina, USAView all articles by this authorAndreas TeskeDepartment of Marine Sciences, University of North Carolina—Chapel Hill, Chapel Hill, North Carolina, USAView all articles by this authorNotesAddress correspondence to Barbara J. MacGregor, [email protected].Metrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleOctober 2019Plastics: Environmental and Biotechnological Perspectives on Microbial Degradation Dominik Danso, Jennifer Chowand Wolfgang R. StreitPlastics: Environmental and Biotechnological Perspectives on Microbial DegradationAuthors: Dominik Danso https://orcid.org/0000-0003-2214-2002, Jennifer Chow https://orcid.org/0000-0002-7499-5325, and Wolfgang R. Streit https://orcid.org/0000-0001-7617-7396DOI: https://doi.org/10.1128/AEM.01095-19Volume 85, Number 191 October 2019ABSTRACTREFERENCESABSTRACTPlastics are widely used in the global economy, and each year, at least 350 to 400 million tons are being produced. Due to poor recycling and low circular use, millions of tons accumulate annually in terrestrial or marine environments. Today it has become clear that plastic causes adverse effects in all ecosystems and that microplastics are of particular concern to our health. Therefore, recent microbial research has addressed the question of if and to what extent microorganisms can degrade plastics in the environment. This review summarizes current knowledge on microbial plastic degradation. Enzymes available act mainly on the high-molecular-weight polymers of polyethylene terephthalate (PET) and ester-based polyurethane (PUR). Unfortunately, the best PUR- and PET-active enzymes and microorganisms known still have moderate turnover rates. While many reports describing microbial communities degrading chemical additives have been published, no enzymes acting on the high-molecular-weight polymers polystyrene, polyamide, polyvinylchloride, polypropylene, ether-based polyurethane, and polyethylene are known. Together, these polymers comprise more than 80% of annual plastic production. Thus, further research is needed to significantly increase the diversity of enzymes and microorganisms acting on these polymers. This can be achieved by tapping into the global metagenomes of noncultivated microorganisms and dark matter proteins. Only then can novel biocatalysts and organisms be delivered that allow rapid degradation, recycling, or value-added use of the vast majority of most human-made polymers.REFERENCES1.Geyer R, Jambeck JR, Law KL. 2017. Production, use, and fate of all plastics ever made. Sci Adv 3:e1700782.CrossrefPubMedGoogle Scholar2.PlasticsEurope. 2018. PlasticsEurope, plastics—the facts 2018: an analysis of European plastics production, demand and waste data. PlasticsEurope, Brussels, Belgium.Google Scholar3.Ellen MacArthur Foundation. 2017. The new plastics economy: rethinking the future of plastics and catalysing action. Ellen MacArthur Foundation, Cowes, United Kingdom.Google Scholar4.Jambeck JR, Geyer R, Wilcox C, Siegler TR, Perryman M, Andrady A, Narayan R, Law KL. 2015. Marine pollution. Plastic waste inputs from land into the ocean. Science 347:768–771.CrossrefPubMedGoogle Scholar5.Derraik J. 2002. The pollution of the marine environment by plastic debris: a review. Mar Pollut Bull 44:842–852.CrossrefPubMedGoogle Scholar6.Cózar A, Echevarría F, González-Gordillo JI, Irigoien X, Úbeda B, Hernández-León S, Palma ÁT, Navarro S, García-de-Lomas J, Ruiz A, Fernández-de-Puelles ML, Duarte CM. 2014. Plastic debris in the open ocean. Proc Natl Acad Sci U S A 111:10239–10244.CrossrefPubMedGoogle Scholar7.Lebreton L, Slat B, Ferrari F, Sainte-Rose B, Aitken J, Marthouse R, Hajbane S, Cunsolo S, Schwarz A, Levivier A, Noble K, Debeljak P, Maral H, Schoeneich-Argent R, Brambini R, Reisser J. 2018. Evidence that the Great Pacific Garbage Patch is rapidly accumulating plastic. Sci Rep 8:4666.CrossrefPubMedGoogle Scholar8.Wei R, Zimmermann W. 2017. Microbial enzymes for the recycling of recalcitrant petroleum-based plastics: how far are we? Microb Biotechnol 10:1308.CrossrefPubMedGoogle Scholar9.Day M, Wiles DM. 1972. Photochemical degradation of poly(ethylene terephthalate). II. Effect of wavelength and environment on the decomposition process. J Appl Polym Sci 16:191–202.CrossrefGoogle Scholar10.Mohammadian M, Allen NS, Edge M, Jones K. 1991. Environmental degradation of poly (ethylene terephthalate). Textile Res J 61:690–696.CrossrefGoogle Scholar11.Welzel K, Müller RJ, Deckwer WD. 2002. Enzymatischer Abbau von Polyester-Nanopartikeln. Chemie Ingenieur Technik 74:1496–1500.CrossrefGoogle Scholar12.Smith M, Love DC, Rochman CM, Neff RA. 2018. Microplastics in seafood and the implications for human health. Curr Environ Health Rep 5:375–386.CrossrefPubMedGoogle Scholar13.de Souza Machado AA, Kloas W, Zarfl C, Hempel S, Rillig MC. 2018. Microplastics as an emerging threat to terrestrial ecosystems. Glob Chang Biol 24:1405–1416.CrossrefPubMedGoogle Scholar14.Gubbels E, Heitz T, Yamamoto M, Chilekar V, Zarbakhsh S, Gepraegs M, Köpnick H, Schmidt M, Brügging W, Rüter J, Kaminsky W. 2018. Polyesters. In Ullmann’s encyclopedia of industrial chemistry. Wiley-VCH, Weinheim, Germany.CrossrefGoogle Scholar15.Acero EH, Ribitsch D, Steinkellner G, Gruber K, Greimel K, Eiteljoerg I, Trotscha E, Wei R, Zimmermann W, Zinn M, Cavaco-Paulo A, Freddi G, Schwab H, Guebitz G. 2011. Enzymatic surface hydrolysis of PET: effect of structural diversity on kinetic properties of cutinases from Thermobifida. Macromolecules 44:4632–4640.CrossrefGoogle Scholar16.Kleeberg I, Hetz C, Kroppenstedt RM, Muller RJ, Deckwer WD. 1998. Biodegradation of aliphatic-aromatic copolyesters by Thermomonospora fusca and other thermophilic compost isolates. Appl Environ Microbiol 64:1731–1735.CrossrefPubMedGoogle Scholar17.Hu X, Thumarat U, Zhang X, Tang M, Kawai F. 2010. Diversity of polyester-degrading bacteria in compost and molecular analysis of a thermoactive esterase from Thermobifida alba AHK119. Appl Microbiol Biotechnol 87:771–779.CrossrefPubMedGoogle Scholar18.Wei R, Oeser T, Then J, Kuhn N, Barth M, Schmidt J, Zimmermann W. 2014. Functional characterization and structural modeling of synthetic polyester-degrading hydrolases from Thermomonospora curvata. AMB Express 4:44.CrossrefPubMedGoogle Scholar19.Wei R, Oeser T, Zimmermann W. 2014. Synthetic polyester-hydrolyzing enzymes from thermophilic actinomycetes. Adv Appl Microbiol 89:267–305.CrossrefPubMedGoogle Scholar20.Chen S, Tong X, Woodard RW, Du GC, Wu J, Chen J. 2008. Identification and characterization of bacterial cutinase. J Biol Chem 283:25854–25862.CrossrefPubMedGoogle Scholar21.Zimmermann W, Billig S. 2011. Enzymes for the biofunctionalization of poly(ethylene terephthalate). Adv Biochem Eng Biotechnol 125:97–120.CrossrefPubMedGoogle Scholar22.Ribitsch D, Acero EH, Greimel K, Dellacher A, Zitzenbacher S, Marold A, Rodriguez RD, Steinkellner G, Gruber K, Schwab H, Guebitz GM. 2012. A new esterase from Thermobifida halotolerans hydrolyses polyethylene terephthalate (PET) and polylactic acid (PLA). Polymers 4:617–629.CrossrefGoogle Scholar23.Kawai F, Oda M, Tamashiro T, Waku T, Tanaka N, Yamamoto M, Mizushima H, Miyakawa T, Tanokura M. 2014. A novel Ca2+-activated, thermostabilized polyesterase capable of hydrolyzing polyethylene terephthalate from Saccharomonospora viridis AHK190. Appl Microbiol Biotechnol 98:10053–10064.CrossrefPubMedGoogle Scholar24.Ollis DL, Cheah E, Cygler M, Dijkstra B, Frolow F, Franken SM, Harel M, Remington SJ, Silman I, Schrag J. 1992. The alpha/beta hydrolase fold. Protein Eng 5:197–211.CrossrefPubMedGoogle Scholar25.Yoshida S, Hiraga K, Takehana T, Taniguchi I, Yamaji H, Maeda Y, Toyohara K, Miyamoto K, Kimura Y, Oda K. 2016. A bacterium that degrades and assimilates poly(ethylene terephthalate). Science 351:1196–1199.CrossrefPubMedGoogle Scholar26.Austin HP, Allen MD, Donohoe BS, Rorrer NA, Kearns FL, Silveira RL, Pollard BC, Dominick G, Duman R, El Omari K, Mykhaylyk V, Wagner A, Michener WE, Amore A, Skaf MS, Crowley MF, Thorne AW, Johnson CW, Woodcock HL, McGeehan JE, Beckham GT. 2018. Characterization and engineering of a plastic-degrading aromatic polyesterase. Proc Natl Acad Sci U S A 115:E4350–E4357.CrossrefPubMedGoogle Scholar27.Roth C, Wei R, Oeser T, Then J, Follner C, Zimmermann W, Strater N. 2014. Structural and functional studies on a thermostable polyethylene terephthalate degrading hydrolase from Thermobifida fusca. Appl Microbiol Biotechnol 98:7815–7823.CrossrefPubMedGoogle Scholar28.Sulaiman S, You DJ, Kanaya E, Koga Y, Kanaya S. 2014. Crystal structure and thermodynamic and kinetic stability of metagenome-derived LC-cutinase. Biochemistry 53:1858–1869.CrossrefPubMedGoogle Scholar29.Then J, Wei R, Oeser T, Gerdts A, Schmidt J, Barth M, Zimmermann W. 2016. A disulfide bridge in the calcium binding site of a polyester hydrolase increases its thermal stability and activity against polyethylene terephthalate. FEBS Open Bio 6:425–432.CrossrefPubMedGoogle Scholar30.Ribitsch D, Heumann S, Trotscha E, Acero EH, Greimel K, Leber R, Birner-Gruenberger R, Deller S, Eiteljoerg I, Remler P, Weber T, Siegert P, Maurer KH, Donelli I, Freddi G, Schwab H, Guebitz GM. 2011. Hydrolysis of polyethyleneterephthalate by p-nitrobenzylesterase from Bacillus subtilis. Biotechnol Progress 27:951–960.CrossrefPubMedGoogle Scholar31.Danso D, Schmeisser C, Chow J, Zimmermann W, Wei R, Leggewie C, Li X, Hazen T, Streit WR. 2018. New insights into the function and global distribution of polyethylene terephthalate (PET)-degrading bacteria and enzymes in marine and terrestrial metagenomes. Appl Environ Microbiol 84:e02773-17.CrossrefPubMedGoogle Scholar32.Palm GJ, Reisky L, Böttcher D, Müller H, Michels EAP, Walczak MC, Berndt L, Weiss MS, Bornscheuer UT, Weber G. 2019. Structure of the plastic-degrading Ideonella sakaiensis MHETase bound to a substrate. Nat Commun 10:1717.CrossrefPubMedGoogle Scholar33.Barth M, Oeser T, Wei R, Then J, Schmidt J, Zimmermann W. 2015. Effect of hydrolysis products on the enzymatic degradation of polyethylene terephthalate nanoparticles by a polyester hydrolase from Thermobifida fusca. Biochem Eng J 93:222–228.CrossrefGoogle Scholar34.Carniel A, Valoni E, Nicomedes J, Gomes AD, de Castro AM. 2017. Lipase from Candida antarctica (CALB) and cutinase from Humicola insolens act synergistically for PET hydrolysis to terephthalic acid. Process Biochem 59:84–90.CrossrefGoogle Scholar35.Wei R, Oeser T, Schmidt J, Meier R, Barth M, Then J, Zimmermann W. 2016. Engineered bacterial polyester hydrolases efficiently degrade polyethylene terephthalate due to relieved product inhibition. Biotechnol Bioeng 113:1658–1665.CrossrefPubMedGoogle Scholar36.Haernvall K, Zitzenbacher S, Wallig K, Yamamoto M, Schick MB, Ribitsch D, Guebitz GM. 2017. Hydrolysis of ionic phthalic acid based polyesters by wastewater microorganisms and their enzymes. Environ Sci Technol 51:4596–4605.CrossrefPubMedGoogle Scholar37.Bollinger A, Thies S, Katzke N, Jaeger KE. 25 June 2018. The biotechnological potential of marine bacteria in the novel lineage of Pseudomonas pertucinogena. Microb Biotechnol doi:CrossrefPubMedGoogle Scholar38.Hajighasemi M, Tchigvintsev A, Nocek BP, Flick R, Popovic A, Hai T, Khusnutdinova AN, Brown G, Xu X, Cui H, Anstett J, Chernikova TN, Bruls T, Le Paslier D, Yakimov MM, Joachimiak A, Golyshina OV, Savchenko A, Golyshin PN, Edwards EA, Yakunin AF. 2018. Screening and characterization of novel polyesterases from environmental metagenomes with high hydrolytic activity against synthetic polyesters. Environ Sci Technol 52:12388–12401.CrossrefPubMedGoogle Scholar39.Seymour RB, Kauffman GB. 1992. Polyurethanes: a class of modern versatile materials. J Chem Educ 69:909.CrossrefGoogle Scholar40.Darby RT, Kaplan AM. 1968. Fungal susceptibility of polyurethanes. Appl Microbiol 16:900–905.CrossrefPubMedGoogle Scholar41.Russell JR, Huang J, Anand P, Kucera K, Sandoval AG, Dantzler KW, Hickman D, Jee J, Kimovec FM, Koppstein D, Marks DH, Mittermiller PA, Núñez SJ, Santiago M, Townes MA, Vishnevetsky M, Williams NE, Vargas MPN, Boulanger L-A, Bascom-Slack C, Strobel SA. 2011. Biodegradation of polyester polyurethane by endophytic fungi. Appl Environ Microbiol 77:6076–6084.CrossrefPubMedGoogle Scholar42.Howard GT, Crother B, Vicknair J. 2001. Cloning, nucleotide sequencing and characterization of a polyurethanase gene (pueB) from Pseudomonas chlororaphis. Int Biodeterior Biodegrad 47:141–149.CrossrefGoogle Scholar43.Howard GT, Blake RC. 1998. Growth of Pseudomonas fluorescens on a polyester–polyurethane and the purification and characterization of a polyurethanase–protease enzyme. Int Biodeterior Biodegrad 42:213–220.CrossrefGoogle Scholar44.Stern RV, Howard GT. 2000. The polyester polyurethanase gene (pueA) from Pseudomonas chlororaphis encodes a lipase. FEMS Microbiol Lett 185:163–168.CrossrefPubMedGoogle Scholar45.Howard GT, Mackie RI, Cann IK, Ohene-Adjei S, Aboudehen KS, Duos BG, Childers GW. 2007. Effect of insertional mutations in the pueA and pueB genes encoding two polyurethanases in Pseudomonas chlororaphis contained within a gene cluster. J Appl Microbiol 103:2074–2083.CrossrefPubMedGoogle Scholar46.Hung CS, Zingarelli S, Nadeau LJ, Biffinger JC, Drake CA, Crouch AL, Barlow DE, Russell JN, Jr, Crookes-Goodson WJ. 2016. Carbon catabolite repression and impranil polyurethane degradation in Pseudomonas protegens strain Pf-5. Appl Environ Microbiol 82:6080–6090.CrossrefPubMedGoogle Scholar47.Peng YH, Shih YH, Lai YC, Liu YZ, Liu YT, Lin NC. 2014. Degradation of polyurethane by bacterium isolated from soil and assessment of polyurethanolytic activity of a Pseudomonas putida strain. Environ Sci Pollut Res Int 21:9529–9537.CrossrefPubMedGoogle Scholar48.Akutsu Y, Nakajima-Kambe T, Nomura N, Nakahara T. 1998. Purification and properties of a polyester polyurethane-degrading enzyme from Comamonas acidovorans TB-35. Appl Environ Microbiol 64:62–67.CrossrefPubMedGoogle Scholar49.Shigeno-Akutsu Y, Nakajima-Kambe T, Nomura N, Nakahara T. 1999. Purification and properties of culture-broth-secreted esterase from the polyurethane degrader Comamonas acidovorans TB-35. J Biosci Bioeng 88:484–487.CrossrefPubMedGoogle Scholar50.Biffinger JC, Barlow DE, Cockrell AL, Cusick KD, Hervey WJ, Fitzgerald LA, Nadeau LJ, Hung CS, Crookes-Goodson WJ, Russell JN. 2015. The applicability of Impranil® DLN for gauging the biodegradation of polyurethanes. Polym Degradation Stab 120:178–185.CrossrefGoogle Scholar51.Shah Z, Krumholz L, Aktas DF, Hasan F, Khattak M, Shah AA. 2013. Degradation of polyester polyurethane by a newly isolated soil bacterium, Bacillus subtilis strain MZA-75. Biodegradation 24:865–877.CrossrefPubMedGoogle Scholar52.Rowe L, Howard GT. 2002. Growth of Bacillus subtilis on polyurethane and the purification and characterization of a polyurethanase-lipase enzyme. Int Biodeterior Biodegrad 50:33–40.CrossrefGoogle Scholar53.Oceguera-Cervantes A, Carrillo-García A, López N, Bolaños-Nuñez S, Cruz-Gómez MJ, Wacher C, Loza-Tavera H. 2007. Characterization of the polyurethanolytic activity of two Alicycliphilus sp. strains able to degrade polyurethane and n-methylpyrrolidone. Appl Environ Microbiol 73:6214–6223.CrossrefPubMedGoogle Scholar54.Schmidt J, Wei R, Oeser T, Dedavid e Silva L, Breite D, Schulze A, Zimmermann W. 2017. Degradation of polyester polyurethane by bacterial polyester hydrolases. Polymers 9:65.CrossrefGoogle Scholar55.Martinez-Martinez M, Coscolin C, Santiago G, Chow J, Stogios PJ, Bargiela R, Gertler C, Navarro-Fernandez J, Bollinger A, Thies S, Mendez-Garcia C, Popovic A, Brown G, Chernikova TN, Garcia-Moyano A, Bjerga GEK, Perez-Garcia P, Hai T, Del Pozo MV, Stokke R, Steen IH, Cui H, Xu X, Nocek BP, Alcaide M, Distaso M, Mesa V, Pelaez AI, Sanchez J, Buchholz PCF, Pleiss J, Fernandez-Guerra A, Glockner FO, Golyshina OV, Yakimov MM, Savchenko A, Jaeger KE, Yakunin AF, Streit WR, Golyshin PN, Guallar V, Ferrer M, The Inmare Consortium. 2018. Determinants and prediction of esterase substrate promiscuity patterns. ACS Chem Biol 13:225–234.CrossrefPubMedGoogle Scholar56.Zafar U, Houlden A, Robson GD. 2013. Fungal communities associated with the biodegradation of polyester polyurethane buried under compost at different temperatures. Appl Environ Microbiol 79:7313–7324.CrossrefPubMedGoogle Scholar57.Gautam R, Bassi AS, Yanful EK. 2007. Candida rugosa lipase-catalyzed polyurethane degradation in aqueous medium. Biotechnol Lett 29:1081–1086.CrossrefPubMedGoogle Scholar58.Álvarez-Barragán J, Domínguez-Malfavón L, Vargas-Suárez M, González-Hernández R, Aguilar-Osorio G, Loza-Tavera H. 2016. Biodegradative activities of selected environmental fungi on a polyester polyurethane varnish and polyether polyurethane foams. Appl Environ Microbiol 82:5225–5235.CrossrefPubMedGoogle Scholar59.Mathur G, Prasad R. 2012. Degradation of polyurethane by Aspergillus flavus (ITCC 6051) isolated from soil. Appl Biochem Biotechnol 167:1595–1602.CrossrefPubMedGoogle Scholar60.Khan S, Nadir S, Shah ZU, Shah AA, Karunarathna SC, Xu J, Khan A, Munir S, Hasan F. 2017. Biodegradation of polyester polyurethane by Aspergillus tubingensis. Environ Pollut 225:469–480.CrossrefPubMedGoogle Scholar61.Nowlin T. 2014. Global polyethylene business overview. In Nowlin TE (ed), Business and technology of the global polyethylene industry. Wiley-VCH, Weinheim, Germany.CrossrefGoogle Scholar62.Sen SK, Raut S. 2015. Microbial degradation of low density polyethylene (LDPE): a review. J Environ Chem Eng 3:462–473.CrossrefGoogle Scholar63.Restrepo-Florez JM, Bassi A, Thompson MR. 2014. Microbial degradation and deterioration of polyethylene—a review. Int Biodeterior Biodegrad 88:83–90.CrossrefGoogle Scholar64.Pathak VM, Navneet. 2017. Review on the current status of polymer degradation: a microbial approach. Bioresource Bioprocess 4:15.CrossrefGoogle Scholar65.Ojha N, Pradhan N, Singh S, Barla A, Shrivastava A, Khatua P, Rai V, Bose S. 2017. Evaluation of HDPE and LDPE degradation by fungus, implemented by statistical optimization. Sci Rep 7:39515.CrossrefPubMedGoogle Scholar66.Yamada-Onodera K, Mukumoto H, Katsuyaya Y, Saiganji A, Tani Y. 2001. Degradation of polyethylene by a fungus, Penicillium simplicissimum YK. Polym Degradation Stab 72:323–327.CrossrefGoogle Scholar67.Bonhomme S, Cuer A, Delort A, Lemaire J, Sancelme M, Scott G. 2003. Environmental biodegradation of polyethylene. Polym Degradation Stab 81:441–452.CrossrefGoogle Scholar68.Veethahavya KS, Rajath BS, Noobia S, Kumar BM. 2016. Biodegradation of low density polyethylene in aqueous media. Procedia Environ Sci 35:709–713.CrossrefGoogle Scholar69.Vimala PP, Mathew L. 2016. Biodegradation of polyethylene using Bacillus subtilis. Procedia Technol 24:232–239.CrossrefGoogle Scholar70.Yang J, Yang Y, Wu W-M, Zhao J, Jiang L. 2014. Evidence of polyethylene biodegradation by bacterial strains from the guts of plastic-eating waxworms. Environ Sci Technol 48:13776–13784.CrossrefPubMedGoogle Scholar71.Yang Y, Yang J, Wu W-M, Zhao J, Song Y, Gao L, Yang R, Jiang L. 2015. Biodegradation and mineralization of polystyrene by plastic-eating mealworms: Part 1. Chemical and physical characterization and isotopic tests. Environ Sci Technol 49:12080–12086.CrossrefPubMedGoogle Scholar72.Sowmya HV, Ramalingappa B, Krishnappa M, Thippeswamy B. 2015. Degradation of polyethylene by Penicillium simplicissimum isolated from local dumpsite of Shivamogga district. Environ Dev Sustain 17:731–745.CrossrefGoogle Scholar73.Palmer R. 2001. Polyamides, plastics. In Encyclopedia of polymer science and technology. Wiley, Hoboken, NJ.CrossrefGoogle Scholar74.Tosa T, Chibata I. 1965. Utilization of cyclic amides and formation of omega-amino acids by microorganisms. J Bacteriol 89:919–920.CrossrefPubMedGoogle Scholar75.Takehara I, Kato DI, Takeo M, Negoro S. 2017. Draft genome sequence of the nylon oligomer-degrading bacterium Arthrobacter sp. strain KI72. Genome Announc 5:e00217-17.CrossrefPubMedGoogle Scholar76.Negoro S, Taniguchi T, Kanaoka M, Kimura H, Okada H. 1983. Plasmid-determined enzymatic degradation of nylon oligomers. J Bacteriol 155:22–31.CrossrefPubMedGoogle Scholar77.Negoro S, Kakudo S, Urabe I, Okada H. 1992. A new nylon oligomer degradation gene (nylC) on plasmid pOAD2 from a Flavobacterium sp. J Bacteriol 174:7948–7953.CrossrefPubMedGoogle Scholar78.Kakudo S, Negoro S, Urabe I, Okada H. 1993. Nylon oligomer degradation gene, nylC, on plasmid pOAD2 from a Flavobacterium strain encodes endo-type 6-aminohexanoate oligomer hydrolase: purification and characterization of the nylC gene product. Appl Environ Microbiol 59:3978.CrossrefPubMedGoogle Scholar79.Negoro S, Ohki T, Shibata N, Sasa K, Hayashi H, Nakano H, Yasuhira K, Kato D-i, Takeo M, Higuchi Y. 2007. Nylon-oligomer degrading enzyme/substrate complex: catalytic mechanism of 6-aminohexanoate-dimer hydrolase. J Mol Biol 370:142–156.CrossrefPubMedGoogle Scholar80.Yasuhira K, Uedo Y, Shibata N, Negoro S, Takeo M, Higuchi Y. 2006. Crystallization and X-ray diffraction analysis of 6-aminohexanoate-cyclic-dimer hydrolase from Arthrobacter sp. KI72. Acta Crystallogr Sect F Struct Biol Cryst Commun 62:1209–1211.CrossrefPubMedGoogle Scholar81.Ohki T, Mizuno N, Shibata N, Takeo M, Negoro S, Higuchi Y. 2005. Crystallization and X-ray diffraction analysis of 6-aminohexanoate-dimer hydrolase from Arthrobacter sp. KI72. Acta Crystallogr Sect F Struct Biol Cryst Commun 61:928–930.CrossrefPubMedGoogle Scholar82.Nagai K, Yasuhira K, Tanaka Y, Kato D, Takeo M, Higuchi Y, Negoro S, Shibata N. 2013. Crystallization and X-ray diffraction analysis of nylon hydrolase (NylC) from Arthrobacter sp. KI72. Acta Crystallogr Sect F Struct Biol Cryst Commun 69:1151–1154.CrossrefPubMedGoogle Scholar83.Kinoshita S, Terada T, Taniguchi T, Takene Y, Masuda S, Matsunaga N, Okada H. 1981. Purification and characterization of 6-aminohexanoic-acid-oligomer hydrolase of Flavobacterium sp. Ki72. Eur J Biochem 116:547–551.CrossrefPubMedGoogle Scholar84.Yasuhira K, Tanaka Y, Shibata H, Kawashima Y, Ohara A, Kato D, Takeo M, Negoro S. 2007. 6-Aminohexanoate oligomer hydrolases from the alkalophilic bacteria Agromyces sp. strain KY5R and Kocuria sp. strain KY2. Appl Environ Microbiol 73:7099–7102.CrossrefPubMedGoogle Scholar85.Negoro S, Ohki T, Shibata N, Mizuno N, Wakitani Y, Tsurukame J, Matsumoto K, Kawamoto I, Takeo M, Higuchi Y. 2005. X-ray crystallographic analysis of 6-aminohexanoate-dimer hydrolase: molecular basis for the birth of a nylon oligomer-degrading enzyme. J Biol Chem 280:39644–39652.CrossrefPubMedGoogle Scholar86.Kinoshita S, Negoro S, Muramatsu M, Bisaria VS, Sawada S, Okada H. 1977. 6-Aminohexanoic acid cyclic dimer hydrolase. A new cyclic amide hydrolase produced by Achromobacter guttatus KI74. Eur J Biochem 80:489–495.CrossrefPubMedGoogle Scholar87.Kinoshita S, Kageyama S, Iba K, Yamada Y, Okada H. 1975. Utilization of a cyclic dimer and linear oligomers of ε-aminocaproic acid by Achromobacter guttatus KI 72. Agric Biol Chem 39:1219–1223.CrossrefGoogle Scholar88.Takehara I, Fujii T, Tanimoto Y, Kato DI, Takeo M, Negoro S. 2018. Correction to: Metabolic pathway of 6-aminohexanoate in the nylon oligomer-degrading bacterium Arthrobacter sp. KI72: identification of the enzymes responsible for the conversion of 6-aminohexanoate to adipate. Appl Microbiol Biotechnol 102:815.CrossrefPubMedGoogle Scholar89.Takehara I, Fujii T, Tanimoto Y, Kato DI, Takeo M, Negoro S. 2018. Metabolic pathway of 6-aminohexanoate in the nylon oligomer-degrading bacterium Arthrobacter sp. KI72: identification of the enzymes responsible for the conversion of 6-aminohexanoate to adipate. Appl Microbiol Biotechnol 102:801–814.CrossrefPubMedGoogle Scholar90.Sudhakar M, Priyadarshini C, Doble M, Sriyutha Murthy P, Venkatesan R. 2007. Marine bacteria mediated degradation of nylon 66 and 6. Int Biodeterior Biodegrad 60:144–151.CrossrefGoogle Scholar91.Oppermann FB, Pickartz S, Steinbüchel A. 1998. Biodegradation of polyamides. Polym Degradation Stab 59:337–344.CrossrefGoogle Scholar92.Deguchi T, Kitaoka Y, Kakezawa M, Nishida T. 1998. Purification and characterization of a nylon-degrading enzyme. Appl Environ Microbiol 64:1366–1371.CrossrefPubMedGoogle Scholar93.Prijambada ID, Negoro S, Yomo T, Urabe I. 1995. Emergence of nylon oligomer degradation enzymes in Pseudomonas aeruginosa PAO through experimental evolution. Appl Environ Microbiol 61:2020–2022.CrossrefPubMedGoogle Scholar94.Kanagawa K, Oishi M, Negoro S, Urabe I, Okada H. 1993. Characterization of the 6-aminohexanoate-dimer hydrolase from Pseudomonas sp. NK87. J Gen Microbiol 139:787–795.CrossrefPubMedGoogle Scholar95.Maul J, Frushour BG, Kontoff JR, Eichenauer H, Ott K-H, Schade C. 2007. Polystyrene and styrene copolymers. In Ullmann’s encyclopedia of industrial chemistry. Wiley-VCH, Weinheim, Germany.CrossrefGoogle Scholar96.Krueger MC, Hofmann U, Moeder M, Schlosser D. 2015. Potential of wood-rotting fungi to attack polystyrene sulfonate and its depolymerisation by Gloeophyllum trabeum via hydroquinone-driven fenton chemistry. PLoS One 10:e0131773.CrossrefPubMedGoogle Scholar97.Milstein O, Gersonde R, Huttermann A, Chen MJ, Meister JJ. 1992. Fungal biodegradation of lignopolystyrene graft copolymers. Appl Environ Microbiol 58:3225–3232.CrossrefPubMedGoogle Scholar98.Ho BT, Roberts TK, Lucas S. 2018. An overview on biodegradation of polystyrene and modified polystyrene: the microbial approach. Crit Rev Biotechnol 38:308–320.CrossrefPubMedGoogle Scholar99.Chauhan D, Agrawal G, Deshmukh S, Roy SS, Priyadarshini R. 2018. Biofilm formation by Exiguobacterium sp. DR11 and DR14 alter polystyrene surface properties and initiate biodegradation. RSC Adv 8:37590–37599.CrossrefGoogle Scholar100.Mooney A, Ward PG, O’Connor KE. 2006. Microbial degradation of styrene: biochemistry, molecular genetics, and perspectives for biotechnological applications. Appl Microbiol Biotechnol 72:1.CrossrefPubMedGoogle Scholar101.Dobson ADW, O’Leary ND, O\'Connor KE. 2002. Biochemistry, genetics and physiology of microbial styrene degradation. FEMS Microbiol Rev 26:403–417.CrossrefPubMedGoogle Scholar102.Tischler D. 2015. Microbial styrene degradation, p 7–22. Springer International Publishing, Cham, Switzerland.CrossrefGoogle Scholar103.Oelschlägel M, Zimmerling J, Tischler D. 2018. A review: the styrene metabolizing cascade of side-chain oxygenation as biotechnological basis to gain various valuable compounds. Front Microbiol 9:490.CrossrefPubMedGoogle Scholar104.Tischler D, Eulberg D, Lakner S, Kaschabek SR, van Berkel WJH, Schlomann M. 2009. Identification of a novel self-sufficient styrene monooxygenase from Rhodococcus opacus 1CP. J Bacteriol 191:4996–5009.CrossrefPubMedGoogle Scholar105.Velasco A, Alonso S, García JL, Perera J, Díaz E. 1998. Genetic and functional analysis of the styrene catabolic cluster of Pseudomonas sp. strain Y2. J Bacteriol 180:1063–1071.CrossrefPubMedGoogle Scholar106.Morrison E, Kantz A, Gassner GT, Sazinsky MH. 2013. Structure and mechanism of styrene monooxygenase reductase: new insight into the FAD-transfer reaction. Biochemistry 52:6063–6075.CrossrefPubMedGoogle Scholar107.Oelschlägel M, Gröning JAD, Tischler D, Kaschabek SR, Schlömann M. 2012. Styrene oxide isomerase of Rhodococcus opacus 1CP, a highly stable and considerably active enzyme. Appl Environ Microbiol 78:4330–4337.CrossrefPubMedGoogle Scholar108.Crabo AG, Singh B, Nguyen T, Emami S, Gassner GT, Sazinsky MH. 2017. Structure and biochemistry of phenylacetaldehyde dehydrogenase from the Pseudomonas putida S12 styrene catabolic pathway. Arch Biochem Biophys 616:47–58.CrossrefPubMedGoogle Scholar109.O\'Leary ND, O\'Mahony MM, Dobson AD. 2011. Regulation of phenylacetic acid uptake is sigma54 dependent in Pseudomonas putida CA-3. BMC Microbiol 11:229.CrossrefPubMedGoogle Scholar110.O’Leary ND, Duetz WA, Dobson AD, O’Connor KE. 2002. Induction and repression of the sty operon in Pseudomonas putida CA-3 during growth on phenylacetic acid under organic and inorganic nutrient-limiting continuous culture conditions. FEMS Microbiol Lett 208:263–268.CrossrefPubMedGoogle Scholar111.O’Leary ND, Mooney A, O\'Mahony M, Dobson AD. 2014. Functional characterization of a StyS sensor kinase reveals distinct domains associated with intracellular and extracellular sensing of styrene in P. putida CA-3. Bioengineered 5:114–122.CrossrefPubMedGoogle Scholar112.Sheldon RA, Van Bekkum H. 2008. Fine chemicals through heterogeneous catalysis. John Wiley Sons, Hoboken, NJ.Google Scholar113.O’Leary ND, O’Connor KE, Ward P, Goff M, Dobson AD. 2005. Genetic characterization of accumulation of polyhydroxyalkanoate from styrene in Pseudomonas putida CA-3. Appl Environ Microbiol 71:4380–4387.CrossrefPubMedGoogle Scholar114.Ward PG, Goff M, Donner M, Kaminsky W, O’Connor KE. 2006. A two step chemo-biotechnological conversion of polystyrene to a biodegradable thermoplastic. Environ Sci Technol 40:2433–2437.CrossrefPubMedGoogle Scholar115.Savoldelli J, Tomback D, Savoldelli H. 2017. Breaking down polystyrene through the application of a two-step thermal degradation and bacterial method to produce usable byproducts. Waste Manage 60:123–126.CrossrefPubMedGoogle Scholar116.Fischer I, Schmitt WF, Porth H, Allsopp MW, Vianello G. 2014. Poly(vinyl chloride). In Ullmann’s encyclopedia of industrial chemistry. Wiley‐VCH, Weinheim, Germany.CrossrefGoogle Scholar117.Karger-Kocsis J, Bárány T. 2019. Polypropylene handbook. Springer Nature Switzerland, Basel, Switzerland.CrossrefGoogle Scholar118.Cacciari I, Quatrini P, Zirletta G, Mincione E, Vinciguerra V, Lupattelli P, Giovannozzi Sermanni G. 1993. Isotactic polypropylene biodegradation by a microbial community: physicochemical characterization of metabolites produced. Appl Environ Microbiol 59:3695–3700.CrossrefPubMedGoogle Scholar119.Iakovlev VV, Guelcher SA, Bendavid R. 2017. Degradation of polypropylene in vivo: a microscopic analysis of meshes explanted from patients. J Biomed Mater Res B Appl Biomater 105:237–248.CrossrefPubMedGoogle Scholar120.Bombelli P, Howe CJ, Bertocchini F. 2017. Polyethylene bio-degradation by caterpillars of the wax moth Galleria mellonella. Curr Biol 27:R292–R293.CrossrefPubMedGoogle Scholar121.Weber C, Pusch S, Opatz T. 2017. Polyethylene bio-degradation by caterpillars? Curr Biol 27:R744–R745.CrossrefPubMedGoogle Scholar122.Yang Y, Chen J, Wu W-M, Zhao J, Yang J. 2015. Complete genome sequence of Bacillus sp. YP1, a polyethylene-degrading bacterium from waxworm’s gut. J Biotechnol 200:77–78.CrossrefPubMedGoogle Scholar123.Brandon AM, Gao S-H, Tian R, Ning D, Yang S-S, Zhou J, Wu W-M, Criddle CS. 2018. Biodegradation of polyethylene and plastic mixtures in mealworms (larvae of Tenebrio molitor) and effects on the gut microbiome. Environ Sci Technol 52:6526–6533.CrossrefPubMedGoogle Scholar124.Notredame C, Higgins DG, Heringa J. 2000. T-Coffee: a novel method for fast and accurate multiple sequence alignment. J Mol Biol 302:205–217.CrossrefPubMedGoogle Scholar125.Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. 2013. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Mol Biol Evol 30:2725–2729.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 85 • Number 19 • 1 October 2019eLocator: e01095-19Editor: Harold L. DrakeUniversity of BayreuthHistoryPublished online: 19 July 2019Copyright© 2019 Danso et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license.PermissionsRequest permissions for this article.Request PermissionsKEYWORDSPETcutinasemicrobial plastic degradationpolyamidespolyethylenepolyethylene terephthalatepolypropylenepolystyrenepolyurethanepolyvinylchlorideContributorsAuthorsDominik Danso https://orcid.org/0000-0003-2214-2002Department of Microbiology and Biotechnology, University of Hamburg, Hamburg, GermanyView all articles by this authorJennifer Chow https://orcid.org/0000-0002-7499-5325Department of Microbiology and Biotechnology, University of Hamburg, Hamburg, GermanyView all articles by this authorWolfgang R. Streit https://orcid.org/0000-0001-7617-7396Department of Microbiology and Biotechnology, University of Hamburg, Hamburg, GermanyView all articles by this authorEditorHarold L. DrakeEditorUniversity of BayreuthNotesAddress correspondence to Wolfgang R. Streit, [email protected].Metrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleDecember 2009Introducing mothur: Open-Source, Platform-Independent, Community-Supported Software for Describing and Comparing Microbial Communities Patrick D. Schloss, Sarah L. Westcott, Thomas Ryabin, Justine R. Hall, Martin Hartmann, Emily B. Hollister, Ryan A. Lesniewski, Brian B. Oakley, Donovan H. Parks, Courtney J. Robinson, Jason W. Sahl, Blaz Stres, Gerhard G. Thallinger, David J. Van Hornand Carolyn F. WeberIntroducing mothur: Open-Source, Platform-Independent, Community-Supported Software for Describing and Comparing Microbial CommunitiesAuthors: Patrick D. Schloss [email protected], Sarah L. Westcott, Thomas Ryabin, Justine R. Hall, Martin Hartmann, Emily B. Hollister, Ryan A. Lesniewski, … Show All … , Brian B. Oakley, Donovan H. Parks, Courtney J. Robinson, Jason W. Sahl, Blaz Stres, Gerhard G. Thallinger, David J. Van Horn, and Carolyn F. Weber Show FewerDOI: https://doi.org/10.1128/AEM.01541-09Volume 75, Number 231 December 2009ABSTRACTREFERENCESABSTRACTmothur aims to be a comprehensive software package that allows users to use a single piece of software to analyze community sequence data. It builds upon previous tools to provide a flexible and powerful software package for analyzing sequencing data. As a case study, we used mothur to trim, screen, and align sequences; calculate distances; assign sequences to operational taxonomic units; and describe the α and β diversity of eight marine samples previously characterized by pyrosequencing of 16S rRNA gene fragments. This analysis of more than 222,000 sequences was completed in less than 2 h with a laptop computer.REFERENCES1.Antonopoulos, D. A., S. M. Huse, H. G. Morrison, T. M. Schmidt, M. L. Sogin, and V. B. Young.2009. Reproducible community dynamics of the gastrointestinal microbiota following antibiotic perturbation. Infect. Immun.77:2367-2375.CrossrefPubMedGoogle Scholar2.Borneman, J.1999. Culture-independent identification of microorganisms that respond to specified stimuli. Appl. Environ. Microbiol.65:3398-3400.CrossrefPubMedGoogle Scholar3.Cole, J. R., Q. Wang, E. Cardenas, J. Fish, B. Chai, et al.2009. The Ribosomal Database Project: improved alignments and new tools for rRNA analysis. Nucleic Acids Res.37:D141-D145.CrossrefPubMedGoogle Scholar4.DeSantis, T. Z., P. Hugenholtz, N. Larsen, M. Rojas, E. L. Brodie, K. Keller, T. Huber, D. Dalevi, P. Hu, and G. L. Andersen.2006. Greengenes, a chimera-checked 16S rRNA gene database and workbench compatible with ARB. Appl. Environ. Microbiol.72:5069-5072.CrossrefPubMedGoogle Scholar5.DeSantis, T. Z., Jr., P. Hugenholtz, K. Keller, E. L. Brodie, N. Larsen, et al.2006. NAST: a multiple sequence alignment server for comparative analysis of 16S rRNA genes. Nucleic Acids Res.34:W394-W939.PubMedGoogle Scholar6.Felsenstein, J.1989. PHYLIP—Phylogeny Inference Package. Cladistics5:164-166.Google Scholar7.Gamma, E., R. Helm, R. Johnson, and J. M. Vlissides.1995. Design patterns: elements of reusable object-oriented software. Addison-Wesley, Reading, MA.Google Scholar8.Hall, J. R., K. R. Mitchell, O. Jackson-Weaver, A. S. Kooser, B. R. Cron, L. J. Crossey, and C. D. Takacs-Vesbach.2008. Molecular characterization of the diversity and distribution of a thermal spring microbial community by using rRNA and metabolic genes. Appl. Environ. Microbiol.74:4910-4922.CrossrefPubMedGoogle Scholar9.Hartmann, M., and F. Widmer.2006. Community structure analyses are more sensitive to differences in soil bacterial communities than anonymous diversity indices. Appl. Environ. Microbiol.72:7804-7812.CrossrefPubMedGoogle Scholar10.Li, W., and A. Godzik.2006. CD-HIT: a fast program for clustering and comparing large sets of protein or nucleotide sequences. Bioinformatics22:1658-1659.PubMedGoogle Scholar11.Lozupone, C., M. Hamady, and R. Knight.2006. UniFrac—an online tool for comparing microbial community diversity in a phylogenetic context. BMC Bioinformatics7:371.PubMedGoogle Scholar12.Lozupone, C., and R. Knight.2005. UniFrac: a new phylogenetic method for comparing microbial communities. Appl. Environ. Microbiol.71:8228-8235.CrossrefPubMedGoogle Scholar13.Ludwig, W., O. Strunk, R. Westram, L. Richter, H. Meier, et al.2004. ARB: a software environment for sequence data. Nucleic Acids Res.32:1363-1371.PubMedGoogle Scholar14.Maddison, W. P., and M. Slatkin.1991. Null models for the number of evolutionary steps in a character on a phylogenetic tree. Evolution45:1184-1197.PubMedGoogle Scholar15.Martin, A. P.2002. Phylogenetic approaches for describing and comparing the diversity of microbial communities. Appl. Environ. Microbiol.68:3673-3682.CrossrefPubMedGoogle Scholar16.McCaig, A. E., L. A. Glover, and J. I. Prosser.1999. Molecular analysis of bacterial community structure and diversity in unimproved and improved upland grass pastures. Appl. Environ. Microbiol.65:1721-1730.CrossrefPubMedGoogle Scholar17.McConnell, S.2004. Code complete, 2nd ed. Microsoft Press, Redmond, WA.Google Scholar18.Pace, N. R., D. A. Stahl, D. J. Lane, and G. J. Olsen.1985. Analyzing natural microbial populations by rRNA sequences. ASM News51:4-12.Google Scholar19.Pilone, D., and R. Miles.2008. Head first software development. O\'Reilly, Sebastopol, CA.Google Scholar20.Pruesse, E., C. Quast, K. Knittel, B. M. Fuchs, W. Ludwig, et al.2007. SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Res.35:7188-7196.PubMedGoogle Scholar21.Schloss, P. D.2008. Evaluating different approaches that test whether microbial communities have the same structure. ISME J.2:265-275.PubMedGoogle Scholar22.Schloss, P. D., and J. Handelsman.2005. Introducing DOTUR, a computer program for defining operational taxonomic units and estimating species richness. Appl. Environ. Microbiol.71:1501-1506.CrossrefPubMedGoogle Scholar23.Schloss, P. D., and J. Handelsman.2006. Introducing SONS, a tool that compares the membership of microbial communities. Appl. Environ. Microbiol.72:6773-6779.CrossrefPubMedGoogle Scholar24.Schloss, P. D., and J. Handelsman.2006. Introducing TreeClimber, a test to compare microbial community structure. Appl. Environ. Microbiol.72:2379-2384.CrossrefPubMedGoogle Scholar25.Schloss, P. D., B. R. Larget, and J. Handelsman.2004. Integration of microbial ecology and statistics: a test to compare gene libraries. Appl. Environ. Microbiol.70:5485-5492.CrossrefPubMedGoogle Scholar26.Singleton, D. R., M. A. Furlong, S. L. Rathbun, and W. B. Whitman.2001. Quantitative comparisons of 16S rRNA gene sequence libraries from environmental samples. Appl. Environ. Microbiol.67:4374-4376.CrossrefPubMedGoogle Scholar27.Sogin, M. L., H. G. Morrison, J. A. Huber, D. M. Welch, S. M. Huse, et al.2006. Microbial diversity in the deep sea and the underexplored \"rare biosphere.” Proc. Natl. Acad. Sci. USA103:12115-12120.PubMedGoogle Scholar28.Turnbaugh, P. J., M. Hamady, T. Yatsunenko, B. L. Cantarel, A. Duncan, et al.2009. A core gut microbiome in obese and lean twins. Nature457:480-484.PubMedGoogle Scholar29.Turnbaugh, P. J., R. E. Ley, M. Hamady, C. M. Fraser-Liggett, R. Knight, et al.2007. The human microbiome project. Nature449:804-810.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 75 • Number 23 • 1 December 2009Pages: 7537 - 7541HistoryReceived: 30 June 2009Accepted: 26 September 2009Published online: 2 October 2009Copyright© 2009.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsPatrick D. Schloss [email protected]Department of Microbiology, University of Massachusetts, Amherst, MassachusettsDepartment of Microbiology and Immunology, University of Michigan, Ann Arbor, MichiganView all articles by this authorSarah L. WestcottDepartment of Microbiology, University of Massachusetts, Amherst, MassachusettsDepartment of Microbiology and Immunology, University of Michigan, Ann Arbor, MichiganView all articles by this authorThomas RyabinDepartment of Microbiology, University of Massachusetts, Amherst, MassachusettsView all articles by this authorJustine R. HallDepartment of Biology, University of New Mexico, Albuquerque, New MexicoView all articles by this authorMartin HartmannDepartment of Microbiology and Immunology, University of British Columbia, Vancouver, BC, CanadaView all articles by this authorEmily B. HollisterDepartment of Soil and Crop Sciences, Texas A M University, College Station, TexasView all articles by this authorRyan A. LesniewskiDepartment of Soil, Water, and Climate, University of Minnesota, St. Paul, MinnesotaView all articles by this authorBrian B. OakleyDepartment of Biological Sciences, University of Warwick, Coventry, United KingdomView all articles by this authorDonovan H. ParksFaculty of Computer Science, Dalhousie University, Halifax, NS, CanadaView all articles by this authorCourtney J. RobinsonDepartment of Microbiology and Immunology, University of Michigan, Ann Arbor, MichiganView all articles by this authorJason W. SahlEnvironmental Science and Engineering, Colorado School of Mines, Golden, ColoradoView all articles by this authorBlaz StresDepartment of Animal Science, University of Ljubljana, Ljubljana, SloveniaView all articles by this authorGerhard G. ThallingerInstitute for Genomics and Bioinformatics, Graz University of Technology, Graz, AustriaView all articles by this authorDavid J. Van HornDepartment of Microbiology and Immunology, University of Michigan, Ann Arbor, MichiganView all articles by this authorCarolyn F. WeberDepartment of Biological Sciences, Louisiana State University, Baton Rouge, LousianaView all articles by this authorMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleMay 2010Effects of Air Temperature and Relative Humidity on Coronavirus Survival on Surfaces Lisa M. Casanova, Soyoung Jeon, William A. Rutala, David J. Weberand Mark D. SobseyEffects of Air Temperature and Relative Humidity on Coronavirus Survival on SurfacesAuthors: Lisa M. Casanova [email protected], Soyoung Jeon, William A. Rutala, David J. Weber, and Mark D. SobseyDOI: https://doi.org/10.1128/AEM.02291-09Volume 76, Number 91 May 2010ABSTRACTREFERENCESABSTRACTAssessment of the risks posed by severe acute respiratory syndrome (SARS) coronavirus (SARS-CoV) on surfaces requires data on survival of this virus on environmental surfaces and on how survival is affected by environmental variables, such as air temperature (AT) and relative humidity (RH). The use of surrogate viruses has the potential to overcome the challenges of working with SARS-CoV and to increase the available data on coronavirus survival on surfaces. Two potential surrogates were evaluated in this study; transmissible gastroenteritis virus (TGEV) and mouse hepatitis virus (MHV) were used to determine effects of AT and RH on the survival of coronaviruses on stainless steel. At 4°C, infectious virus persisted for as long as 28 days, and the lowest level of inactivation occurred at 20% RH. Inactivation was more rapid at 20°C than at 4°C at all humidity levels; the viruses persisted for 5 to 28 days, and the slowest inactivation occurred at low RH. Both viruses were inactivated more rapidly at 40°C than at 20°C. The relationship between inactivation and RH was not monotonic, and there was greater survival or a greater protective effect at low RH (20%) and high RH (80%) than at moderate RH (50%). There was also evidence of an interaction between AT and RH. The results show that when high numbers of viruses are deposited, TGEV and MHV may survive for days on surfaces at ATs and RHs typical of indoor environments. TGEV and MHV could serve as conservative surrogates for modeling exposure, the risk of transmission, and control measures for pathogenic enveloped viruses, such as SARS-CoV and influenza virus, on health care surfaces.REFERENCES1.Abad, F., R. Pinto, and A. Bosch.1994. Survival of enteric viruses on environmental fomites. Appl. Environ. Microbiol.60:3704-3710.CrossrefPubMedGoogle Scholar2.Reference deleted.Google Scholar3.Bean, B., B. Moore, B. Sterner, L. Peterson, D. Gerding, and H. Balfour, Jr.1982. Survival of influenza viruses on environmental surfaces. J. Infect. Dis.146:47-51.PubMedGoogle Scholar4.Blachere, F., W. Lindsley, T. Pearce, S. Anderson, M. Fisher, R. Khakoo, B. Meade, O. Lander, S. Davis, and R. Thewlis.2009. Measurement of airborne influenza virus in a hospital emergency department. Clin. Infect. Dis.48:438-440.PubMedGoogle Scholar5.Reference deleted.Google Scholar6.Booth, T., B. Kournikakis, N. Bastien, J. Ho, D. Kobasa, L. Stadnyk, Y. Li, M. Spence, S. Paton, and B. Henry.2005. Detection of airborne severe acute respiratory syndrome (SARS) coronavirus and environmental contamination in SARS outbreak units. J. Infect. Dis.191:1472-1477.PubMedGoogle Scholar7.Casanova, L., W. Rutala, D. Weber, and M. Sobsey.2009. Survival of surrogate coronaviruses in water. Water Res.43:1893-1898.PubMedGoogle Scholar8.Chu, C., V. Cheng, I. Hung, K. Chan, B. Tang, T. Tsang, K. Chan, and K. Yuen.2005. Viral load distribution in SARS outbreak. Emerg. Infect. Dis.11:1882-1886.PubMedGoogle Scholar9.Cox, C.1993. Roles of water molecules in bacteria and viruses. Origins Life Evol. Biosph.23:29-36.PubMedGoogle Scholar10.Dowell, S., J. Simmerman, D. Erdman, J. Wu, A. Chaovavanich, M. Javadi, J. Yang, L. Anderson, S. Tong, and M. Ho.2004. Severe acute respiratory syndrome coronavirus on hospital surfaces. Clin. Infect. Dis.39:652-657.PubMedGoogle Scholar11.Harper, G.1961. Airborne micro-organisms: survival tests with four viruses. J. Hyg.59:479-486.PubMedGoogle Scholar12.Harper, G.1963. The influence of environment on the survival of airborne virus particles in the laboratory. Arch. Virol.13:64-71.Google Scholar13.Hemmes, J., K. C. Winkler, and S. M. Kool.1960. Virus survival as a seasonal factor in influenza and poliomyelitis. Nature188:430-431.PubMedGoogle Scholar14.Hung, I. F., V. C. Cheng, A. K. Wu, B. S. Tang, K. H. Chan, C. M. Chu, M. M. Wong, W. T. Hui, L. L. Poon, D. M. Tse, K. S. Chan, P. C. Woo, S. K. Lau, J. S. Peiris, and K. Y. Yuen.2004. Viral loads in clinical specimens and SARS manifestations. Emerg. Infect. Dis.10:1550-1557.PubMedGoogle Scholar15.Ijaz, M., A. Brunner, S. Sattar, R. Nair, and C. Johnson-Lussenburg.1985. Survival characteristics of airborne human coronavirus 229E. J. Gen. Virol.66:2743-2748.PubMedGoogle Scholar16.Jackwood, M. W.2006. The relationship of severe acute respiratory syndrome coronavirus with avian and other coronaviruses. Avian Dis.50:315-320.PubMedGoogle Scholar17.Kim, S., M. Ramakrishnan, P. Raynor, and S. Goyal.2007. Effects of humidity and other factors on the generation and sampling of a coronavirus aerosol. Aerobiologia23:239-248.PubMedGoogle Scholar18.Reference deleted.Google Scholar19.Mbithi, J., V. Springthorpe, and S. Sattar.1991. Effect of relative humidity and air temperature on survival of hepatitis A virus on environmental surfaces. Appl. Environ. Microbiol.57:1394-1399.CrossrefPubMedGoogle Scholar20.McDonald, L., A. Simor, S. IhJen, S. Maloney, M. Ofner, C. KowTong, J. Lando, A. McGeer, L. MinLing, and D. Jernigan.2004. SARS in healthcare facilities, Toronto and Taiwan. Emerg. Infect. Dis.10:777-781.PubMedGoogle Scholar21.Noyce, J., H. Michels, and C. Keevil.2007. Inactivation of influenza A virus on copper versus stainless steel surfaces? Appl. Environ. Microbiol.73:2748-2750.CrossrefPubMedGoogle Scholar22.Peiris, J., S. Lai, L. Poon, Y. Guan, L. Yam, W. Lim, J. Nicholls, W. Yee, W. Yan, and M. Cheung.2003. Coronavirus as a possible cause of severe acute respiratory syndrome. Lancet361:1319-1325.PubMedGoogle Scholar23.Rabenau, H. F., J. Cinatl, B. Morgenstern, G. Bauer, W. Preiser, and H. W. Doerr.2005. Stability and inactivation of SARS coronavirus. Med. Microbiol. Immunol.194:1-6.PubMedGoogle Scholar24.Reference deleted.Google Scholar25.Sattar, S.2004. Microbicides and the environmental control of nosocomial viral infections. J. Hosp. Infect.56(Suppl. 2):S64-S69.PubMedGoogle Scholar26.Schaffer, F., M. Soergel, and D. Straube.1976. Survival of airborne influenza virus: effects of propagating host, relative humidity, and composition of spray fluids. Arch. Virol.51:263-273.PubMedGoogle Scholar27.Shechmeister, I.1950. Studies on the experimental epidemiology of respiratory infections. III. Certain aspects of the behavior of type A influenza virus as an air-borne cloud. J. Infect. Dis.87:128-132.PubMedGoogle Scholar28.Sizun, J., M. Yu, and P. Talbot.2000. Survival of human coronaviruses 229E and OC43 in suspension and after drying on surfaces: a possible source of hospital-acquired infections. J. Hosp. Infect.46:55-60.PubMedGoogle Scholar29.Tennant, B., R. Gaskell, and C. Gaskell.1994. Studies on the survival of canine coronavirus under different environmental conditions. Vet. Microbiol.42:255-259.PubMedGoogle Scholar30.Thompson, S., M. Flury, M. Yates, and W. Jury.1998. Role of the air-water-solid interface in bacteriophage sorption experiments. Appl. Environ. Microbiol.64:304-309.CrossrefPubMedGoogle Scholar31.Thompson, S., and M. Yates.1999. Bacteriophage inactivation at the air-water-solid interface in dynamic batch systems. Appl. Environ. Microbiol.65:1186.CrossrefPubMedGoogle Scholar32.Reference deleted.Google Scholar33.Trouwborst, T., S. Kuyper, J. C. de Jong, and A. Plantinga.1974. Inactivation of some bacterial and animal viruses by exposure to liquid-air interfaces. J. Gen. Virol.24:155-165.PubMedGoogle Scholar34.Wong, S. C., J. K. Chan, K. C. Lee, E. S. Lo, and D. N. Tsang.2005. Development of a quantitative assay for SARS coronavirus and correlation of GAPDH mRNA with SARS coronavirus in clinical specimens. J. Clin. Pathol.58:276-280.PubMedGoogle Scholar35.World Health Organization.2003. First data on stability and resistance of SARS coronavirus compiled by members of WHO laboratory network. World Health Organization, Geneva, Switzerland. http://www.who.int/csr/sars/survival_2003_05_04/en/index.html.Google ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 76 • Number 9 • 1 May 2010Pages: 2712 - 2717HistoryReceived: 23 September 2009Accepted: 26 February 2010Published online: 12 March 2010Copyright© 2010.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsLisa M. Casanova [email protected]Department of Environmental Sciences and Engineering, Gillings School of Global Public Health, University of North Carolina at Chapel Hill, Chapel Hill, North CarolinaView all articles by this authorSoyoung JeonDepartment of Statistics and Operations Research, University of North Carolina at Chapel Hill, Chapel Hill, North CarolinaView all articles by this authorWilliam A. RutalaDepartment of Medicine, University of North Carolina at Chapel Hill, Chapel Hill, North CarolinaView all articles by this authorDavid J. WeberDepartment of Medicine, University of North Carolina at Chapel Hill, Chapel Hill, North CarolinaView all articles by this authorMark D. SobseyDepartment of Environmental Sciences and Engineering, Gillings School of Global Public Health, University of North Carolina at Chapel Hill, Chapel Hill, North CarolinaView all articles by this authorMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleFebruary 2006Comparative Genomics of DNA Fragments from Six Antarctic Marine Planktonic Bacteria Joseph J. Grzymski, Brandon J. Carter, Edward F. DeLong, Robert A. Feldman, Amir Ghadiriand Alison E. MurrayComparative Genomics of DNA Fragments from Six Antarctic Marine Planktonic BacteriaAuthors: Joseph J. Grzymski, Brandon J. Carter, Edward F. DeLong, Robert A. Feldman, Amir Ghadiri, and Alison E. Murray [email protected]DOI: https://doi.org/10.1128/AEM.72.2.1532-1541.2006Volume 72, Number 2February 2006ABSTRACTREFERENCESABSTRACTSixenvironmental fosmid clones from Antarctic coastal waterbacterioplankton were completely sequenced. The genome fragmentsharbored small-subunit rRNA genes that were between 85 and 91% similarto those of their nearest cultivated relatives. The six fragments spanfour phyla, including the Gemmatimonadetes,Proteobacteria (α and γ),Bacteroidetes, and high-G+C gram-positive bacteria.Gene-finding and annotation analyses identified 244 total open readingframes. Amino acid comparisons of 123 and 113 Antarcticbacterial amino acid sequences to mesophilic homologs fromG+C-specific and SwissProt/UniProt databases, respectively,revealed widespread adaptation to the cold. The most significantchanges in these Antarctic bacterial protein sequences included areduction in salt-bridge-forming residues such as arginine, glutamicacid, and aspartic acid, reduced proline contents, and a reduction instabilizing hydrophobic clusters. Stretches of disordered amino acidswere significantly longer in the Antarctic sequences than in themesophilic sequences. These characteristics were not specific to anyone phylum, COG role category, or G+C content and imply thatunderlying genotypic and biochemical adaptations to the cold areinherent to life in the permanently subzero Antarcticwaters.REFERENCES1.Abell,G. C. J., and J. Bowman.2005.Ecological and biogeographic relationships of class Flavobacteria inthe Southern Ocean. FEMS Microbiol. Ecol.51:265-277.PubMedGoogle Scholar2.Bano,N., and J. T. Hollibaugh.2002. Phylogeneticcomposition of bacterioplankton assemblages from the Arctic Ocean.Appl. Environ. Microbiol.68:505-518.CrossrefPubMedGoogle Scholar3.Béjà,O., L. Aravind, E. V. Koonin, M. T. Suzuki, A.Hadd, L. P. Nguyen, S. Jovanovich, C. M. Gates,R. A. Feldman, J. L. Spudich, E. N.Spudich, and E. F. DeLong.2000. Bacterialrhodopsin: evidence for a new type of phototrophy in the sea.Science289:1902-1906.PubMedGoogle Scholar4.Béjà,O., E. V. Koonin, L. Aravind, L. T. Taylor, H.Seitz, J. L. Stein, D. C. Bensen, R. A.Feldman, R. V. Swanson, and E. F. DeLong.2002. Comparative genomic analysis of archaeal genotypicvariants in a single population and in two different oceanic provinces.Appl. Environ. Microbiol.68:335-345.CrossrefPubMedGoogle Scholar5.Béjà,O., M. T. Suzuki, E. V. Koonin, L. Aravind, A.Hadd, L. P. Nguyen, R. Villacorta, M. Amjadi, C. Garrigues,S. B. Jovanovich, R. A. Feldman, and E. F. DeLong.2000. Construction and analysis ofbacterial artificial chromosome libraries from a marine microbialassemblage. Environ. Microbiol.2:516-529.CrossrefPubMedGoogle Scholar6.Courtois,S., C. M. Cappellano, M. Ball, F. X. Francou, P.Normand, G. Helynck, A. Martinez, S. J. Kolvek, J.Hopke, M. S. Osburne, P. R. August, R. Nalin, M.Guerineau, P. Jeannin, P. Simonet, and J. L. Pernodet.2003. Recombinant environmental libraries provide accessto microbial diversity for drug discovery from natural products.Appl. Environ. Microbiol.69:49-55.CrossrefPubMedGoogle Scholar7.Creighton,T. E.1994. The energetic ups and downs ofprotein folding. Nat. Struct. Biol.1:135-138.PubMedGoogle Scholar8.dela Torre, J. R., L. M. Christianson, O. Beja,M. T. Suzuki, D. M. Karl, J. Heidelberg,and E. F. DeLong.2003. Proteorhodopsingenes are distributed among divergent marine bacterial taxa.Proc. Natl. Acad. Sci. USA100:12830-12835.CrossrefPubMedGoogle Scholar9.Delong,E. F., D. G. Franks, and A. L.Alldredge.1993. Phylogenetic diversity Ofaggregate-attached vs free-living marine bacterial assemblages.Limnol. Oceanogr.38:924-934.Google Scholar10.Ducklow,H.2000. Bacterial production and biomass in theoceans, p. 542. In D. L.Kirchman (ed.), Microbial ecology of the oceans.Wiley-Liss, New York,N.Y.Google Scholar11.Dunker,A. K., C. J. Brown, J. D. Lawson,L. M. Iakoucheva, and Z. Obradovic.2002. Intrinsic disorder and protein function.Biochemistry41:6573-6582.PubMedGoogle Scholar12.Feller,G.2003. Molecular adaptations to cold inpsychrophilic enzymes. Cell Mol. Life Sci.60:648-662.PubMedGoogle Scholar13.Feller,G., J. L. Arpigny, E. Narinx, and C. Gerday.1997. Molecular adaptations of enzymes from psychrophilicorganisms. Comp. Biochem. Physiol.118A:495-499.Google Scholar14.Feller,G., and C. Gerday.2003. Psychrophilic enzymes: hottopics in cold adaptation. Nat. Rev. Microbiol.1:200-208.PubMedGoogle Scholar15.Feller,G., and C. Gerday.1997. Psychrophilic enzymes:molecular basis of cold adaptation. Cell. Mol. Life Sci.53:830-841.PubMedGoogle Scholar16.Fuhrman,J. A., and F. Azam.1980. Bacterioplanktonsecondary production estimates for coastal waters of British Columbia,Antarctica, and California. Appl. Environ. Microbiol.39:1085-1095.CrossrefPubMedGoogle Scholar17.Gerday,C., M. Aittaleb, J. L. Arpigny, E. Baise, J. P.Chessa, G. Garsoux, I. Petrescu, and G. Feller.1997.Psychrophilic enzymes: a thermodynamic challenge.Biochim. Biophys. Acta1342:119-131.PubMedGoogle Scholar18.Gianese,G., P. Argo, and S. Pascarella.2001. Structuraladaptation of enzymes to low temperatures. Protein Eng.14:141-148.PubMedGoogle Scholar19.Hallam,S. J., N. Putnam, C. M. Preston, J. C.Detter, D. Rokhsar, P. M. Richardson, and E. F.DeLong.2004. Reverse methanogenesis: testing thehypothesis with environmental genomics. Science305:1457-1462.PubMedGoogle Scholar20.Haney,P. J., J. H. Badger, G. L. Buldak,C. I. Reich, C. R. Woese, and G. J.Olsen.1999. Thermal adaptation analyzed by comparisonof protein sequences from mesophilic and extremely thermophilicMethanococcus species. Proc. Natl. Acad. Sci.USA96:3578-3583.PubMedGoogle Scholar21.Hill,J. E., S. L. Penny, K. G. Crowell,S. H. Goh, and S. M. Hemmingsen.2004. cpnDB: a chaperonin sequence database. GenomeRes.14:1669-1675.PubMedGoogle Scholar22.Karl,D. M.1993. Microbial processes in theSouthern Ocean, p. 634. InE. I. Friedmann (ed.), Antarctic microbiology.Wiley-Liss, New York,N.Y.Google Scholar23.Lopez-Garcia,P., A. Lopez-Lopez, D. Moreira, and F. Rodriguez-Valera.2001. Diversity of free-living prokaryotes froma deep-sea site at the Antarctic polar front. FEMSMicrobiol. Ecol.36:193-202.PubMedGoogle Scholar24.Martin,J. H., R. M. Gordon, and S. E.Fitzwater.1990. Iron in Antarctic waters.Nature345:156-158.Google Scholar25.Methe,B. A., K. E. Nelson, J. W. Deming, B.Momen, E. Melamud, X. Zhang, J. Moult, R. Madupu, W. C.Nelson, R. J. Dodson, L. M. Brinkac, S. C. Daugherty, A. S. Durkin, R. T. DeBoy,J. F. Kolonay, S. A. Sullivan, L. Zhou,T. M. Davidsen, M. Wu, A. L. Huston, M. Lewis, B.Weaver, J. F. Weidman, H. Khouri, T. R. Utterback,T. V. Feldblyum, and C. M. Fraser.2005. The psychrophilic lifestyle as revealed by thegenome sequence of Colwellia psychrerythraea 34H throughgenomic and proteomic analyses. Proc. Natl. Acad. Sci.USA102:10913-10918.PubMedGoogle Scholar26.Moran,M. A., A. Buchan, J. M. Gonzalez, J. F.Heidelberg, W. B. Whitman, R. P. Kiene,J. R. Henriksen, G. M. King, R. Belas, C. Fuqua, L.Brinkac, M.Lewis, S. Johri, B. Weaver, G. Pai,J. A. Eisen, E. Rahe, W. M. Sheldon, W. Y. Ye, T. R. Miller, J. Carlton, D. A.Rasko, I. T. Paulsen, Q. H. Ren, S. C.Daugherty, R. T. Deboy, R. J. Dodson, A. S. Durkin, R. Madupu, W. C. Nelson, S. A. Sullivan,M. J. Rosovitz, D. H. Haft, J. Selengut, and N.Ward.2004. Genome sequence of Silicibacterpomeroyi reveals adaptations to the marine environment.Nature432:910-913.PubMedGoogle Scholar27.Moreira,D., F. Rodriguez-Valera, and P. Lopez-Garcia.2004.Analysis of a genome fragment of a deep-sea uncultivated group IIeuryarchaeote containing 16S rDNA, a spectinomycin-like operon andseveral energy metabolism genes. Environ. Microbiol.6:959-969.PubMedGoogle Scholar28.Murray,A. E., J. T. Hollibaugh, and C. Orrego.1996. Phylogenetic compositions of bacterioplankton fromtwo California estuaries compared by denaturing gradient gelelectrophoresis of 16S rDNA fragments. Appl. Environ.Microbiol.62:2676-2680.CrossrefPubMedGoogle Scholar29.Murray,A. E., C. M. Preston, R. Massana, L. T.Taylor, A. Blakis, K. Wu, and E. F. DeLong.1998. Seasonal and spatial variability of bacterial andarchaeal assemblages in the coastal waters near Anvers Island,Antarctica. Appl. Environ. Microbiol.64:2585-2595.CrossrefPubMedGoogle Scholar30.Muyzer,G., E. C. Dewaal, and A. G. Uitterlinden.1993. Profiling of complex microbial populations bydenaturing gradient gel electrophoresis analysis of polymerasechain reaction-amplified genes coding for 16S rRNA. Appl.Environ. Microbiol.59:695-700.CrossrefPubMedGoogle Scholar31.Rondon,M. R., P. R. August, A. D. Bettermann,S. F. Brady, T. H. Grossman, M. R. Liles,K. A. Loiacono, B. A. Lynch, I. A.MacNeil, C. Minor, C. L. Tiong, M. Gilman, M. S.Osburne, J. Clardy, J. Handelsman, and R. M.Goodman.2000. Cloning the soil metagenome: a strategyfor accessing the genetic and functional diversity of unculturedmicroorganisms. Appl. Environ. Microbiol.66:2541-2547.CrossrefPubMedGoogle Scholar32.Russell,N. J.1990. Cold adaptation ofmicroorganisms. Philos. Trans. R. Soc. Lond. B326:595-611.PubMedGoogle Scholar33.Russell,N. J.2000. Toward a molecular understandingof cold activity of enzymes from psychrophiles.Extremophiles4:83-90.PubMedGoogle Scholar34.Saunders,N. F. W., T. Thomas, P. M. G.Curmi, J. S. Mattick, E. Kuczek, R. Slade, J. Davis,P. D. Franzmann, D. Boone, K. Rusterholtz, R. Feldman, C.Gates, S. Bench, K. Sowers, K. Kadner, A. Aerts, P. Dehal, C. Detter,T. Glavina, S. Lucas, P. Richardson, F. Larimer, L. Hauser, M. Land,and R. Cavicchioli.2003. Mechanisms ofthermal adaptation revealed from the genomes of the Antarctic archaeaMethanogenium frigidum and Methanococcoides burtonii.Genome Res.13:1580-1588.PubMedGoogle Scholar35.Schleper,C., E. F. DeLong, C. M. Preston, R. A.Feldman, K. Y. Wu, and R. V. Swanson.1998. Genomic analysis reveals chromosomal variation innatural populations of the uncultured psychrophilic archaeonCenarchaeum symbiosum. J. Bacteriol.180:5003-5009.CrossrefPubMedGoogle Scholar36.Stein,J. L., T. L. Marsh, K. Y. Wu, H. Shizuya,and E. F. DeLong.1996. Characterization ofuncultivated prokaryotes: isolation and analysis of a 40-kilobase-pairgenome fragment from a planktonic marine archaeon. J.Bacteriol.178:591-599.CrossrefPubMedGoogle Scholar37.Suzuki,M. T., and E. F. DeLong.2002.Marine prokaryote diversity, p.209-234. In J. T.Staley and A. L. Reysenbach (ed.), Biodiversity ofmicrobial life. Wiley-Liss Inc., New York,N.Y.Google Scholar38.Treusch,A. H., A. Kletzin, G. Raddatz, T. Ochsenreiter, A. Quaiser,G. Meurer, S. C. Schuster, and C. Schleper.2004. Characterization of large-insert DNA libraries fromsoil for environmental genomic studies of Archaea. Environ.Microbiol.6:970-980.PubMedGoogle Scholar39.Tyson,G. W., J. Chapman, P. Hugenholtz, E. E.Allen, R. J. Ram, P. M. Richardson, V. V.Solovyev, E. M. Rubin, D. S. Rokhsar, andJ. F. Banfield.2004. Community structureand metabolism through reconstruction of microbial genomes from theenvironment. Nature428:37-43.CrossrefPubMedGoogle Scholar40.Venter,J. C., K. Remington, J. F. Heidelberg, A. L. Halpern, D. Rusch, J. A. Eisen, D. Y. Wu, I.Paulsen, K. E. Nelson, W. Nelson, D. E. Fouts, S.Levy, A. H. Knap, M. W. Lomas, K. Nealson, O.White, J. Peterson, J. Hoffman, R. Parsons, H. Baden-Tillson, C.Pfannkoch, Y. H. Rogers, and H. O. Smith.2004. Environmental genome shotgun sequencing of theSargasso Sea. Science304:66-74.PubMedGoogle Scholar41.Zhang,H., Y. Sekiguchi, S. Hanada, P. Hugenholtz, H. Kim, Y. Kamagata, and K.Nakamura.2003. Gemmatimonas aurantiaca gen.nov., sp. nov., a gram-negative, aerobic, polyphosphate-accumulatingmicro-organism, the first cultured representative of the new bacterialphylum Gemmatimonadetes phyl. nov. Int. J. Syst. Evol.Microbiol.53:1155-1163.PubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 72 • Number 2 • February 2006Pages: 1532 - 1541HistoryReceived: 25 April 2005Accepted: 8 November 2005Copyright© 2006.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsJoseph J. GrzymskiDesert Research Institute, 2215 Raggio Parkway, Reno, Nevada 89512View all articles by this authorBrandon J. CarterDesert Research Institute, 2215 Raggio Parkway, Reno, Nevada 89512View all articles by this authorEdward F. DeLongMassachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, Massachusetts 02139View all articles by this authorRobert A. FeldmanAmersham Biosciences, 928 E. Argues Avenue, Sunnyvale, California 94086Present address: Symbio Corporation, 1455 Adams Dr., Menlo Park, CA 94025.View all articles by this authorAmir GhadiriAmersham Biosciences, 928 E. Argues Avenue, Sunnyvale, California 94086View all articles by this authorAlison E. Murray [email protected]Desert Research Institute, 2215 Raggio Parkway, Reno, Nevada 89512View all articles by this authorNotes†Supplemental material for this article may be found at http://aem.asm.org/.Metrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleDecember 2005UniFrac: a New Phylogenetic Method for Comparing Microbial Communities Catherine Lozuponeand Rob KnightUniFrac: a New Phylogenetic Method for Comparing Microbial CommunitiesAuthors: Catherine Lozupone and Rob Knight [email protected]DOI: https://doi.org/10.1128/AEM.71.12.8228-8235.2005Volume 71, Number 12December 2005ABSTRACTREFERENCESABSTRACTWe introduce here a new method for computing differences between microbial communities based on phylogenetic information. This method, UniFrac, measures the phylogenetic distance between sets of taxa in a phylogenetic tree as the fraction of the branch length of the tree that leads to descendants from either one environment or the other, but not both. UniFrac can be used to determine whether communities are significantly different, to compare many communities simultaneously using clustering and ordination techniques, and to measure the relative contributions of different factors, such as chemistry and geography, to similarities between samples. We demonstrate the utility of UniFrac by applying it to published 16S rRNA gene libraries from cultured isolates and environmental clones of bacteria in marine sediment, water, and ice. Our results reveal that (i) cultured isolates from ice, water, and sediment resemble each other and environmental clone sequences from sea ice, but not environmental clone sequences from sediment and water; (ii) the geographical location does not correlate strongly with bacterial community differences in ice and sediment from the Arctic and Antarctic; and (iii) bacterial communities differ between terrestrially impacted seawater (whether polar or temperate) and warm oligotrophic seawater, whereas those in individual seawater samples are not more similar to each other than to those in sediment or ice samples. These results illustrate that UniFrac provides a new way of characterizing microbial communities, using the wealth of environmental rRNA sequences, and allows quantitative insight into the factors that underlie the distribution of lineages among environments.REFERENCES1.Acinas, S. G., V. Klepac-Ceraj, D. E. Hunt, C. Pharino, I. Ceraj, D. L. Distel, and M. F. Polz.2004. Fine-scale phylogenetic architecture of a complex bacterial community. Nature430:551-554.PubMedGoogle Scholar2.Bano, N., and J. T. Hollibaugh.2002. Phylogenetic composition of bacterioplankton assemblages from the Arctic Ocean. Appl. Environ. Microbiol.68:505-518.CrossrefPubMedGoogle Scholar3.Bowman, J. P., S. A. McCammon, M. V. Brown, D. S. Nichols, and T. A. McMeekin.1997. Diversity and association of psychrophilic bacteria in Antarctic sea ice. Appl. Environ. Microbiol.63:3068-3078.CrossrefPubMedGoogle Scholar4.Bowman, J. P., S. A. McCammon, J. A. Gibson, L. Robertson, and P. D. Nichols.2003. Prokaryotic metabolic activity and community structure in Antarctic continental shelf sediments. Appl. Environ. Microbiol.69:2448-2462.CrossrefPubMedGoogle Scholar5.Bowman, J. P., and R. D. McCuaig.2003. Biodiversity, community structural shifts, and biogeography of prokaryotes within Antarctic continental shelf sediment. Appl. Environ. Microbiol.69:2463-2483.CrossrefPubMedGoogle Scholar6.Brinkmeyer, R., K. Knittel, J. Jurgens, H. Weyland, R. Amann, and E. Helmke.2003. Diversity and structure of bacterial communities in Arctic versus Antarctic pack ice. Appl. Environ. Microbiol.69:6610-6619.CrossrefPubMedGoogle Scholar7.Brown, M. V., and J. P. Bowman.2001. A molecular phylogenetic survey of sea-ice microbial communities (SIMCO). FEMS Microbiol. Ecol.35:267-275.PubMedGoogle Scholar8.Eilers, H., J. Pernthaler, F. O. Glöckner, and R. Amann.2000. Culturability and in situ abundance of pelagic bacteria from the North Sea. Appl. Environ. Microbiol.66:3044-3051.CrossrefPubMedGoogle Scholar9.Felsenstein, J.2004. Inferring phylogenies. Sinauer Associates, Inc., Sunderland, Mass.Google Scholar10.Ferguson, R. L., E. N. Buckley, and A. V. Palumbo.1984. Response of marine bacterioplankton to differential filtration and confinement. Appl. Environ. Microbiol.47:49-55.CrossrefPubMedGoogle Scholar11.Fuhrman, J. A., K. McCallum, and A. A. Davis.1993. Phylogenetic diversity of subsurface marine microbial communities from the Atlantic and Pacific oceans. Appl. Environ. Microbiol.59:1294-1302.CrossrefPubMedGoogle Scholar12.Giovannoni, S. J., and M. Rappé.2000. Evolution, diversity, and molecular ecology of marine prokaryotes, p. 47-84. In D. L. Kirchman (ed.), Microbial ecology of the oceans. John Wiley Sons, Inc., New York, N.Y.Google Scholar13.Glöckner, F. O., E. Zaichikov, N. Belkova, L. Denissova, J. Pernthaler, A. Pernthaler, and R. Amann.2000. Comparative 16S rRNA analysis of lake bacterioplankton reveals globally distributed phylogenetic clusters including an abundant group of actinobacteria. Appl. Environ. Microbiol.66:5053-5065.CrossrefPubMedGoogle Scholar14.Helmke, E., and H. Weyland.1995. Bacteria in sea ice and underlying water of the eastern Weddell Sea in midwinter. Mar. Ecol. Prog. Ser.117:269-287.Google Scholar15.Hugenholtz, P.2002. Exploring prokaryotic diversity in the genomic era. Genome Biol.3:reviews0003.PubMedGoogle Scholar16.Hughes, J. B., J. J. Hellmann, T. H. Ricketts, and B. J. Bohannan.2001. Counting the uncountable: statistical approaches to estimating microbial diversity. Appl. Environ. Microbiol.67:4399-4406.CrossrefPubMedGoogle Scholar17.Hur, I., and J. Chun.2004. A method for comparing multiple bacterial community structures from 16S rDNA clone library sequences. J. Microbiol.42:9-13.PubMedGoogle Scholar18.Jannasch, H. W., and G. E. Jones.1959. Bacterial populations in sea water as determined by different methods of enumeration. Limnol. Oceanogr.4:128-139.Google Scholar19.Junge, K., F. Imhoff, T. Staley, and J. W. Deming.2002. Phylogenetic diversity of numerically important Arctic sea-ice bacteria cultured at subzero temperature. Microb. Ecol.43:315-328.PubMedGoogle Scholar20.Juretschko, S., A. Loy, A. Lehner, and M. Wagner.2002. The microbial community composition of a nitrifying-denitrifying activated sludge from an industrial sewage treatment plant analyzed by the full-cycle rRNA approach. Syst. Appl. Microbiol.25:84-99.PubMedGoogle Scholar21.Kanawaga, T.2003. Bias and artifacts in multitemplate polymerase chain reaction. J. Biosci. Bioeng.96:317-323.PubMedGoogle Scholar22.Kelly, K. M., and A. Y. Chistoserdov.2001. Phylogenetic analysis of the succession of bacterial communities in the Great South Bay (Long Island). FEMS Microbiol. Ecol.35:85-95.PubMedGoogle Scholar23.Krzanowski, W. J.2000. Principles of multivariate analysis. A user\'s perspective. Oxford University Press, Oxford, United Kingdom.Google Scholar24.Ley, R. E., F. Backhed, P. Turnbaugh, C. A. Lozupone, R. D. Knight, and J. I. Gordon.2005. Obesity alters gut microbial ecology. Proc. Natl. Acad. Sci. USA102:11070-11075.PubMedGoogle Scholar25.Li, L., C. Kato, and K. Horikoshi.1999. Microbial diversity in sediments collected from the deepest cold-seep area, the Japan Trench. Mar. Biotechnol. (New York)1:391-400.Google Scholar26.Ludwig, W., O. Strunk, R. Westram, L. Richter, H. Meier, Yadhukumar, A. Buchner, T. Lai, S. Steppi, G. Jobb, W. Förster, I. Brettske, S. Gerber, A. W. Ginhart, O. Gross, S. Grumann, S. Hermann, R. Jost, A. König, T. Liss, R. Lüssmann, M. May, B. Nonhoff, B. Reichel, R. Strehlow, A. Stamatakis, N. Stuckmann, A. Vilbig, M. Lenke, T. Ludwig, A. Bode, and K. H. Schleifer.2004. ARB: a software environment for sequence data. Nucleic Acids Res.32:1363-1371.PubMedGoogle Scholar27.Magurran, A. E.1988. Ecological diversity and its measurement. Princeton University Press, Princeton, N.J.Google Scholar28.Magurran, A. E.2004. Measuring biological diversity. Blackwell, Oxford, United Kingdom.Google Scholar29.Maidak, B. L., J. R. Cole, T. G. Lilburn, C. T. Parker, Jr., P. R. Saxman, R. J. Farris, G. M. Garrity, G. J. Olsen, T. M. Schmidt, and J. M. Tiedje.2001. The RDP-II (Ribosomal Database Project). Nucleic Acids Res.29:173-174.CrossrefPubMedGoogle Scholar30.Martin, A. P.2002. Phylogenetic approaches for describing and comparing the diversity of microbial communities. Appl. Environ. Microbiol.68:3673-3682.CrossrefPubMedGoogle Scholar31.Massana, R., A. E. Murray, C. M. Preston, and E. F. DeLong.1997. Vertical distribution and phylogenetic characterization of marine planktonic archaea in the Santa Barbara Channel. Appl. Environ. Microbiol.63:50-56.CrossrefPubMedGoogle Scholar32.Mullins, T. D., T. B. Britschgi, R. L. Krest, and S. J. Giovannoni.1995. Genetic comparisons reveal the same unknown bacterial lineages in Atlantic and Pacific bacterioplankton communities. Limnol. Oceanogr.40:148-158.Google Scholar33.Nübel, U., F. Garcia-Pichel, M. Kuhl, and G. Muyzer.1999. Quantifying microbial diversity: morphotypes, 16S rRNA genes, and carotenoids of oxygenic phototrophs in microbial mats. Appl. Environ. Microbiol.65:422-430.CrossrefPubMedGoogle Scholar34.Pace, N. R.1997. A molecular view of microbial diversity and the biosphere. Science276:734-740.CrossrefPubMedGoogle Scholar35.Pace, N. R., D. A. Stahl, D. J. Lane, and G. J. Olsen.1986. The analysis of natural microbial populations by ribosomal RNA sequences. Adv. Microb. Ecol.9:1-55.Google Scholar36.Pace, N. R., D. A. Stahl, D. J. Lane, and G. J. Olsen.1985. Analyzing natural microbial populations by rRNA sequences. ASM News51:4-12.Google Scholar37.Rappé, M. S., and S. J. Giovannoni.2003. The uncultured microbial majority. Annu. Rev. Microbiol.57:369-394.PubMedGoogle Scholar38.Ravenschlag, K., K. Sahm, J. Pernthaler, and R. Amann.1999. High bacterial diversity in permanently cold marine sediments. Appl. Environ. Microbiol.65:3982-3989.CrossrefPubMedGoogle Scholar39.Schloss, P. D., B. R. Larget, and J. Handelsman.2004. Integration of microbial ecology and statistics: a test to compare gene libraries. Appl. Environ. Microbiol.70:5485-5492.CrossrefPubMedGoogle Scholar40.Singleton, D. R., M. A. Furlong, S. L. Rathbun, and W. B. Whitman.2001. Quantitative comparisons of 16S rRNA gene sequence libraries from environmental samples. Appl. Environ. Microbiol.67:4374-4376.CrossrefPubMedGoogle Scholar41.Spear, J. R., J. J. Walker, T. M. McCollom, and N. R. Pace.2005. Hydrogen and bioenergetics in the Yellowstone geothermal ecosystem. Proc. Natl. Acad. Sci. USA102:2555-2560.PubMedGoogle Scholar42.Staley, J. T., and J. J. Gosink.1999. Poles apart: biodiversity and biogeography of sea ice bacteria. Annu. Rev. Microbiol.53:189-215.CrossrefPubMedGoogle Scholar43.von Wintzingerode, F., U. B. Göbel, and E. Stackebrandt.1997. Determination of microbial diversity in environmental samples: pitfalls of PCR-based rRNA analysis. FEMS Microbiol. Rev.21:213-229.CrossrefPubMedGoogle Scholar44.Wheeler, P. A., M. Gosselin, E. Sherr, D. Thibault, D. L. Kirchman, R. Benner, and T. E. Whitledge.1996. Active cycling of organic carbon in the central Arctic Ocean. Nature380:697-699.Google Scholar45.Zwart, G., W. D. Hiorns, B. A. Methe, M. P. Van Agterveld, R. Huismans, S. C. Nold, J. P. Zehr, and H. J. Laanbroek.1998. Nearly identical 16S rRNA sequences recovered from lakes in North America and Europe indicate the existence of clades of globally distributed freshwater bacteria. Syst. Appl. Microbiol.21:546-556.PubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 71 • Number 12 • December 2005Pages: 8228 - 8235HistoryReceived: 3 May 2005Accepted: 26 August 2005Copyright© 2005.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsCatherine LozuponeDepartment of Molecular, Cellular, and Developmental Biology, University of Colorado, Boulder, Colorado 80309View all articles by this authorRob Knight [email protected]Department of Chemistry and Biochemistry, University of Colorado, Boulder, Colorado 80309View all articles by this authorMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleSeptember 2011Biodegradation of Polyester Polyurethane by Endophytic Fungi Jonathan R. Russell, Jeffrey Huang, Pria Anand, Kaury Kucera, Amanda G. Sandoval, Kathleen W. Dantzler, DaShawn Hickman, Justin Jee, Farrah M. Kimovec, David Koppstein, Daniel H. Marks, Paul A. Mittermiller, Salvador Joel Núñez, Marina Santiago, Maria A. Townes, Michael Vishnevetsky, Neely E. Williams, Mario Percy Núñez Vargas, Lori-Ann Boulanger, Carol Bascom-Slackand Scott A. StrobelBiodegradation of Polyester Polyurethane by Endophytic FungiAuthors: Jonathan R. Russell, Jeffrey Huang, Pria Anand, Kaury Kucera, Amanda G. Sandoval, Kathleen W. Dantzler, DaShawn Hickman, … Show All … , Justin Jee, Farrah M. Kimovec, David Koppstein, Daniel H. Marks, Paul A. Mittermiller, Salvador Joel Núñez, Marina Santiago, Maria A. Townes, Michael Vishnevetsky, Neely E. Williams, Mario Percy Núñez Vargas, Lori-Ann Boulanger, Carol Bascom-Slack, and Scott A. Strobel [email protected] Show FewerDOI: https://doi.org/10.1128/AEM.00521-11Volume 77, Number 171 September 2011ABSTRACTREFERENCESABSTRACTBioremediation is an important approach to waste reduction that relies on biological processes to break down a variety of pollutants. This is made possible by the vast metabolic diversity of the microbial world. To explore this diversity for the breakdown of plastic, we screened several dozen endophytic fungi for their ability to degrade the synthetic polymer polyester polyurethane (PUR). Several organisms demonstrated the ability to efficiently degrade PUR in both solid and liquid suspensions. Particularly robust activity was observed among several isolates in the genus Pestalotiopsis, although it was not a universal feature of this genus. Two Pestalotiopsis microspora isolates were uniquely able to grow on PUR as the sole carbon source under both aerobic and anaerobic conditions. Molecular characterization of this activity suggests that a serine hydrolase is responsible for degradation of PUR. The broad distribution of activity observed and the unprecedented case of anaerobic growth using PUR as the sole carbon source suggest that endophytes are a promising source of biodiversity from which to screen for metabolic properties useful for bioremediation.REFERENCES1.Akutsu Y., Nakajima-Kambe T., Nomura N., and Nakahara T.. 1998. Purification and properties of a polyester-polyurethane degrading enzyme from Comamonas acidovorans TB-35. Appl. Environ. Microbiol. 64:62–67.CrossrefPubMedGoogle Scholar2.Allen B. A., Hilliard N. P., and Howard G. T.. 1999. Purification and characterization of a soluble polyurethane degrading enzyme from Comamonas acidovorans. Int. Biodeterior. Biodegrad. 43:37–41.Google Scholar3.Bacon C. and White J.. 2000. Microbial endophytes. Marcel Dekker, New York, NY.Google Scholar4.Cosgrove L., McGeechan P. L., Robson G. D., and Handley P. S.. 2007. Fungal communities associated with degradation of polyester polyurethane in soil. Appl. Environ. Microbiol. 73:5817–5824.CrossrefPubMedGoogle Scholar5.Crabbe J. R., Campbell J. R., Thompson L., Walz S. L., and Schultz W. W.. 1994. Biodegradation of a colloidal ester-based polyurethane by soil fungi. Int. Biodeterior. Biodegrad. 33:103–113.Google Scholar6.Darby R. T. and Kaplan A. T.. 1968. Fungal susceptibility of polyurethanes. Appl. Microbiol. 16:900–905.CrossrefPubMedGoogle Scholar7.Filip Z.. 1979. Polyurethane as the sole nutrient source for Aspergillus niger and Cladosporium herbarum. Eur. J. Appl. Microbiol. Biotechnol. 7:277–280.Google Scholar8.Gautam R., Bassi A. S., and Yanful E. K.. 2007. Candida rugosa lipase-catalyzed polyurethane degradation in aqueous medium. Biotechnol. Lett. 29:1081–1086.PubMedGoogle Scholar9.Hanauer D. I., Jacobs-Sera D., and Pedulla M. L.. 2006. Teaching scientific inquiry. Science 314:1880–1881.PubMedGoogle Scholar10.Howard G. T.. 2002. Biodegradation of polyurethane: a review. Int. Biodeterior. Biodegrad. 49:245–252.Google Scholar11.Howard G. T. and Blake R. C.. 1998. Growth of Pseudomonas fluorescens on a polyester-polyurethane and the purification and characterization of a polyurethanase-protease enzyme. Int. Biodeterior. Biodegrad. 42:213–220.Google Scholar12.Howard G. T. and Hilliard N. P.. 1999. Use of Coomassie blue-polyurethane interaction in screening of polyurethanase proteins and polyurethanolytic bacteria. Int. Biodeterior. Biodegrad. 43:23–30.Google Scholar13.Howard G. T., Ruiz C., and Hilliard N. P.. 1999. Growth of Pseudomonas chlororaphis on a polyester-polyurethane and the purification and characterization of a polyurethanase-esterase enzyme. Int. Biodeterior. Biodegrad. 43:7–12.Google Scholar14.Howard G. T., Vicknair J., and Mackie R. I.. 2001. Sensitive plate assay for screening and detection of bacterial polyurethanase activity. Lett. Appl. Microbiol. 32:211–214.PubMedGoogle Scholar15.Kay M. J., Morton L. H. G., and Prince E. L.. 1991. Bacterial degradation of polyester polyurethane. Int. Biodeterior. Biodegrad. 27:205–222.Google Scholar16.Kay M. J., McCabe R. W., and Morton L. H. G.. 1993. Chemical and physical changes occurring in polyester polyurethane during biodegradation. Int. Biodeterior. Biodegrad. 31:209–225.Google Scholar17.Nakajima-Kambe T., Shigeno-Akutsu Y., Nomura N., Onuma F., and Nakahara T.. 1999. Microbial degradation of polyurethane, polyester polyurethanes, and polyether polyurethanes. Appl. Microbiol. Biotechnol. 51:134–140.PubMedGoogle Scholar18.Oceguera-Cervantes A. et al. 2007. Characterization of the polyurethanolytic activity of two Alicycliphilus sp. strains able to degrade polyurethane and N-methylpyrrolidone. Appl. Environ. Microbiol. 73:6214–6223.CrossrefPubMedGoogle Scholar19.Pathirana R. A. and Seal K. J.. 1984. Studies on polyurethane deteriorating fungi. Int. Biodeterior. Biodegrad. 20:163–168.Google Scholar20.Patricelli M. P., Giang D. K., Stamp L. M., and Burbaum J. J.. 2001. Direct visualization of serine hydrolase activities in complex proteomes using fluorescent active site-directed probes. Proteomics 1:1067–1071.PubMedGoogle Scholar21.PlasticsEurope. 1 2008. The compelling facts about plastics, an analysis of plastics production, demand and recovery for 2006 in Europe. PlasticsEurope, Brussels, Belgium. http://www.plasticsrecyclers.eu/docs/press%20release/080123CfaPpdfVersion.pdf.Google Scholar22.Ramirez-Coronel M. A., Viniegra-Gonzalez G., Darvill A., and Augur C.. 2003. A novel tannase from Aspergillus niger with β-glucosidase activity. Microbiology 149:2941–2946.PubMedGoogle Scholar23.Rowe L. and Howard G. T.. 2002. Growth of Bacillus subtilis on polyurethane and the purification and characterization of a polyurethanase-lipase enzyme. Int. Biodeterior. Biodegrad. 50:33–40.Google Scholar24.Shah A. A., Hassan F., Hameed A., and Ahmed S.. 2008. Biological degradation of plastics: a comprehensive review. Biotechnol. Adv. 26:246–265.PubMedGoogle Scholar25.Smith S. A. et al. 2008. Bioactive endophytes warrant intensified exploration and conservation. PLoS One 3:e3052.PubMedGoogle Scholar26.Stern R. V. and Howard G. T.. 2000. The polyester polyurethane gene (pueA) from Pseudomonas clororaphis encodes lipase. FEMS Microbiol. Lett. 185:163–168.PubMedGoogle Scholar27.Strobel G. et al. 1996. Taxol from Pestalotiopsis microspora, an endophytic fungus of Taxus wallachiana. Microbiology 142:435–440.PubMedGoogle Scholar28.Strobel S. A. and Strobel G.. 2007. Plant endophytes as a platform for discovery-based undergraduate science education. Nat. Chem. Biol. 3:356–359.PubMedGoogle Scholar29.Uchida J. Y.. 2004. Pestalotiopsis diseases, p. 27–28. In Elliott M. L., Broschat T. K., Uchida J. Y., and Simone G. W. (ed.), Diseases and disorders of ornamental palms. American Phytopathological Society, St. Paul, MN.Google Scholar30.White T. J., Bruns T., Lee S., and Taylor J. W.. 1990. Amplification and direct sequencing of fungal rRNA genes for phylogenetics, p. 315–322. In Innis M. A., Gelfand D. H., Sninsky J. J., and White T. J. (ed.), PCR protocols: a guide to methods and applications. Academic Press, Inc., New York, NY.Google ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 77 • Number 17 • 1 September 2011Pages: 6076 - 6084HistoryReceived: 7 March 2011Accepted: 21 June 2011Published online: 24 August 2011Copyright© 2011 American Society for Microbiology.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsJonathan R. RussellDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorJeffrey HuangDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorPria AnandDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorKaury KuceraDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorAmanda G. SandovalDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorKathleen W. DantzlerDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorDaShawn HickmanDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorJustin JeeDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorFarrah M. KimovecDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorDavid KoppsteinDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorDaniel H. MarksDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorPaul A. MittermillerDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorSalvador Joel NúñezDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorMarina SantiagoDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorMaria A. TownesDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorMichael VishnevetskyDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorNeely E. WilliamsDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorMario Percy Núñez VargasUniversidad Nacional San Antõnio Abad del Cusco, Peru Escuela Post Grado, Facultad de Biologia, Andes Amazon Guianas Herbario Vargas (CUZ), Cusco, PeruView all articles by this authorLori-Ann BoulangerDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorCarol Bascom-SlackDepartment of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorScott A. Strobel [email protected]Department of Molecular Biophysics and Biochemistry, Yale University, New Haven, Connecticut 06520View all articles by this authorMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleMay 2021Determining Gut Microbial Dysbiosis: a Review of Applied Indexes for Assessment of Intestinal Microbiota Imbalances Shaodong Wei, Martin Iain Bahl, Simon Mark Dahl Baunwall, Christian Lodberg Hvasand Tine Rask LichtDetermining Gut Microbial Dysbiosis: a Review of Applied Indexes for Assessment of Intestinal Microbiota ImbalancesAuthors: Shaodong Wei, Martin Iain Bahl, Simon Mark Dahl Baunwall, Christian Lodberg Hvas, and Tine Rask Licht https://orcid.org/0000-0002-6399-9574 [email protected]DOI: https://doi.org/10.1128/AEM.00395-21Volume 87, Number 1111 May 2021ABSTRACTREFERENCESABSTRACTAssessing \"dysbiosis” in intestinal microbial communities is increasingly considered a routine analysis in microbiota studies, and it has added relevant information to the prediction and characterization of diseases and other adverse conditions. However, dysbiosis is not a well-defined condition. A variety of different dysbiosis indexes have been suggested and applied, but their underlying methodologies, as well as the cohorts and conditions for which they have been developed, differ considerably. To date, no comprehensive overview and comparison of all the different methodologies and applications of such indexes is available. Here, we list all types of dysbiosis indexes identified in the literature, introduce their methodology, group them into categories, and discuss their potential descriptive and clinical applications as well as their limitations. Thus, our focus is not on the implications of dysbiosis for disease but on the methodological approaches available to determine and quantify this condition.REFERENCES1.Santiago M, Eysenbach L, Allegretti J, Aroniadis O, Brandt L, Fischer M, Grinspan A, Kelly C, Morrow C, Rodriguez M, Osman M, Kassam Z, Smith M, Timberlake S. 2019. Microbiome predictors of dysbiosis and VRE decolonization in patients with recurrent C. difficile infections in a multi-center retrospective study. AIMS Microbiol 5:1–18.CrossrefPubMedGoogle Scholar2.Ley RE, Turnbaugh PJ, Klein S, Gordon JI. 2006. Microbial ecology: human gut microbes associated with obesity. Nature 444:1022–1023.CrossrefPubMedGoogle Scholar3.Forslund K, Hildebrand F, Nielsen T, Falony G, Le Chatelier E, Sunagawa S, Prifti E, Vieira-Silva S, Gudmundsdottir V, Pedersen HK, Arumugam M, Kristiansen K, Voigt AY, Vestergaard H, Hercog R, Costea PI, Kultima JR, Li J, Jørgensen T, Levenez F, Dore J, Nielsen HB, Brunak S, Raes J, Hansen T, Wang J, Ehrlich SD, Bork P, Pedersen O, MetaHIT Consortium. 2015. Disentangling type 2 diabetes and metformin treatment signatures in the human gut microbiota. Nature 528:262–266.CrossrefPubMedGoogle Scholar4.Manichanh C, Borruel N, Casellas F, Guarner F. 2012. The gut microbiota in IBD. Nat Rev Gastroenterol Hepatol 9:599–608.CrossrefPubMedGoogle Scholar5.Nakatsu G, Li X, Zhou H, Sheng J, Wong SH, Wu WKK, Ng SC, Tsoi H, Dong Y, Zhang N, He Y, Kang Q, Cao L, Wang K, Zhang J, Liang Q, Yu J, Sung JJY. 2015. Gut mucosal microbiome across stages of colorectal carcinogenesis. Nat Commun 6:30–32.CrossrefGoogle Scholar6.Wang J, Wang Y, Zhang X, Liu J, Zhang Q, Zhao Y, Peng J, Feng Q, Dai J, Sun S, Zhao Y, Zhao L, Zhang Y, Hu Y, Zhang M. 2017. Gut microbial dysbiosis is associated with altered hepatic functions and serum metabolites in chronic hepatitis B patients. Front Microbiol 8:2222.CrossrefPubMedGoogle Scholar7.Jørgensen SF, Trøseid M, Kummen M, Anmarkrud JA, Michelsen AE, Osnes LT, Holm K, Høivik ML, Rashidi A, Dahl CP, Vesterhus M, Halvorsen B, Mollnes TE, Berge RK, Moum B, Lundin KEA, Fevang B, Ueland T, Karlsen TH, Aukrust P, Hov JR. 2016. Altered gut microbiota profile in common variable immunodeficiency associates with levels of lipopolysaccharide and markers of systemic immune activation. Mucosal Immunol 9:1455–1465.CrossrefPubMedGoogle Scholar8.Shen Y, Xu J, Li Z, Huang Y, Yuan Y, Wang J, Zhang M, Hu S, Liang Y. 2018. Analysis of gut microbiota diversity and auxiliary diagnosis as a biomarker in patients with schizophrenia: a cross-sectional study. Schizophr Res 197:470–477.CrossrefPubMedGoogle Scholar9.Huang S, Li R, Zeng X, He T, Zhao H, Chang A, Bo C, Chen J, Yang F, Knight R, Liu J, Davis C, Xu J. 2014. Predictive modeling of gingivitis severity and susceptibility via oral microbiota. ISME J 8:1768–1780.CrossrefPubMedGoogle Scholar10.Kim YJ, Choi YS, Baek KJ, Yoon SH, Park HK, Choi Y. 2016. Mucosal and salivary microbiota associated with recurrent aphthous stomatitis. BMC Microbiol 16:1–10.CrossrefPubMedGoogle Scholar11.Soares RC, Camargo-Penna PH, De Moraes VCS, De Vecchi R, Clavaud C, Breton L, Braz ASK, Paulino LC. 2016. Dysbiotic bacterial and fungal communities not restricted to clinically affected skin sites in dandruff. Front Cell Infect Microbiol 6:157.CrossrefPubMedGoogle Scholar12.Olesen SW, Alm EJ. 2016. Dysbiosis is not an answer. Nat Microbiol 1:16228.CrossrefPubMedGoogle Scholar13.Farup PG, Aasbrenn M, Valeur J. 2018. Separating \"good” from \"bad” faecal dysbiosis–evidence from two cross-sectional studies. BMC Obes 5:30.CrossrefPubMedGoogle Scholar14.Falony G, Joossens M, Vieira-Silva S, Wang J, Darzi Y, Faust K, Kurilshikov A, Bonder MJ, Valles-Colomer M, Vandeputte D, Tito RY, Chaffron S, Rymenans L, Verspecht C, De Sutter L, Lima-Mendez G, D\'hoe K, Jonckheere K, Homola D, Garcia R, Tigchelaar EF, Eeckhaudt L, Fu J, Henckaerts L, Zhernakova A, Wijmenga C, Raes J. 2016. Population-level analysis of gut microbiome variation. Science 352:560–564.CrossrefPubMedGoogle Scholar15.Sarangi AN, Goel A, Aggarwal R. 2019. Methods for studying gut microbiota: a primer for physicians. J Clin Exp Hepatol 9:62–73.CrossrefPubMedGoogle Scholar16.Casén C, Vebø HC, Sekelja M, Hegge FT, Karlsson MK, Ciemniejewska E, Dzankovic S, Frøyland C, Nestestog R, Engstrand L, Munkholm P, Nielsen OH, Rogler G, Simrén M, Öhman L, Vatn MH, Rudi K. 2015. Deviations in human gut microbiota: a novel diagnostic test for determining dysbiosis in patients with IBS or IBD. Aliment Pharmacol Ther 42:71–83.CrossrefPubMedGoogle Scholar17.El-Salhy M, Hausken T, Hatlebakk JG. 2019. Increasing the dose and/or repeating faecal microbiota transplantation (FMT) increases the response in patients with irritable bowel syndrome (IBS). Nutrients 11:1415.CrossrefGoogle Scholar18.El-Salhy M, Hatlebakk JG, Gilja OH, Bråthen Kristoffersen A, Hausken T. 2019. Efficacy of faecal microbiota transplantation for patients with irritable bowel syndrome in a randomised, double-blind, placebo-controlled study. Gut 69:859–867.CrossrefPubMedGoogle Scholar19.Mazzawi T, Lied GA, Sangnes DA, El-Salhy M, Hov JR, Gilja OH, Hatlebakk JG, Hausken T. 2018. The kinetics of gut microbial community composition in patients with irritable bowel syndrome following fecal microbiota transplantation. PLoS One 13:e0194904-17.CrossrefGoogle Scholar20.Bennet SMP, Böhn L, Störsrud S, Liljebo T, Collin L, Lindfors P, Törnblom H, Öhman L, Simrén M. 2018. Multivariate modelling of faecal bacterial profiles of patients with IBS predicts responsiveness to a diet low in FODMAPs. Gut 67:872–881.CrossrefPubMedGoogle Scholar21.Hustoft TN, Hausken T, Ystad SO, Valeur J, Brokstad K, Hatlebakk JG, Lied GA. 2017. Effects of varying dietary content of fermentable short-chain carbohydrates on symptoms, fecal microenvironment, and cytokine profiles in patients with irritable bowel syndrome. Neurogastroenterol Motil 29:e12969.CrossrefGoogle Scholar22.Valeur J, Småstuen MC, Knudsen T, Lied GA, Røseth AG. 2018. Exploring gut microbiota composition as an indicator of clinical response to dietary FODMAP restriction in patients with irritable bowel syndrome. Dig Dis Sci 63:429–436.CrossrefPubMedGoogle Scholar23.Magnusson MK, Strid H, Sapnara M, Lasson A, Bajor A, Ung KA, Öhman L. 2016. Anti-TNF therapy response in patients with ulcerative colitis is associated with colonic antimicrobial peptide expression and microbiota composition. J Crohns Colitis 10:943–952.CrossrefPubMedGoogle Scholar24.Aasbrenn M, Valeur J, Farup PG. 2018. Evaluation of a faecal dysbiosis test for irritable bowel syndrome in subjects with and without obesity. Scand J Clin Lab Invest 78:109–113.CrossrefPubMedGoogle Scholar25.Farup PG, Valeur J. 2020. Changes in faecal short-chain fatty acids after weight-loss interventions in subjects with morbid obesity. Nutrients 12:802–814.CrossrefGoogle Scholar26.Farup PG, Lydersen S, Valeur J. 2019. Are nonnutritive sweeteners obesogenic? Associations between diet, faecal microbiota, and short-chain fatty acids in morbidly obese subjects. J Obes 2019:4608315–4608326.CrossrefPubMedGoogle Scholar27.Mandl T, Marsal J, Olsson P, Ohlsson B, Andréasson K. 2017. Severe intestinal dysbiosis is prevalent in primary Sjögren’s syndrome and is associated with systemic disease activity. Arthritis Res Ther 19:237.CrossrefPubMedGoogle Scholar28.Ferrannini E. 2014. The target of metformin in type 2 diabetes. N Engl J Med 371:1547–1548.CrossrefPubMedGoogle Scholar29.Desilets AR, Dhakal-Karki S, Dunican KC. 2008. Role of metformin for weight management in patients without type 2 diabetes. Ann Pharmacother 42:817–826.CrossrefPubMedGoogle Scholar30.Suez J, Korem T, Zeevi D, Zilberman-Schapira G, Thaiss CA, Maza O, Israeli D, Zmora N, Gilad S, Weinberger A, Kuperman Y, Harmelin A, Kolodkin-Gal I, Shapiro H, Halpern Z, Segal E, Elinav E. 2014. Artificial sweeteners induce glucose intolerance by altering the gut microbiota. Nature 514:181–186.CrossrefPubMedGoogle Scholar31.Thorsen J, Brejnrod A, Mortensen M, Rasmussen MA, Stokholm J, Al-Soud WA, Sørensen S, Bisgaard H, Waage J. 2016. Large-scale benchmarking reveals false discoveries and count transformation sensitivity in 16S rRNA gene amplicon data analysis methods used in microbiome studies. Microbiome 4:62.CrossrefPubMedGoogle Scholar32.Gevers D, Kugathasan S, Denson LA, Vázquez-Baeza Y, Van Treuren W, Ren B, Schwager E, Knights D, Song SJ, Yassour M, Morgan XC, Kostic AD, Luo C, González A, McDonald D, Haberman Y, Walters T, Baker S, Rosh J, Stephens M, Heyman M, Markowitz J, Baldassano R, Griffiths A, Sylvester F, Mack D, Kim S, Crandall W, Hyams J, Huttenhower C, Knight R, Xavier RJ. 2014. The treatment-naive microbiome in new-onset Crohn’s disease. Cell Host Microbe 15:382–392.CrossrefPubMedGoogle Scholar33.Khor B, Gardet A, Xavier RJ. 2011. Genetics and pathogenesis of inflammatory bowel disease. Nature 474:307–317.CrossrefPubMedGoogle Scholar34.Bajaj JS, Heuman DM, Hylemon PB, Sanyal AJ, White MB, Monteith P, Noble NA, Unser AB, Daita K, Fisher AR, Sikaroodi M, Gillevet PM. 2014. Altered profile of human gut microbiome is associated with cirrhosis and its complications. J Hepatol 60:940–947.CrossrefPubMedGoogle Scholar35.Xia GH, You C, Gao XX, Zeng XL, Zhu JJ, Xu KY, Tan CH, Xu RT, Wu QH, Zhou HW, He Y, Yin J. 2019. Stroke dysbiosis index (SDI) in gut microbiome are associated with brain injury and prognosis of stroke. Front Neurol 10:397.CrossrefPubMedGoogle Scholar36.Guo Z, Zhang J, Wang Z, Ang KY, Huang S, Hou Q, Su X, Qiao J, Zheng Y, Wang L, Koh E, Danliang H, Xu J, Lee YK, Zhang H. 2016. Intestinal microbiota distinguish gout patients from healthy humans. Sci Rep 6:20602–20610.CrossrefPubMedGoogle Scholar37.Jeffery IB, O\'Toole PW, Öhman L, Claesson MJ, Deane J, Quigley EMM, Simrén M. 2012. An irritable bowel syndrome subtype defined by species-specific alterations in faecal microbiota. Gut 61:997–1006.CrossrefPubMedGoogle Scholar38.Liu Y, Jin Y, Li J, Zhao L, Li Z, Xu J, Zhao F, Feng J, Chen H, Fang C, Shilpakar R, Wei Y. 2018. Small bowel transit and altered gut microbiota in patients with liver cirrhosis. Front Physiol 9:470.CrossrefPubMedGoogle Scholar39.Mayerhofer CCK, Kummen M, Holm K, Broch K, Awoyemi A, Vestad B, Storm-Larsen C, Seljeflot I, Ueland T, Bohov P, Berge RK, Svardal A, Gullestad L, Yndestad A, Aukrust P, Hov JR, Trøseid M. 2020. Low fibre intake is associated with gut microbiota alterations in chronic heart failure. ESC Hear Fail 7:456–466.CrossrefPubMedGoogle Scholar40.Sze MA, Schloss PD. 2016. Looking for a signal in the noise: revisiting obesity and the microbiome. mBio 7:e01018-16.CrossrefPubMedGoogle Scholar41.Castaner O, Goday A, Park YM, Lee SH, Magkos F, Shiow SATE, Schröder H. 2018. The gut microbiome profile in obesity: a systematic review. Int J Endocrinol 2018:1–9.CrossrefGoogle Scholar42.Sokol H, Leducq V, Aschard H, Pham H, Jegou S, Landman C, Cohen D, Liguori G, Bourrier A, Nion-Larmurier I, Cosnes J, Seksik P, Langella P, Skurnik D, Richard ML, Beaugerie L. 2017. Fungal microbiota dysbiosis in IBD. Gut 66:1039–1048.CrossrefPubMedGoogle Scholar43.Lloyd-Price J, Arze C, Ananthakrishnan AN, Schirmer M, Avila-Pacheco J, Poon TW, Andrews E, Ajami NJ, Bonham KS, Brislawn CJ, Casero D, Courtney H, Gonzalez A, Graeber TG, Hall AB, Lake K, Landers CJ, Mallick H, Plichta DR, Prasad M, Rahnavard G, Sauk J, Shungin D, Vázquez-Baeza Y, White RA, Bishai J, Bullock K, Deik A, Dennis C, Kaplan JL, Khalili H, McIver LJ, Moran CJ, Nguyen L, Pierce KA, Schwager R, Sirota-Madi A, Stevens BW, Tan W, ten Hoeve JJ, Weingart G, Wilson RG, Yajnik V, Braun J, Denson LA, Jansson JK, Knight R, Kugathasan S, McGovern DPB, Petrosino JF, Stappenbeck TS, Winter HS, IBDMDB Investigators. 2019. Multi-omics of the gut microbial ecosystem in inflammatory bowel diseases. Nature 569:655–662.CrossrefPubMedGoogle Scholar44.Truong DT, Franzosa EA, Tickle TL, Scholz M, Weingart G, Pasolli E, Tett A, Huttenhower C, Segata N. 2015. MetaPhlAn2 for enhanced metagenomic taxonomic profiling. Nat Methods 12:902–903.CrossrefPubMedGoogle Scholar45.Franzosa EA, McIver LJ, Rahnavard G, Thompson LR, Schirmer M, Weingart G, Lipson KS, Knight R, Caporaso JG, Segata N, Huttenhower C. 2018. Species-level functional profiling of metagenomes and metatranscriptomes. Nat Methods 15:962–968.CrossrefPubMedGoogle Scholar46.AlShawaqfeh MK, Wajid B, Minamoto Y, Markel M, Lidbury JA, Steiner JM, Serpedin E, Suchodolski JS. 2017. A dysbiosis index to assess microbial changes in fecal samples of dogs with chronic inflammatory enteropathy. FEMS Microbiol Ecol 93:1–8.CrossrefGoogle Scholar47.Whittemore JC, Stokes JE, Price JM, Suchodolski JS. 2019. Effects of a synbiotic on the fecal microbiome and metabolomic profiles of healthy research cats administered clindamycin: a randomized, controlled trial. Gut Microbes 10:521–539.CrossrefPubMedGoogle Scholar48.Blake AB, Guard BC, Honneffer JB, Lidbury JA, Steiner JM, Suchodolski JS. 2019. Altered microbiota, fecal lactate, and fecal bile acids in dogs with gastrointestinal disease. PLoS One 14:e0224454.CrossrefPubMedGoogle Scholar49.Schmidt M, Unterer S, Suchodolski JS, Honneffer JB, Guard BC, Lidbury JA, Steiner JM, Fritz J, Kölle P. 2018. The fecal microbiome and metabolome differs between dogs fed bones and raw food (BARF) diets and dogs fed commercial diets. PLoS One 13:e0201279.CrossrefPubMedGoogle Scholar50.Rossi G, Cerquetella M, Gavazza A, Galosi L, Berardi S, Mangiaterra S, Mari S, Suchodolski JS, Lidbury JA, Steiner JM, Pengo G. 2020. Rapid resolution of large bowel diarrhea after the administration of a combination of a high-fiber diet and a probiotic mixture in 30 dogs. Vet Sci 7:21.CrossrefGoogle Scholar51.Fujishiro MA, Lidbury JA, Pilla R, Steiner JM, Lappin MR, Suchodolski JS. 2020. Evaluation of the effects of anthelmintic administration on the fecal microbiome of healthy dogs with and without subclinical Giardia spp. and Cryptosporidium canis infections. PLoS One 15:e0228145.CrossrefPubMedGoogle Scholar52.Chaitman J, Ziese A-L, Pilla R, Minamoto Y, Blake AB, Guard BC, Isaiah A, Lidbury JA, Steiner JM, Unterer S, Suchodolski JS. 2020. Fecal microbial and metabolic profiles in dogs with acute diarrhea receiving either fecal microbiota transplantation or oral metronidazole. Front Vet Sci 7:192.CrossrefPubMedGoogle Scholar53.Tysnes KR, Angell IL, Fjellanger I, Larsen SD, Søfteland SR, Robertson LJ, Skancke E, Rudi K. 2020. Pre- and post-race intestinal microbiota in long-distance sled dogs and associations with performance. Animals 10:204.CrossrefGoogle Scholar54.Montassier E, Al-Ghalith GA, Hillmann B, Viskocil K, Kabage AJ, McKinlay CE, Sadowsky MJ, Khoruts A, Knights D. 2018. CLOUD: a non-parametric detection test for microbiome outliers. Microbiome 6:137.CrossrefPubMedGoogle Scholar55.Saffouri GB, Shields-Cutler RR, Chen J, Yang Y, Lekatz HR, Hale VL, Cho JM, Battaglioli EJ, Bhattarai Y, Thompson KJ, Kalari KK, Behera G, Berry JC, Peters SA, Patel R, Schuetz AN, Faith JJ, Camilleri M, Sonnenburg JL, Farrugia G, Swann JR, Grover M, Knights D, Kashyap PC. 2019. Small intestinal microbial dysbiosis underlies symptoms associated with functional gastrointestinal disorders. Nat Commun 10:2012.CrossrefPubMedGoogle Scholar56.Chen L, Reeve J, Zhang L, Huang S, Wang X, Chen J. 2018. GMPR: a robust normalization method for zero-inflated count data with application to microbiome sequencing data. PeerJ 6:e4600.CrossrefPubMedGoogle Scholar57.Brahe LK, Le Chatelier E, Prifti E, Pons N, Kennedy S, Blædel T, Håkansson J, Dalsgaard TK, Hansen T, Pedersen O, Astrup A, Ehrlich SD, Larsen LH. 2015. Dietary modulation of the gut microbiota–a randomised controlled trial in obese postmenopausal women. Br J Nutr 114:406–417.CrossrefPubMedGoogle Scholar58.Le Chatelier E, Nielsen T, Qin J, Prifti E, Hildebrand F, Falony G, Almeida M, Arumugam M, Batto J-M, Kennedy S, Leonard P, Li J, Burgdorf K, Grarup N, Jørgensen T, Brandslund I, Nielsen HB, Juncker AS, Bertalan M, Levenez F, Pons N, Rasmussen S, Sunagawa S, Tap J, Tims S, Zoetendal EG, Brunak S, Clément K, Doré J, Kleerebezem M, Kristiansen K, Renault P, Sicheritz-Ponten T, de Vos WM, Zucker J-D, Raes J, Hansen T, Bork P, Wang J, Ehrlich SD, Pedersen O, MetaHIT Consortium. 2013. Richness of human gut microbiome correlates with metabolic markers. Nature 500:541–546.CrossrefPubMedGoogle Scholar59.Qi Y. 2012. Random forest for bioinformatics, p 307–323. In Ensemble machine learning. Springer, New York, NY.CrossrefGoogle Scholar60.David LA, Maurice CF, Carmody RN, Gootenberg DB, Button JE, Wolfe BE, Ling AV, Devlin AS, Varma Y, Fischbach MA, Biddinger SB, Dutton RJ, Turnbaugh PJ. 2014. Diet rapidly and reproducibly alters the human gut microbiome. Nature 505:559–563.CrossrefPubMedGoogle Scholar61.Litvak Y, Byndloss MX, Tsolis RM, Bäumler AJ. 2017. Dysbiotic Proteobacteria expansion: a microbial signature of epithelial dysfunction. Curr Opin Microbiol 39:1–6.CrossrefPubMedGoogle Scholar62.Levy M, Kolodziejczyk AA, Thaiss CA, Elinav E. 2017. Dysbiosis and the immune system. Nat Rev Immunol 17:219–232.CrossrefPubMedGoogle Scholar63.Vich Vila A, Collij V, Sanna S, Sinha T, Imhann F, Bourgonje AR, Mujagic Z, Jonkers DMAE, Masclee AAM, Fu J, Kurilshikov A, Wijmenga C, Zhernakova A, Weersma RK. 2020. Impact of commonly used drugs on the composition and metabolic function of the gut microbiota. Nat Commun 11:1–11.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 87 • Number 11 • 11 May 2021eLocator: e00395-21Editor: Harold L. DrakeUniversity of BayreuthHistoryPublished online: 19 March 2021Copyright© 2021 Wei et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license.PermissionsRequest permissions for this article.Request PermissionsKEYWORDSdysbiosisdysbiosis indexgutimbalanceintestinemicrobiomemicrobiotaContributorsAuthorsShaodong WeiNational Food Institute, Technical University of Denmark, Kgs Lyngby, DenmarkView all articles by this authorMartin Iain BahlNational Food Institute, Technical University of Denmark, Kgs Lyngby, DenmarkView all articles by this authorSimon Mark Dahl BaunwallDepartment of Hepatology and Gastroenterology, Aarhus University Hospital, Aarhus, DenmarkView all articles by this authorChristian Lodberg HvasDepartment of Hepatology and Gastroenterology, Aarhus University Hospital, Aarhus, DenmarkView all articles by this authorTine Rask Licht https://orcid.org/0000-0002-6399-9574 [email protected]National Food Institute, Technical University of Denmark, Kgs Lyngby, DenmarkView all articles by this authorEditorHarold L. DrakeEditorUniversity of BayreuthMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied MicrobiologyArticleNovember 1956Analytical Microbiology John J. GavinAnalytical Microbiology: I. The Test OrganismAuthor: John J. GavinDOI: https://doi.org/10.1128/am.4.6.323-331.1956Volume 4, Number 6November 1956Information ContributorsInformationPublished InApplied MicrobiologyVolume 4 • Number 6 • November 1956Pages: 323 - 331PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorJohn J. GavinFood Research Laboratories, Inc., Long Island City, N. Y.View all articles by this authorMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleJune 2005Inactivation of Enteric Adenovirus and Feline Calicivirus by Chlorine Dioxide Jeanette A. Thurston-Enriquez, Charles N. Haas, Joseph Jacangeloand Charles P. GerbaInactivation of Enteric Adenovirus and Feline Calicivirus by Chlorine DioxideAuthors: Jeanette A. Thurston-Enriquez [email protected], Charles N. Haas, Joseph Jacangelo, and Charles P. GerbaDOI: https://doi.org/10.1128/AEM.71.6.3100-3105.2005Volume 71, Number 6June 2005ABSTRACTREFERENCESABSTRACTChlorine dioxide (ClO2) inactivation experiments were conducted with adenovirus type 40 (AD40) and feline calicivirus (FCV). Experiments were carried out in buffered, disinfectant demand-free water under high- and low-pH and -temperature conditions. Ct values (the concentration of ClO2 multiplied by contact time with the virus) were calculated directly from bench-scale experiments and from application of the efficiency factor Hom (EFH) model. AD40 Ct ranges for 4-log inactivation (Ct99.99%) at 5°C were 0.77 to 1.53 mg/liter × min and 0.80 to 1.59 mg/liter × min for pH 6 and 8, respectively. For 15°C AD40 experiments, 0.49 to 0.74 mg/liter × min and 0.12 mg/liter × min Ct99.99% ranges were observed for pH 6 and 8, respectively. FCV Ct99.99% ranges for 5°C experiments were 20.20 to 30.30 mg/liter × min and 0.68 mg/liter × min for pH 6 and 8, respectively. For 15°C FCV experiments, Ct99.99% ranges were 4.20 to 6.72 and 0.18 mg/liter × min for pH 6 and 8, respectively. Viral inactivation was higher at pH 8 than at pH 6 and at 15°C than at 5°C. Comparison of Ct values and inactivation curves demonstrated that the EFH model described bench-scale experiment data very well. Observed bench-scale Ct99.99% ranges and EFH model Ct99.99% values demonstrated that FCV is more resistant to ClO2 than AD40 for the conditions studied. U.S. Environmental Protection Agency guidance manual Ct99.99% values are higher than Ct99.99% values calculated from bench-scale experiments and from EFH model application.REFERENCES1.Aieta, E. M., and J. D. Berg.1986. A review of chlorine dioxide in drinking water treatment. J. Am. Water Works Assoc.June:62-72.Google Scholar2.Albert, M. J.1986. Enteric adenoviruses. Arch. Virol.88:1-17.PubMedGoogle Scholar3.Alvarez, M. E., and R. T. O\'Brien.1982. Mechanisms of inactivation of poliovirus by chlorine dioxide and iodine. Appl. Environ. Microbiol.44:1064-1071.CrossrefPubMedGoogle Scholar4.American Public Health Association.1998. Standard methods for the examination of water and wastewater, 20th ed. American Public Health Association, Washington, D.C.Google Scholar5.American Water Works Association.1995. Water treatment: principles and practices of water supply operations, 2nd ed. American Water Works Association, Denver, Colo.Google Scholar6.Clancy, J.2000. UV rises to the Cryptosporidium challenge. Water2110:14-16.Google Scholar7.Clarke, I. N., and P. R. Lambden.2000. Organization and expression of Calicivirus genes. J. Infect. Dis.181:S309-S316.PubMedGoogle Scholar8.Cronier, S., P. V. Scarpino, and M. L. Zink.1978. Chlorine dioxide destruction of viruses and bacteria in water, p. 651-658. In R. L. Jolly, H. Gorchev, and D. M. Hamilton (ed.), Water chlorination: environmental impacts and health effects,vol. 2. Ann Arbor Science Publishers, Ann Arbor, Mich.Google Scholar9.Doultree, J. C., J. D. Druce, C. J. Birch, D. S. Bowden, and J. A. Marshall.1999. Inactivation of feline calicivirus, a Norwalk virus surrogate. J. Hosp. Infect.41:51-57.PubMedGoogle Scholar10.Enriquez, C. E., C. J. Hurst, and C. P. Gerba.1995. Survival of the enteric adenovirus-40 and adenovirus-41 in tap, sea, and waste-water. Water Res.29:2548-2553.Google Scholar11.Fankhauser, R. L., J. S. Noel, S. S. Monroe, T. Ando, and R. I. Glass.1998. Molecular epidemiology of \"Norwalk-like viruses” in outbreaks of gastroenteritis in the United States. J. Infect. Dis.178:1571-1578.PubMedGoogle Scholar12.Federal Register.1998. Announcement of the drinking water contaminant candidate list. Fed. Regist.63:10273-10287.Google Scholar13.Foy, H. M., M. K. Cooney, and J. B. Hatlen.1968. Adenovirus type 3 epidemic associated with intermittent chlorination of a swimming pool. Arch. Environ. Health17:795-802.PubMedGoogle Scholar14.Fujioka, R. S., M. A. Dow, and B. S. Yoneyama.1986. Comparative disinfection of indicator bacteria and poliovirus by chlorine dioxide. Water Sci. Technol.18:125-132.Google Scholar15.Gerba, C. P.2000. Disinfection, p. 543-556. In R. M. Maier, I. L. Pepper, and C. P. Gerba (ed.), Environmental microbiology. Academic Press, San Diego, Calif..Google Scholar16.Glass, R. I., J. Noel, T. Ando, R. Fankhauser, G. Belliot, A. Mounts, U. D. Parashar, J. S. Bresee, and S. S. Monroe.2000. The epidemiology of enteric caliciviruses from humans: a reassessment using new diagnostics. J. Infect. Dis.181:S254-S261.PubMedGoogle Scholar17.Gyurek, L. L., and G. R. Finch.1998. Modeling water treatment chemical disinfection kinetics. J. Environ. Eng.124:783-793.Google Scholar18.Gyurek, L. L., L. Hanbin, M. Belosevic, and G. B. Finch.1999. Ozone inactivation kinetics of Cryptosporidium in phosphate buffer. J. Environ. Eng.125:913-924.Google Scholar19.Haas, C. N., and J. Joffe.1994. Disinfection under dynamic conditions: modification of Hom\'s model for decay. Environ. Sci. Technol.28:1367-1369.PubMedGoogle Scholar20.Hafliger, D., P. Hubner, and J. Luthy.2000. Outbreak of viral gastroenteritis due to sewage-contaminated drinking water. Int. J. Food Microbiol.54:123-126.PubMedGoogle Scholar21.Harakeh, S.1987. The behavior of viruses on disinfection by chlorine dioxide and other disinfectants in effluent. FEMS Microbiol. Lett.44:335-341.Google Scholar22.Hoff, J. C.1986. Inactivation of microbial agents by chemical disinfectants. EPA 600-S2-86-067. Office of Water, U.S. Environmental Protection Agency, Washington, D.C.Google Scholar23.Horowitz, M. S.1996. Adenoviruses, p. 2149-2171. In B. N. Fields, D. M. Knipe, and P. M. Howley (ed.), Fields Virology, 3rd ed., vol. 1. Lippincott-Raven, Philadelphia, Pa.Google Scholar24.Hurst, C. J., K. A. McClellan, and W. H. Benton.1988. Comparison of cytopathogenicity, immunofluorescence and in situ DNA hybridization as methods for the detection of adenoviruses. Water Res.22:1547-1552.Google Scholar25.Irving, L. G., and P. A. Smith.1981. One-year survey of enteroviruses, adenoviruses, and reoviruses isolated from effluent at an activated-sludge purification plant. Appl. Environ. Microbiol.41:51-59.CrossrefPubMedGoogle Scholar26.Jiang, X., M. Wang, K. Wang, and M. K. Estes.1993. Sequence and genomic organization of Norwalk virus. Virology195:51-61.PubMedGoogle Scholar27.Junli, H., L. Wang, R. Nenqu, L. Li, S. Fun, and Y. Guanle.1997. Disinfection effect of chlorine dioxide on viruses, algae, and animal planktons in water. Water Res.31:455-460.Google Scholar28.Kapikian, A. Z., M. K. Estes, and R. M. Chanock.1996. Norwalk group of viruses, p. 783-810. In B. N. Fields, D. M. Knipe, and P. M. Howley (ed.), Fields Virology, 3rd ed., vol. 1. Lippincott-Raven, Philadelphia, Pa.Google Scholar29.Kaplan, J. E., R. A. Goodman, L. B. Schonberger, E. C. Lippy, and G. W. Gary.1982. Gastroenteritis due to Norwalk virus: an outbreak associated with a municipal water system. J. Infect. Dis.146:190-197.PubMedGoogle Scholar30.Keswick, B. H., T. K. Satterwhite, P. C. Johnson, H. L. DuPont, S. L. Secor, J. A. Bitsura, G. W. Gary, and J. C. Hoff.1985. Inactivation of Norwalk virus in drinking water by chlorine. Appl. Environ. Microbiol.50:261-264.CrossrefPubMedGoogle Scholar31.Kukkula, M., P. Arstila, M. L. Klossner, L. Maunula, C. H. V. Bonsdorff, and P. Jaatinen.1997. Waterborne outbreak of viral gastroenteritis. Scand. J. Infect. Dis.29:415-418.PubMedGoogle Scholar32.Kukkula, M., L. Maunula, E. Silvennoinen, and C.-H. v. Bonsdorff.1999. Outbreak of viral gastroenteritis due to drinking water contaminated by Norwalk-like viruses. J. Infect. Dis.180:1771-1776.PubMedGoogle Scholar33.Li, H., L. L. Gyurek, G. B. Finch, D. W. Smith, and M. Belosevic.2001. Effect of temperature on ozone inactivation of Cryptosporidium parvum in oxidant demand-free phosphate buffer. J. Environ. Eng.127:456-467.Google Scholar34.Morita, S., A. Namikoshi, T. Hirata, K. Oguma, H. Katayama, and e. al.2002. Efficacy of UV irradiation in inactivating Cryptosporidium parvum oocysts. Appl. Environ. Microbiol.68:5387-5393.CrossrefPubMedGoogle Scholar35.Moss, C. I., and V. P. Olivieri.1985. Disinfecting capabilities of oxychlorine compounds. Appl. Environ. Microbiol.50:1162-1168.CrossrefPubMedGoogle Scholar36.Nuanualsuwan, S., and D. O. Cliver.2003. Capsid functions of inactivated human picornaviruses and feline calicivirus. Appl. Environ. Microbiol.69:350-357.CrossrefPubMedGoogle Scholar37.Papapetropoulou, M., and A. C. Vantarakis.1998. Detection of adenovirus outbreak at a municipal swimming pool by nested PCR amplification. J. Infect.36:101-103.PubMedGoogle Scholar38.Prasad, B. V. V., M. E. Hardy, and M. K. Estes.2000. Structural studies of recombinant Norwalk capsids. J. Infect. Dis.181:S317-S321.PubMedGoogle Scholar39.Rizet, M., N. Dumoutier, and D. Bellahcen.1986. Techniques for the elimination of viral particles. Aqua6:343-348.Google Scholar40.Scarpino, P. V., F. A. Brigano, S. Cronier, and M. L. Zink.1979. Effects of particulates on disinfection of enteroviruses in water by chlorine dioxide. EPA 600-2-79-054. U.S. Environmental Protection Agency, Washington, D.C.Google Scholar41.Shin, G. A., D. Battigelli, and M. D. Sobsey.1998. Reduction of norwalk virus, poliovirus 1, and coliphage MS2 by free chlorine, chlorine dioxide, and ozone disinfection of water. Proceedings of the Water Quality Technology Conference. American Water Works Association, Denver, CO.Google Scholar42.Slomka, M. J., and H. Appleton.1998. Feline calicivirus as a model system for heat inactivation studies of small round structured viruses in shellfish. Epidemiol. Infect.121:401-407.PubMedGoogle Scholar43.Sobsey, M. D.1989. Inactivation of health-related microorganisms in water by disinfection processes. Water Sci. Tech.21:171-195.Google Scholar44.Thurston-Enriquez, J. A., C. N. Haas, J. G. Jacangelo, and C. P. Gerba.2003. Inactivation of feline calicivirus and adenovirus type 40 by UV radiation. Appl. Environ. Microbiol.69:577-582.CrossrefPubMedGoogle Scholar45.Thurston-Enriquez, J. A., C. N. Haas, J. G. Jacangelo, and C. P. Gerba.2003. Chlorine inactivation of adenovirus type 40 and feline calicivirus. Appl. Environ. Microbiol.69:3979-3985.CrossrefPubMedGoogle Scholar46.U.S. Environmental Protection Agency.1989. Guidance manual for compliance with the filtration and disinfection requirements for public water systems using surface water sources. Office of Water, U.S. Environmental Protection Agency, Washington, D.C.Google Scholar47.U.S. Environmental Protection Agency.1999. EPA guidance manual: alternative disinfectants and oxidants. EPA 815-R-99-014. Office of Water, U.S. Environmental Protection Agency, Washington, D.C.Google Scholar48.U.S. Environmental Protection Agency.2001. National primary drinking water standards. EPA 816-F-01-007. Office of Water, U.S. Environmental Protection Agency, Washington, D.C.Google Scholar49.White, G. C.1999. Handbook of chlorination and alternative disinfectants, 4th ed. John Wiley and Sons, Inc., New York, N.Y.Google ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 71 • Number 6 • June 2005Pages: 3100 - 3105HistoryReceived: 12 July 2004Accepted: 20 December 2004Copyright© 2005.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsJeanette A. Thurston-Enriquez [email protected]U.S. Department of Agriculture-Agricultural Research Service, 120 Keim Hall, University of Nebraska East Campus, Lincoln, Nebraska 68583-0934View all articles by this authorCharles N. HaasSchool of Environmental Science, Engineering, and Policy, Drexel University, Philadelphia, Pennsylvania 19104View all articles by this authorJoseph JacangeloMontgomery Watson Harza, Lovettsville, Virginia 20180View all articles by this authorCharles P. GerbaDepartment of Soil, Water, and Environmental Science, University of Arizona, Tucson, Arizona 85721View all articles by this authorMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleApril 2015Multigene Editing in the Escherichia coli Genome via the CRISPR-Cas9 System Yu Jiang, Biao Chen, Chunlan Duan, Bingbing Sun, Junjie Yangand Sheng YangMultigene Editing in the Escherichia coli Genome via the CRISPR-Cas9 SystemAuthors: Yu Jiang, Biao Chen, Chunlan Duan, Bingbing Sun, Junjie Yang, and Sheng YangDOI: https://doi.org/10.1128/AEM.04023-14Volume 81, Number 71 April 2015ABSTRACTREFERENCESABSTRACTAn efficient genome-scale editing tool is required for construction of industrially useful microbes. We describe a targeted, continual multigene editing strategy that was applied to the Escherichia coli genome by using the Streptococcus pyogenes type II CRISPR-Cas9 system to realize a variety of precise genome modifications, including gene deletion and insertion, with a highest efficiency of 100%, which was able to achieve simultaneous multigene editing of up to three targets. The system also demonstrated successful targeted chromosomal deletions in Tatumella citrea, another species of the Enterobacteriaceae, with highest efficiency of 100%.REFERENCES1.Ingram LO, Gomez PF, Lai X, Moniruzzaman M, Wood BE, Yomano LP, York SW. 1998. Metabolic engineering of bacteria for ethanol production. Biotechnol Bioeng 58:204–214.CrossrefPubMedGoogle Scholar2.Atsumi S, Hanai T, Liao JC. 2008. Non-fermentative pathways for synthesis of branched-chain higher alcohols as biofuels. Nature 451:86–89.CrossrefPubMedGoogle Scholar3.Steen EJ, Kang Y, Bokinsky G, Hu Z, Schirmer A, McClure A, Del Cardayre SB, Keasling JD. 2010. Microbial production of fatty-acid-derived fuels and chemicals from plant biomass. Nature 463:559–562.CrossrefPubMedGoogle Scholar4.Leuchtenberger W, Huthmacher K, Drauz K. 2005. Biotechnological production of amino acids and derivatives: current status and prospects. Appl Microbiol Biotechnol 69:1–8.CrossrefPubMedGoogle Scholar5.Bongaerts J, Kramer M, Muller U, Raeven L, Wubbolts M. 2001. Metabolic engineering for microbial production of aromatic amino acids and derived compounds. Metab Eng 3:289–300.CrossrefPubMedGoogle Scholar6.Martin VJ, Pitera DJ, Withers ST, Newman JD, Keasling JD. 2003. Engineering a mevalonate pathway in Escherichia coli for production of terpenoids. Nat Biotechnol 21:796–802.CrossrefPubMedGoogle Scholar7.McDaniel R, Thamchaipenet A, Gustafsson C, Fu H, Betlach M, Ashley G. 1999. Multiple genetic modifications of the erythromycin polyketide synthase to produce a library of novel \"unnatural” natural products. Proc Natl Acad Sci U S A 96:1846–1851.CrossrefPubMedGoogle Scholar8.Yim H, Haselbeck R, Niu W, Pujol-Baxley C, Burgard A, Boldt J, Khandurina J, Trawick JD, Osterhout RE, Stephen R, Estadilla J, Teisan S, Schreyer HB, Andrae S, Yang TH, Lee SY, Burk MJ, Van Dien S. 2011. Metabolic engineering of Escherichia coli for direct production of 1,4-butanediol. Nat Chem Biol 7:445–452.CrossrefPubMedGoogle Scholar9.Nakamura CE, Whited GM. 2003. Metabolic engineering for the microbial production of 1,3-propanediol. Curr Opin Biotechnol 14:454–459.CrossrefPubMedGoogle Scholar10.Esvelt KM, Wang HH. 2013. Genome-scale engineering for systems and synthetic biology. Mol Syst Biol 9:641.CrossrefPubMedGoogle Scholar11.Enyeart PJ, Chirieleison SM, Dao MN, Perutka J, Quandt EM, Yao J, Whitt JT, Keatinge-Clay AT, Lambowitz AM, Ellington AD. 2013. Generalized bacterial genome editing using mobile group II introns and Cre-lox. Mol Syst Biol 9:685.CrossrefPubMedGoogle Scholar12.Yu BJ, Kang KH, Lee JH, Sung BH, Kim MS, Kim SC. 2008. Rapid and efficient construction of markerless deletions in the Escherichia coli genome. Nucleic Acids Res 36:e84.CrossrefPubMedGoogle Scholar13.Datsenko KA, Wanner BL. 2000. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci U S A 97:6640–6645.CrossrefPubMedGoogle Scholar14.Yu D, Ellis HM, Lee EC, Jenkins NA, Copeland NG, Court DL. 2000. An efficient recombination system for chromosome engineering in Escherichia coli. Proc Natl Acad Sci U S A 97:5978–5983.CrossrefPubMedGoogle Scholar15.Zhang Y, Buchholz F, Muyrers JP, Stewart AF. 1998. A new logic for DNA engineering using recombination in Escherichia coli. Nat Genet 20:123–128.CrossrefPubMedGoogle Scholar16.Sharan SK, Thomason LC, Kuznetsov SG, Court DL. 2009. Recombineering: a homologous recombination-based method of genetic engineering. Nat Protoc 4:206–223.CrossrefPubMedGoogle Scholar17.Warner JR, Reeder PJ, Karimpour-Fard A, Woodruff LB, Gill RT. 2010. Rapid profiling of a microbial genome using mixtures of barcoded oligonucleotides. Nat Biotechnol 28:856–862.CrossrefPubMedGoogle Scholar18.Costantino N, Court DL. 2003. Enhanced levels of lambda Red-mediated recombinants in mismatch repair mutants. Proc Natl Acad Sci U S A 100:15748–15753.CrossrefPubMedGoogle Scholar19.Posfai G, Kolisnychenko V, Bereczki Z, Blattner FR. 1999. Markerless gene replacement in Escherichia coli stimulated by a double-strand break in the chromosome. Nucleic Acids Res 27:4409–4415.CrossrefPubMedGoogle Scholar20.Yang J, Sun B, Huang H, Jiang Y, Diao L, Chen B, Xu C, Wang X, Liu J, Jiang W, Yang S. 2014. High-efficiency scarless genetic modification in Escherichia coli using lambda-red recombination and I-SceI cleavage. Appl Environ Microbiol 80:3826–3834.CrossrefPubMedGoogle Scholar21.Karberg M, Guo H, Zhong J, Coon R, Perutka J, Lambowitz AM. 2001. Group II introns as controllable gene targeting vectors for genetic manipulation of bacteria. Nat Biotechnol 19:1162–1167.CrossrefPubMedGoogle Scholar22.Wang HH, Church GM. 2011. Multiplexed genome engineering and genotyping methods applications for synthetic biology and metabolic engineering. Methods Enzymol 498:409–426.CrossrefPubMedGoogle Scholar23.Wang HH, Isaacs FJ, Carr PA, Sun ZZ, Xu G, Forest CR, Church GM. 2009. Programming cells by multiplex genome engineering and accelerated evolution. Nature 460:894–898.CrossrefPubMedGoogle Scholar24.Jiang W, Bikard D, Cox D, Zhang F, Marraffini LA. 2013. RNA-guided editing of bacterial genomes using CRISPR-Cas systems. Nat Biotechnol 31:233–239.CrossrefPubMedGoogle Scholar25.DiCarlo JE, Norville JE, Mali P, Rios X, Aach J, Church GM. 2013. Genome engineering in Saccharomyces cerevisiae using CRISPR-Cas systems. Nucleic Acids Res 41:4336–4343.CrossrefPubMedGoogle Scholar26.Cobb RE, Wang Y, Zhao H. 8 December 2014. High-efficiency multiplex genome editing of Streptomyces species using an engineered CRISPR/Cas system. ACS Synth Biol.CrossrefPubMedGoogle Scholar27.Shan Q, Wang Y, Li J, Zhang Y, Chen K, Liang Z, Zhang K, Liu J, Xi JJ, Qiu JL, Gao C. 2013. Targeted genome modification of crop plants using a CRISPR-Cas system. Nat Biotechnol 31:686–688.CrossrefPubMedGoogle Scholar28.Wang Y, Li Z, Xu J, Zeng B, Ling L, You L, Chen Y, Huang Y, Tan A. 2013. The CRISPR/Cas System mediates efficient genome engineering in Bombyx mori. Cell Res 23:1414–1416.CrossrefPubMedGoogle Scholar29.Yu Z, Ren M, Wang Z, Zhang B, Rong YS, Jiao R, Gao G. 2013. Highly efficient genome modifications mediated by CRISPR/Cas9 in Drosophila. Genetics 195:289–291.CrossrefPubMedGoogle Scholar30.Cong L, Ran FA, Cox D, Lin S, Barretto R, Habib N, Hsu PD, Wu X, Jiang W, Marraffini LA, Zhang F. 2013. Multiplex genome engineering using CRISPR/Cas systems. Science 339:819–823.CrossrefPubMedGoogle Scholar31.Mali P, Yang L, Esvelt KM, Aach J, Guell M, DiCarlo JE, Norville JE, Church GM. 2013. RNA-guided human genome engineering via Cas9. Science 339:823–826.CrossrefPubMedGoogle Scholar32.Zhang Q, Rho M, Tang H, Doak TG, Ye Y. 2013. CRISPR-Cas systems target a diverse collection of invasive mobile genetic elements in human microbiomes. Genome Biol 14:R40.CrossrefPubMedGoogle Scholar33.Deltcheva E, Chylinski K, Sharma CM, Gonzales K, Chao Y, Pirzada ZA, Eckert MR, Vogel J, Charpentier E. 2011. CRISPR RNA maturation by trans-encoded small RNA and host factor RNase III. Nature 471:602–607.CrossrefPubMedGoogle Scholar34.Jinek M, Chylinski K, Fonfara I, Hauer M, Doudna JA, Charpentier E. 2012. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 337:816–821.CrossrefPubMedGoogle Scholar35.Qi LS, Larson MH, Gilbert LA, Doudna JA, Weissman JS, Arkin AP, Lim WA. 2013. Repurposing CRISPR as an RNA-guided platform for sequence-specific control of gene expression. Cell 152:1173–1183.CrossrefPubMedGoogle Scholar36.Pujol CJ, Kado CI. 2000. Genetic and biochemical characterization of the pathway in Pantoea citrea leading to pink disease of pineapple. J Bacteriol 182:2230–2237.CrossrefPubMedGoogle Scholar37.Cha JS, Pujol C, Kado CI. 1997. Identification and characterization of a Pantoea citrea gene encoding glucose dehydrogenase that is essential for causing pink disease of pineapple. Appl Environ Microbiol 63:71–76.CrossrefPubMedGoogle Scholar38.Ochman H, Gerber AS, Hartl DL. 1988. Genetic applications of an inverse polymerase chain reaction. Genetics 120:621–623.PubMedGoogle Scholar39.Shetty RP, Endy D, Knight TF, Jr. 2008. Engineering BioBrick vectors from BioBrick parts. J Biol Eng 2:5.CrossrefPubMedGoogle Scholar40.Guzman LM, Belin D, Carson MJ, Beckwith J. 1995. Tight regulation, modulation, and high-level expression by vectors containing the arabinose PBAD promoter. J Bacteriol 177:4121–4130.CrossrefPubMedGoogle Scholar41.Chayot R, Montagne B, Mazel D, Ricchetti M. 2010. An end-joining repair mechanism in Escherichia coli. Proc Natl Acad Sci U S A 107:2141–2146.CrossrefPubMedGoogle Scholar42.Koboldt DC, Steinberg KM, Larson DE, Wilson RK, Mardis ER. 2013. The next-generation sequencing revolution and its impact on genomics. Cell 155:27–38.CrossrefPubMedGoogle Scholar43.Dodge TC, Valle F, Rashid MH. February 2005. Metabolically engineered bacterial strains having enhanced 2-keto-d-gluconate accumulation. US patent WO2005012486-A2.Google Scholar44.Banta S, Boston M, Jarnagin A, Anderson S. 2002. Mathematical modeling of in vitro enzymatic production of 2-keto-l-gulonic acid using NAD(H) or NADP(H) as cofactors. Metab Eng 4:273–284.CrossrefPubMedGoogle Scholar45.Wilson TE, Topper LM, Palmbos PL. 2003. Non-homologous end-joining: bacteria join the chromosome breakdance. Trends Biochem Sci 28:62–66.CrossrefPubMedGoogle Scholar46.Malyarchuk S, Wright D, Castore R, Klepper E, Weiss B, Doherty AJ, Harrison L. 2007. Expression of Mycobacterium tuberculosis Ku and ligase D in Escherichia coli results in RecA and RecB-independent DNA end-joining at regions of microhomology. DNA Repair (Amst) 6:1413–1424.CrossrefPubMedGoogle Scholar47.Siegele DA, Hu JC. 1997. Gene expression from plasmids containing the araBAD promoter at subsaturating inducer concentrations represents mixed populations. Proc Natl Acad Sci U S A 94:8168–8172.CrossrefPubMedGoogle Scholar48.Cradick TJ, Fine EJ, Antico CJ, Bao G. 2013. CRISPR/Cas9 systems targeting beta-globin and CCR5 genes have substantial off-target activity. Nucleic Acids Res 41:9584–9592.CrossrefPubMedGoogle Scholar49.Fu Y, Foden JA, Khayter C, Maeder ML, Reyon D, Joung JK, Sander JD. 2013. High-frequency off-target mutagenesis induced by CRISPR-Cas nucleases in human cells. Nat Biotechnol 31:822–826.CrossrefPubMedGoogle Scholar50.Ran FA, Hsu PD, Lin CY, Gootenberg JS, Konermann S, Trevino AE, Scott DA, Inoue A, Matoba S, Zhang Y, Zhang F. 2013. Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell 154:1380–1389.CrossrefPubMedGoogle Scholar51.Anders C, Niewoehner O, Duerst A, Jinek M. 2014. Structural basis of PAM-dependent target DNA recognition by the Cas9 endonuclease. Nature 513:569–573.CrossrefPubMedGoogle Scholar52.Martinez E, Bartolome B, de la Cruz F. 1988. pACYC184-derived cloning vectors containing the multiple cloning site and lacZ alpha reporter gene of pUC8/9 and pUC18/19 plasmids. Gene 68:159–162.CrossrefPubMedGoogle Scholar53.Amann E, Ochs B, Abel KJ. 1988. Tightly regulated tac promoter vectors useful for the expression of unfused and fused proteins in Escherichia coli. Gene 69:301–315.CrossrefPubMedGoogle Scholar54.Gust B, Challis GL, Fowler K, Kieser T, Chater KF. 2003. PCR-targeted Streptomyces gene replacement identifies a protein domain needed for biosynthesis of the sesquiterpene soil odor geosmin. Proc Natl Acad Sci U S A 100:1541–1546.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 81 • Number 7 • 1 April 2015Pages: 2506 - 2514Editor: R. M. KellyHistoryReceived: 10 December 2014Accepted: 17 January 2015Published online: 30 January 2015Copyright© 2015 American Society for Microbiology.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsYu JiangKey Laboratory of Synthetic Biology, Institute of Plant Physiology and Ecology, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences, Shanghai, ChinaShanghai Research Center of Industrial Biotechnology, Shanghai, ChinaView all articles by this authorBiao ChenKey Laboratory of Synthetic Biology, Institute of Plant Physiology and Ecology, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences, Shanghai, ChinaShanghai Research Center of Industrial Biotechnology, Shanghai, ChinaView all articles by this authorChunlan DuanKey Laboratory of Synthetic Biology, Institute of Plant Physiology and Ecology, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences, Shanghai, ChinaView all articles by this authorBingbing SunKey Laboratory of Synthetic Biology, Institute of Plant Physiology and Ecology, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences, Shanghai, ChinaShanghai Research Center of Industrial Biotechnology, Shanghai, ChinaView all articles by this authorJunjie YangKey Laboratory of Synthetic Biology, Institute of Plant Physiology and Ecology, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences, Shanghai, ChinaShanghai Research Center of Industrial Biotechnology, Shanghai, ChinaView all articles by this authorSheng YangKey Laboratory of Synthetic Biology, Institute of Plant Physiology and Ecology, Shanghai Institutes for Biological Sciences, Chinese Academy of Sciences, Shanghai, ChinaShanghai Research Center of Industrial Biotechnology, Shanghai, ChinaShanghai Collaborative Innovation Center for Biomanufacturing Technology, Shanghai, ChinaView all articles by this authorEditorR. M. KellyEditorNotesAddress correspondence to Sheng Yang, [email protected].Metrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleNovember 2016Longer Contact Times Increase Cross-Contamination of Enterobacter aerogenes from Surfaces to Food Robyn C. Mirandaand Donald W. SchaffnerLonger Contact Times Increase Cross-Contamination of Enterobacter aerogenes from Surfaces to FoodAuthors: Robyn C. Miranda and Donald W. Schaffner https://orcid.org/0000-0001-9200-0400DOI: https://doi.org/10.1128/AEM.01838-16Volume 82, Number 211 November 2016ABSTRACTREFERENCESABSTRACTBacterial cross-contamination from surfaces to food can contribute to foodborne disease. The cross-contamination rate of Enterobacter aerogenes on household surfaces was evaluated by using scenarios that differed by surface type, food type, contact time ( 1, 5, 30, and 300 s), and inoculum matrix (tryptic soy broth or peptone buffer). The surfaces used were stainless steel, tile, wood, and carpet. The food types were watermelon, bread, bread with butter, and gummy candy. Surfaces (25 cm2) were spot inoculated with 1 ml of inoculum and allowed to dry for 5 h, yielding an approximate concentration of 107 CFU/surface. Foods (with a 16-cm2 contact area) were dropped onto the surfaces from a height of 12.5 cm and left to rest as appropriate. Posttransfer, surfaces and foods were placed in sterile filter bags and homogenized or massaged, diluted, and plated on tryptic soy agar. The transfer rate was quantified as the log percent transfer from the surface to the food. Contact time, food, and surface type all had highly significant effects (P 0.000001) on the log percent transfer of bacteria. The inoculum matrix (tryptic soy broth or peptone buffer) also had a significant effect on transfer (P = 0.013), and most interaction terms were significant. More bacteria transferred to watermelon (∼0.2 to 97%) than to any other food, while the least bacteria transferred to gummy candy (∼0.1 to 62%). Transfer of bacteria to bread (∼0.02 to 94%) was similar to transfer of bacteria to bread with butter (∼0.02 to 82%), and these transfer rates under a given set of conditions were more variable than with watermelon and gummy candy.IMPORTANCE The popular notion of the \"five-second rule” is that food dropped on the floor and left there for 5 s is \"safe” because bacteria need time to transfer. The rule has been explored by a single study in the published literature and on at least two television shows. Results from two academic laboratories have been shared through press releases but remain unpublished. We explored this topic by using four different surfaces (stainless steel, ceramic tile, wood, and carpet), four different foods (watermelon, bread, bread with butter, and gummy candy), four different contact times ( 1, 5, 30, and 300 s), and two bacterial preparation methods. Although we found that longer contact times result in more transfer, we also found that other factors, including the nature of the food and the surface, are of equal or greater importance. Some transfer takes place \"instantaneously,” at times of 1 s, disproving the five-second rule.REFERENCES1.Scallan E, Hoekstra RM, Angulo FJ, Tauxe RV, Widdowson MA, Roy SL, Jones JL, Griffin PM. 2011. Foodborne illness acquired in the United States–major pathogens. Emerg Infect Dis 17:7–15.CrossrefPubMedGoogle Scholar2.Gould LH, Walsh KA, Vieira AR, Herman K, Williams IT, Hall AJ, Cole D. 2013. Surveillance for foodborne disease outbreaks—United States, 1998–2008. MMWR Surveill Summ 62:1–34. http://www.cdc.gov/mmwr/preview/mmwrhtml/ss6202a1.htm.Google Scholar3.Centers for Disease Control and Prevention (CDC). 2013. Surveillance for foodborne disease outbreaks—United States, 2009–2010. Morb Mortal Wkly Rep 62:41–47. http://www.cdc.gov/mmwr/preview/mmwrhtml/mm6203a1.htm.Google Scholar4.Centers for Disease Control and Prevention (CDC). 2013. Surveillance for foodborne disease outbreaks—United States, 2011: annual report. Centers for Disease Control and Prevention, Atlanta, GA. http://www.cdc.gov/foodsafety/pdfs/foodborne-disease-outbreaks-annual-report-2011-508c.pdf.Google Scholar5.Centers for Disease Control and Prevention (CDC). 2014. Surveillance for foodborne disease outbreaks—United States, 2012: annual report. Centers for Disease Control and Prevention, Atlanta, GA. http://www.cdc.gov/foodsafety/pdfs/foodborne-disease-outbreaks-annual-report-2012-508c.pdf.Google Scholar6.Centers for Disease Control and Prevention (CDC). 2015. Surveillance for foodborne disease outbreaks—United States, 2013: annual report. Centers for Disease Control and Prevention, Atlanta, GA. http://www.cdc.gov/foodsafety/pdfs/foodborne-disease-outbreaks-annual-report-2013-508c.pdf.Google Scholar7.Jensen DA, Friedrich LM, Harris LJ, Danyluk MD, Schaffner DW. 2013. Quantifying transfer rates of Salmonella and Escherichia coli O157:H7 between fresh-cut produce and common kitchen surfaces. J Food Prot 76:1530–1538.CrossrefPubMedGoogle Scholar8.Dawson P, Han I, Cox M, Black C, Simmons L. 2007. Residence time and food contact time effects on transfer of Salmonella Typhimurium from tile, wood, and carpet: testing the five-second rule. J Appl Microbiol 102:945–953.PubMedGoogle Scholar9.Wendt C, Dietze B, Dietz E, Rüden H. 1997. Survival of Acinetobacter baumannii on dry surfaces. J Clin Microbiol 35:1394–1397.CrossrefPubMedGoogle Scholar10.Kusumaningrum HD, vanAsselt ED, Beumer RR, Zwietering MH. 2004. A quantitative analysis of cross-contamination of Salmonella and Campylobacter spp. via domestic kitchen surfaces. J Food Prot 67:1892–1903.PubMedGoogle Scholar11.Moore CM, Sheldon BW, Jaykus LA. 2003. Transfer of Salmonella and Campylobacter from stainless steel to romaine lettuce. J Food Prot 66:2231–2236.PubMedGoogle Scholar12.Midelet G, Carpentier B. 2002. Transfer of microorganisms, including Listeria monocytogenes, from various materials to beef. Appl Environ Microbiol 68:4015–4024.CrossrefPubMedGoogle Scholar13.Chen Y, Jackson KM, Chea FP, Schaffner DW. 2001. Quantification and variability analysis of bacterial cross-contamination rates in common food service tasks. J Food Prot 64:72–80.PubMedGoogle Scholar14.Zhao P, Zhao T, Doyle MP, Rubino JR, Meng J. 1998. Development of a model for evaluation of microbial cross-contamination in the kitchen. J Food Prot 61:960–963.PubMedGoogle Scholar15.Lankford MG, Collins S, Youngberg L, Rooney DM, Warren JR, Noskin GA. 2006. Assessment of materials commonly utilized in health care: implications for bacterial survival and transmission. Am J Infect Control 34:258–263.CrossrefPubMedGoogle Scholar16.Rice DH, Hancock DD, Szymanski MH, Scheenstra BC, Cady KM, Besser TE, Chudek PA. 2003. Household contamination with Salmonella enterica. Emerg Infect Dis 9:120–122.CrossrefPubMedGoogle Scholar17.Holah JT, Thorpe RH. 1990. Cleanability in relation to bacterial retention on unused and abraded domestic sink materials. J Appl Bacteriol 69:599–608.CrossrefPubMedGoogle Scholar18.Wilks SA, Michels HT, Keevil CW. 2006. Survival of Listeria monocytogenes Scott A on metal surfaces: implications for cross-contamination. Int J Food Microbiol 111:93–98.CrossrefPubMedGoogle Scholar19.Kuhn PJ. 1983. Doorknobs: a source of nosocomial infection. Diagn Med November-December:62–63. http://www.antimicrobialcopper.org/sites/default/files/upload/Media-library/Files/PDFs/UK/Scientific_literature/kuhn-doorknob.pdf.Google Scholar20.Robine E, Boulangé-Petermann L, Derangère D. 2002. Assessing bactericidal properties of materials: the case of metallic surfaces in contact with air. J Microbiol Methods 49:225–234.CrossrefPubMedGoogle Scholar21.Wilks SA, Michels H, Keevil CW. 2005. The survival of Escherichia coli O157 on a range of metal surfaces. Int J Food Microbiol 105:445–454.CrossrefPubMedGoogle Scholar22.Berto AM. 2007. Ceramic tiles: above and beyond traditional applications. J Eur Ceram Soc 27:1607–1613.CrossrefGoogle Scholar23.Ak NO, Cliver DO, Kaspar CW. 1994. Cutting boards of plastic and wood contaminated experimentally with bacteria. J Food Prot 57:16–22.PubMedGoogle Scholar24.Welker C, Faiola N, Davis S, Maffatore I, Batt CA. 1997. Bacterial retention and cleanability of plastic and wood cutting boards with commercial food service maintenance practices. J Food Prot 60:407–413.PubMedGoogle Scholar25.United States Department of Agriculture (USDA). 2013. Cutting boards and food safety. United States Department of Agriculture, Washington, DC. http://www.fsis.usda.gov/wps/portal/fsis/topics/food-safety-education/get-answers/food-safety-fact-sheets/safe-food-handling/cutting-boards-and-food-safety.Google Scholar26.Yu H. 2007. The effect of chemical finishing on the microbial transfer from carpets to human skin and selected fabrics. Ph.D. dissertation. University of Georgia, Athens, GA.Google Scholar27.. 2003. If you drop it, should you eat it? Scientists weigh in on the 5-second rule. University of Illinois at Urbana-Champaign, Urbana, IL. http://news.aces.illinois.edu/news/if-you-drop-it-should-you-eat-it-scientists-weigh-5-second-rule.Google Scholar28.MythBusters. 2005. 5 second rule with food on floor. Discovery Communications, Silver Spring, MD.http://www.discovery.com/tv-shows/mythbusters/mythbusters-database/5-second-rule-with-food/.Google Scholar29.Garbett J. 10 March 2014. Researchers prove the five-second rule is real. Aston University, Birmingham, United Kingdom. http://www.aston.ac.uk/about/news/releases/2014/march/five-second-food-rule-does-exist/.Google Scholar30.Moran L. 2016. Science explains why the 5-second rule is actually true. The Huffington Post, New York, NY. http://www.huffingtonpost.com/entry/5-second-rule-scienceexplainer_us_56b07205e4b057d7d7c809a8.Google Scholar31.Schaffner DW. 2003. Challenges in cross contamination modelling in home and food service settings. Food Aust 55:583–586.Google Scholar32.Sharma M, Taormina PJ, Beuchat LR. 2003. Habituation of foodborne pathogens exposed to extreme pH conditions: genetic basis and implications in foods and food processing environments. Food Sci Technol Res 9:115–127.CrossrefGoogle Scholar33.Kusumaningrum HD, Riboldi G, Hazeleger WC, Beumer RR. 2003. Survival of foodborne pathogens on stainless steel surfaces and cross-contamination to foods. Int J Food Microbiol 85:227–236.CrossrefPubMedGoogle Scholar34.Donlan RM. 2002. Biofilms: microbial life on surfaces. Emerg Infect Dis 8:881–890.CrossrefPubMedGoogle Scholar35.Ryu JH, Beuchat LR. 2005. Biofilm formation by Escherichia coli O157:H7 on stainless steel: effect of exopolysaccharide and curli production on its resistance to chlorine. Appl Environ Microbiol 71:247–254.CrossrefPubMedGoogle Scholar36.Wachtel MR, Charkowski AO. 2002. Cross-contamination of lettuce with Escherichia coli O157:H7. J Food Prot 65:465–470.PubMedGoogle Scholar37.Mbithi JN, Springthorpe VS, Boulet JR, Sattar SA. 1992. Survival of hepatitis A virus on human hands and its transfer on contact with animate and inanimate surfaces. J Clin Microbiol 30:757–763.CrossrefPubMedGoogle Scholar38.D\'Souza DH, Sair A, Williams K, Papafragkou E, Jean J, Moore C, Jaykus L. 2006. Persistence of caliciviruses on environmental surfaces and their transfer to food. Int J Food Microbiol 108:84–91.CrossrefPubMedGoogle Scholar39.Escudero BI, Rawsthorne H, Gensel C, Jaykus LA. 2012. Persistence and transferability of noroviruses on and between common surfaces and foods. J Food Prot 75:927–935.CrossrefPubMedGoogle Scholar40.Pérez-Rodríguez F, Valero A, Carrasco E, García RM, Zurera G. 2008. Understanding and modelling bacterial transfer to foods: a review. Trends Food Sci Technol 19:131–144.CrossrefGoogle Scholar41.Montville R, Schaffner DW. 2003. Inoculum size influences bacterial cross contamination between surfaces. Appl Environ Microbiol 69:7188–7193.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 82 • Number 21 • 1 November 2016Pages: 6490 - 6496Editor: C. A. ElkinsFDA Center for Food Safety and Applied NutritionHistoryReceived: 16 June 2016Accepted: 16 August 2016Published online: 2 September 2016Copyright© 2016 American Society for Microbiology.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsRobyn C. MirandaRutgers, The State University of New Jersey, New Brunswick, New Jersey, USAView all articles by this authorDonald W. Schaffner https://orcid.org/0000-0001-9200-0400Rutgers, The State University of New Jersey, New Brunswick, New Jersey, USAView all articles by this authorEditorC. A. ElkinsEditorFDA Center for Food Safety and Applied NutritionNotesAddress correspondence to Donald W. Schaffner, [email protected].Metrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleApril 2021Seafloor Incubation Experiment with Deep-Sea Hydrothermal Vent Fluid Reveals Effect of Pressure and Lag Time on Autotrophic Microbial Communities Caroline S. Fortunato, David A. Butterfield, Benjamin Larson, Noah Lawrence-Slavas, Christopher K. Algar, Lisa Zeigler Allen, James F. Holden, Giora Proskurowski, Emily Reddington, Lucy C. Stewart, Begüm D. Topçuoğlu, Joseph J. Vallinoand Julie A. HuberSeafloor Incubation Experiment with Deep-Sea Hydrothermal Vent Fluid Reveals Effect of Pressure and Lag Time on Autotrophic Microbial CommunitiesAuthors: Caroline S. Fortunato, David A. Butterfield, Benjamin Larson https://orcid.org/0000-0002-1007-1684, Noah Lawrence-Slavas, Christopher K. Algar, Lisa Zeigler Allen, James F. Holden https://orcid.org/0000-0002-4343-2378, … Show All … , Giora Proskurowski, Emily Reddington, Lucy C. Stewart, Begüm D. Topçuoğlu, Joseph J. Vallino, and Julie A. Huber https://orcid.org/0000-0002-4790-7633 [email protected] Show FewerDOI: https://doi.org/10.1128/AEM.00078-21Volume 87, Number 913 April 2021ABSTRACTREFERENCESABSTRACTDepressurization and sample processing delays may impact the outcome of shipboard microbial incubations of samples collected from the deep sea. To address this knowledge gap, we developed a remotely operated vehicle (ROV)-powered incubator instrument to carry out and compare results from in situ and shipboard RNA stable isotope probing (RNA-SIP) experiments to identify the key chemolithoautotrophic microbes and metabolisms in diffuse, low-temperature venting fluids from Axial Seamount. All the incubations showed microbial uptake of labeled bicarbonate primarily by thermophilic autotrophic Epsilonbacteraeota that oxidized hydrogen coupled with nitrate reduction. However, the in situ seafloor incubations showed higher abundances of transcripts annotated for aerobic processes, suggesting that oxygen was lost from the hydrothermal fluid samples prior to shipboard analysis. Furthermore, transcripts for thermal stress proteins such as heat shock chaperones and proteases were significantly more abundant in the shipboard incubations, suggesting that depressurization induced thermal stress in the metabolically active microbes in these incubations. Together, the results indicate that while the autotrophic microbial communities in the shipboard and seafloor experiments behaved similarly, there were distinct differences that provide new insight into the activities of natural microbial assemblages under nearly native conditions in the ocean.IMPORTANCE Diverse microbial communities drive biogeochemical cycles in Earth’s ocean, yet studying these organisms and processes is often limited by technological capabilities, especially in the deep ocean. In this study, we used a novel marine microbial incubator instrument capable of in situ experimentation to investigate microbial primary producers at deep-sea hydrothermal vents. We carried out identical stable isotope probing experiments coupled to RNA sequencing both on the seafloor and on the ship to examine thermophilic, microbial autotrophs in venting fluids from an active submarine volcano. Our results indicate that microbial communities were significantly impacted by the effects of depressurization and sample processing delays, with shipboard microbial communities being more stressed than seafloor incubations. Differences in metabolism were also apparent and are likely linked to the chemistry of the fluid at the beginning of the experiment. Microbial experimentation in the natural habitat provides new insights into understanding microbial activities in the ocean.REFERENCES1.Butterfield DA, Roe KK, Lilley MD, Huber JA, Baross JA, Embley RW, Massoth GJ. 2004. Mixing, reaction and microbial activity in the sub-seafloor revealed by temporal and spatial variation in diffuse flow vents at Axial Volcano, p 269–289. In Wilcock WSD, DeLong EF, Kelley DS, Baross JA, Cary SC (ed), The subseafloor biosphere at mid-ocean ridges. American Geophysical Union, Washington, DC.CrossrefGoogle Scholar2.Huber JA, Mark Welch D, Morrison HG, Huse SM, Neal PR, Butterfield DA, Sogin ML. 2007. Microbial population structures in the deep marine biosphere. Science 318:97–100.CrossrefPubMedGoogle Scholar3.Jannasch HW, Mottl MJ. 1985. Geomicrobiology of deep-sea hydrothermal vents. Science 229:717–725.CrossrefPubMedGoogle Scholar4.McNichol J, Stryhanyuk H, Sylva SP, Thomas F, Musat N, Seewald JS, Sievert SM. 2018. Primary productivity below the seafloor at deep-sea hot springs. Proc Natl Acad Sci U S A 115:6756–6761.CrossrefPubMedGoogle Scholar5.Perner M, Bach W, Hentscher M, Koschinsky A, Garbe-Schonberg D, Streit WR, Strauss H. 2009. Short-term microbial and physico-chemical variability in low-temperature hydrothermal fluids near 5°S on the Mid-Atlantic Ridge. Environ Microbiol 11:2526–2541.CrossrefPubMedGoogle Scholar6.Fortunato CS, Huber JA. 2016. Coupled RNA-SIP and metatranscriptomics of active chemolithoautotrophic communities at a deep-sea hydrothermal vent. ISME J 10:1925–1938.CrossrefPubMedGoogle Scholar7.Fortunato CS, Larson B, Butterfield DA, Huber JA. 2018. Spatially distinct, temporally stable microbial populations mediate biogeochemical cycling at and below the seafloor in hydrothermal vent fluids. Environ Microbiol 20:769–784.CrossrefPubMedGoogle Scholar8.Galambos D, Anderson RE, Reveillaud J, Huber JA. 2019. Genome-resolved metagenomics and metatranscriptomics reveal niche differentiation in functionally redundant microbial communities at deep-sea hydrothermal vents. Environ Microbiol 21:4395–4410.CrossrefPubMedGoogle Scholar9.Meier DV, Pjevac P, Bach W, Hourdez S, Girguis PR, Vidoudez C, Amann R, Meyerdierks A. 2017. Niche partitioning of diverse sulfur-oxidizing bacteria at hydrothermal vents. ISME J 11:1545–1558.CrossrefPubMedGoogle Scholar10.Olins HC, Rogers DR, Preston C, Ussler W, Pargett D, Jensen S, Roman B, Birch JM, Scholin CA, Haroon MF, Girguis PR. 2017. Co-registered geochemistry and metatranscriptomics reveal unexpected distributions of microbial activity within a hydrothermal vent field. Front Microbiol 8:1042.CrossrefPubMedGoogle Scholar11.Reveillaud J, Reddington E, McDermott J, Algar C, Meyer JL, Sylva S, Seewald J, German CR, Huber JA. 2016. Subseafloor microbial communities in hydrogen-rich vent fluids from hydrothermal systems along the Mid-Cayman Rise. Environ Microbiol 18:1970–1987.CrossrefPubMedGoogle Scholar12.Trembath-Reichert E, Butterfield DA, Huber JA. 2019. Active subseafloor microbial communities from Mariana back-arc venting fluids share metabolic strategies across different thermal niches and taxa. ISME J 13:2264–2279.CrossrefPubMedGoogle Scholar13.Edgcomb VP, Taylor C, Pachiadaki MG, Honjo S, Engstrom I, Yakimov M. 2016. Comparison of Niskin vs. in situ approaches for analysis of gene expression in deep Mediterranean Sea water samples. Deep Sea Res 2 Top Stud Oceanogr 129:213–222.CrossrefGoogle Scholar14.Ottesen EA. 2016. Probing the living ocean with ecogenomic sensors. Curr Opin Microbiol 31:132–139.CrossrefPubMedGoogle Scholar15.Sievert SM, Vetriani C. 2012. Chemoautotrophy at deep-sea vents: past, present, and future. Oceanography 25:218–233.CrossrefGoogle Scholar16.McNichol J, Sylva SP, Thomas F, Taylor CD, Sievert SM, Seewald JS. 2016. Assessing microbial processes in deep-sea hydrothermal systems by incubation at in situ temperature and pressure. Deep Sea Res 1 Oceanogr Res Pap 115:221–232.CrossrefGoogle Scholar17.McQuillan JS, Robidart JC. 2017. Molecular-biological sensing in aquatic environments: recent developments and emerging capabilities. Curr Opin Biotechnol 45:43–50.CrossrefPubMedGoogle Scholar18.Lippsett L. 2014. Bringing a lab to the seafloor: new device probes deep-sea microbial life. Woods Hole Oceanographic Institute, Woods Hole, MA.Google Scholar19.Taylor C, Howes BL, Doherty KW. 1993. Automated instrumentation for time-series measurement of primary production and nutrient status in production platform-accessible environments. Mar Technol Soc J 27:32–44.Google Scholar20.Taylor CD, Doherty KW. 1990. Submersible incubation device (SID), autonomous instrumentation for the in situ measurement of primary production and other microbial rate processes. Deep Sea Res A 37:343–358.CrossrefGoogle Scholar21.Taylor CD, Molongoski JJ, Lohrenz SE. 1983. Instrumentation for the measurement of phytoplankton production. Limnol Oceanogr 28:781–787.CrossrefGoogle Scholar22.Pachiadaki MG, Rédou V, Beaudoin DJ, Burgaud G, Edgcomb VP. 2016. Fungal and prokaryotic activities in the marine subsurface biosphere at Peru Margin and Canterbury basin inferred from RNA-based analyses and microscopy. Front Microbiol 7:846.CrossrefPubMedGoogle Scholar23.Medina LE, Taylor CD, Pachiadaki MG, Henríquez-Castillo C, Ulloa O, Edgcomb VP. 2017. A review of protist grazing below the photic zone emphasizing studies of oxygen-depleted water columns and recent applications of in situ approaches. Front Mar Sci 4:105.CrossrefGoogle Scholar24.Scholin CA, Birch J, Jensen S, Marin R, III, Massion E, Pargett D, Preston C, Roman B, Ussler W, III. 2017. The quest to develop ecogenomic sensors: a 25-year history of the environmental sample processor (ESP) as a case study. Oceanography 30:100–113.CrossrefGoogle Scholar25.Ussler W, Preston C, Tavormina P, Pargett D, Jensen S, Roman B, Marin R, Shah SR, Girguis PR, Birch JM, Orphan V, Scholin C. 2013. Autonomous application of quantitative PCR in the deep sea: in situ surveys of aerobic methanotrophs using the deep-sea environmental sample processor. Environ Sci Technol 47:9339–9346.CrossrefPubMedGoogle Scholar26.Huber JA, Butterfield DA, Baross JA. 2003. Bacterial diversity in a subseafloor habitat following a deep-sea volcanic eruption. FEMS Microbiol Ecol 43:393–409.CrossrefPubMedGoogle Scholar27.Opatkiewicz AD, Butterfield DA, Baross JA. 2009. Individual hydrothermal vents at Axial Seamount harbor distinct subseafloor microbial communities. FEMS Microbiol Ecol 70:413–424.CrossrefPubMedGoogle Scholar28.Topçuoğlu BD, Stewart LC, Morrison HG, Butterfield DA, Huber JA, Holden JF. 2016. Hydrogen limitation and syntrophic growth among natural assemblages of thermophilic methanogens at deep-sea hydrothermal vents. Front Microbiol 7:1240.CrossrefPubMedGoogle Scholar29.Stewart LC, Algar CK, Fortunato CS, Larson BI, Vallino JJ, Huber JA, Butterfield DA, Holden JF. 2019. Fluid geochemistry, local hydrology, and metabolic activity define methanogen community size and composition in deep-sea hydrothermal vents. ISME J 13:1711–1721.CrossrefPubMedGoogle Scholar30.Seewald JS, Doherty KW, Hammar TR, Liberatore SP. 2002. A new gas-tight isobaric sampler for hydrothermal fluids. Deep Sea Res 1 Oceanogr Res Pap 49:189–196.CrossrefGoogle Scholar31.Fang J, Zhang L, Bazylinski DA. 2010. Deep-sea piezosphere and piezophiles: geomicrobiology and biogeochemistry. Trends Microbiol 18:413–422.CrossrefPubMedGoogle Scholar32.Worthington LV. 1982. The loss of dissolved oxygen in Nansen bottle samples from the deep Atlantic Ocean. Deep Sea Res A 29:1259–1266.CrossrefGoogle Scholar33.Cerqueira T, Barroso C, Froufe H, Egas C, Bettencourt R. 2018. Metagenomic signatures of microbial communities in deep-sea hydrothermal sediments of Azores vent fields. Microb Ecol 76:387–403.CrossrefPubMedGoogle Scholar34.Anantharaman K, Breier JA, Dick GJ. 2016. Metagenomic resolution of microbial functions in deep-sea hydrothermal plumes across the Eastern Lau spreading center. ISME J 10:225–239.CrossrefPubMedGoogle Scholar35.Lesniewski RA, Jain S, Anantharaman K, Schloss PD, Dick GJ. 2012. The metatranscriptome of a deep-sea hydrothermal plume is dominated by water column methanotrophs and lithotrophs. ISME J 6:2257–2268.CrossrefPubMedGoogle Scholar36.Li M, Jain S, Baker BJ, Taylor C, Dick GJ. 2014. Novel hydrocarbon monooxygenase genes in the metatranscriptome of a natural deep-sea hydrocarbon plume. Environ Microbiol 16:60–71.CrossrefPubMedGoogle Scholar37.Stewart FJ, Dalsgaard T, Young CR, Thamdrup B, Revsbech NP, Ulloa O, Canfield DE, DeLong EF. 2012. Experimental incubations elicit profound changes in community transcription in OMZ bacterioplankton. PLoS One 7:e37118.CrossrefGoogle Scholar38.Susin MF, Baldini RL, Gueiros-Filho F, Gomes SL. 2006. GroES/GroEL and DnaK/DnaJ have distinct roles in stress responses and during cell cycle progression in Caulobacter crescentus. J Bacteriol 188:8044–8053.CrossrefPubMedGoogle Scholar39.Alain K, Querellou J, Lesongeur F, Pignet P, Crassous P, Raguénès G, Cueff V, Cambon-Bonavita MA. 2002. Caminibacter hydrogeniphilus gen. nov., sp. nov., a novel thermophilic, hydrogen-oxidizing bacterium isolated from an East Pacific Rise hydrothermal vent. Int J Syst Evol Microbiol 52:1317–1323.CrossrefPubMedGoogle Scholar40.Miroshnichenko ML, L’Haridon S, Schumann P, Spring S, Bonch-Osmolovskaya EA, Jeanthon C, Stackebrandt E. 2004. Caminibacter profundus sp. nov., a novel thermophile of Nautiliales ord. nov. within the class ‘Epsilonproteobacteria’, isolated from a deep-sea hydrothermal vent. Int J Syst Evol Microbiol 54:41–45.CrossrefPubMedGoogle Scholar41.Voordeckers JW, Starovoytov V, Vetriani C. 2005. Caminibacter mediatlanticus sp. nov., a thermophilic, chemolithoautotrophic, nitrate-ammonifying bacterium isolated from a deep-sea hydrothermal vent on the Mid-Atlantic Ridge. Int J Syst Evol Microbiol 55:773–779.CrossrefPubMedGoogle Scholar42.Miroshnichenko ML, Kostrikina NA, L’Haridon S, Jeanthon C, Hippe H, Stackebrandt E, Bonch-Osmolovskaya EA. 2002. Nautilia lithotrophica gen. nov., sp. nov., a thermophilic sulfur-reducing epsilon-proteobacterium isolated from a deep-sea hydrothermal vent. Int J Syst Evol Microbiol 52:1299–1304.CrossrefPubMedGoogle Scholar43.Pérez-Rodríguez I, Ricci J, Voordeckers JW, Starovoytov V, Vetriani C. 2010. Nautilia nitratireducens sp. nov., a thermophilic, anaerobic, chemosynthetic, nitrate-ammonifying bacterium isolated from a deep-sea hydrothermal vent. Int J Syst Evol Microbiol 60:1182–1186.CrossrefPubMedGoogle Scholar44.Alain K, Callac N, Guégan M, Lesongeur F, Crassous P, Cambon-Bonavita M-A, Querellou J, Prieur D. 2009. Nautilia abyssi sp. nov., a thermophilic, chemolithoautotrophic, sulfur-reducing bacterium isolated from an East Pacific Rise hydrothermal vent. Int J Syst Evol Microbiol 59:1310–1315.CrossrefPubMedGoogle Scholar45.Takai K, Nealson KH, Horikoshi K. 2004. Hydrogenimonas thermophila gen. nov., sp. nov., a novel thermophilic, hydrogen-oxidizing chemolithoautotroph within the ε-Proteobacteria, isolated from a black smoker in a Central Indian Ridge hydrothermal field. Int J Syst Evol Microbiol 54:25–32.CrossrefPubMedGoogle Scholar46.Jannasch HW, Wirsen CO, Molyneaux SJ, Langworthy TA. 1992. Comparative physiological studies on hyperthermophilic archaea isolated from deep-sea hot vents with emphasis on Pyrococcus strain GB-D. Appl Environ Microbiol 58:3472–3481.CrossrefPubMedGoogle Scholar47.Holden JF, Baross JA. 1995. Enhanced thermotolerance by hydrostatic pressure in the deep-sea hyperthermophile Pyrococcus strain ES4. FEMS Microbiol Ecol 18:27–33.CrossrefGoogle Scholar48.Pledger RJ, Crump BC, Baross JA. 1994. A barophilic response by two hyperthermophilic, hydrothermal vent Archaea: an upward shift in the optimal temperature and acceleration of growth rate at supra-optimal temperatures by elevated pressure. FEMS Microbiol Ecol 14:233–241.CrossrefGoogle Scholar49.Takahashi S, Tomita J, Nishioka K, Hisada T, Nishijima M. 2014. Development of a prokaryotic universal primer for simultaneous analysis of Bacteria and Archaea using next-generation sequencing. PLoS One 9:e105592.CrossrefGoogle Scholar50.Kucukural A, Yukselen O, Ozata DM, Moore MJ, Garber M. 2019. DEBrowser: interactive differential expression analysis and visualization tool for count data. BMC Genomics 20:6.CrossrefPubMedGoogle Scholar51.Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W, Smyth GK. 2015. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res 43:e47.CrossrefGoogle Scholar52.Darling AE, Jospin G, Lowe E, Matsen FA, IV, Bik HM, Eisen JA. 2014. PhyloSift: phylogenetic analysis of genomes and metagenomes. PeerJ 2:e243.CrossrefGoogle Scholar53.Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. 2015. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res 25:1043–1055.CrossrefPubMedGoogle Scholar54.Langmead B, Salzberg SL. 2012. Fast gapped-read alignment with Bowtie 2. Nat Methods 9:357–359.CrossrefPubMedGoogle Scholar55.Eren AM, Esen ÖC, Quince C, Vineis JH, Morrison HG, Sogin ML, Delmont TO. 2015. Anvi’o: an advanced analysis and visualization platform for ‘omics data. PeerJ 3:e1319.CrossrefPubMedGoogle Scholar56.Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, Marth G, Abecasis G, Durbin R, 1000 Genome Project Data Processing Subgroup. 2009. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25:2078–2079.CrossrefPubMedGoogle Scholar57.R Development Core Team. 2011. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria.Google ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 87 • Number 9 • 13 April 2021eLocator: e00078-21Editor: Shuang-Jiang LiuChinese Academy of SciencesHistoryReceived: 13 January 2021Accepted: 10 February 2021Published online: 19 February 2021Copyright© 2021 Fortunato et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license.PermissionsRequest permissions for this article.Request PermissionsKEYWORDSRNA-SIPautotrophydeep seahydrothermal ventinstrumentationmetagenomicsmetatranscriptomicsContributorsAuthorsCaroline S. FortunatoDepartment of Biology, Widener University, Chester, Pennsylvania, USAView all articles by this authorDavid A. ButterfieldJoint Institute for the Study of Atmosphere and Ocean, University of Washington, Seattle, Washington, USAView all articles by this authorBenjamin Larson https://orcid.org/0000-0002-1007-1684NOAA/PMEL, Seattle, Washington, USAView all articles by this authorNoah Lawrence-SlavasJoint Institute for the Study of Atmosphere and Ocean, University of Washington, Seattle, Washington, USAView all articles by this authorChristopher K. AlgarDepartment of Oceanography, Dalhousie University, Halifax, Nova Scotia, CanadaView all articles by this authorLisa Zeigler AllenMicrobial and Environmental Genomics, J. Craig Venter Institute, La Jolla, California, USAView all articles by this authorJames F. Holden https://orcid.org/0000-0002-4343-2378Department of Microbiology, University of Massachusetts, Amherst, Massachusetts, USAView all articles by this authorGiora ProskurowskiMarqMetrix, Inc., Seattle, Washington, USAView all articles by this authorEmily ReddingtonGreat Pond Foundation, Edgartown, Massachusetts, USAView all articles by this authorLucy C. StewartDepartment of Microbiology, University of Massachusetts, Amherst, Massachusetts, USAPresent address: Lucy C. Stewart, Toha Science, Wellington, New Zealand; Begüm D. Topçuoğlu, Merck Exploratory Science Center, Cambridge, Massachusetts, USA.View all articles by this authorBegüm D. TopçuoğluDepartment of Microbiology, University of Massachusetts, Amherst, Massachusetts, USAPresent address: Lucy C. Stewart, Toha Science, Wellington, New Zealand; Begüm D. Topçuoğlu, Merck Exploratory Science Center, Cambridge, Massachusetts, USA.View all articles by this authorJoseph J. VallinoEcosystems Center, Marine Biological Laboratory, Woods Hole, Massachusetts, USAView all articles by this authorJulie A. Huber https://orcid.org/0000-0002-4790-7633 [email protected]Marine Chemistry and Geochemistry, Woods Hole Oceanographic Institution, Woods Hole, Massachusetts, USAView all articles by this authorEditorShuang-Jiang LiuEditorChinese Academy of SciencesMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleSeptember 2013Development of a Dual-Index Sequencing Strategy and Curation Pipeline for Analyzing Amplicon Sequence Data on the MiSeq Illumina Sequencing Platform James J. Kozich, Sarah L. Westcott, Nielson T. Baxter, Sarah K. Highlanderand Patrick D. SchlossDevelopment of a Dual-Index Sequencing Strategy and Curation Pipeline for Analyzing Amplicon Sequence Data on the MiSeq Illumina Sequencing PlatformAuthors: James J. Kozich, Sarah L. Westcott, Nielson T. Baxter, Sarah K. Highlander, and Patrick D. SchlossDOI: https://doi.org/10.1128/AEM.01043-13Volume 79, Number 171 September 2013ABSTRACTREFERENCESABSTRACTRapid advances in sequencing technology have changed the experimental landscape of microbial ecology. In the last 10 years, the field has moved from sequencing hundreds of 16S rRNA gene fragments per study using clone libraries to the sequencing of millions of fragments per study using next-generation sequencing technologies from 454 and Illumina. As these technologies advance, it is critical to assess the strengths, weaknesses, and overall suitability of these platforms for the interrogation of microbial communities. Here, we present an improved method for sequencing variable regions within the 16S rRNA gene using Illumina\'s MiSeq platform, which is currently capable of producing paired 250-nucleotide reads. We evaluated three overlapping regions of the 16S rRNA gene that vary in length (i.e., V34, V4, and V45) by resequencing a mock community and natural samples from human feces, mouse feces, and soil. By titrating the concentration of 16S rRNA gene amplicons applied to the flow cell and using a quality score-based approach to correct discrepancies between reads used to construct contigs, we were able to reduce error rates by as much as two orders of magnitude. Finally, we reprocessed samples from a previous study to demonstrate that large numbers of samples could be multiplexed and sequenced in parallel with shotgun metagenomes. These analyses demonstrate that our approach can provide data that are at least as good as that generated by the 454 platform while providing considerably higher sequencing coverage for a fraction of the cost.REFERENCES1.Hugenholtz P, Pitulle C, Hershberger KL, and Pace NR. 1998. Novel division level bacterial diversity in a Yellowstone hot spring. J. Bacteriol. 180:366–376.CrossrefPubMedGoogle Scholar2.The Human Microbiome Consortium. 2012. Structure, function and diversity of the healthy human microbiome. Nature 486:207–214.PubMedGoogle Scholar3.Sogin ML, Morrison HG, Huber JA, Welch DM, Huse SM, Neal PR, Arrieta JM, and Herndl GJ. 2006. Microbial diversity in the deep sea and the underexplored \"rare biosphere.”. Proc. Natl. Acad. Sci. U. S. A. 103:12115–12120.CrossrefPubMedGoogle Scholar4.Junemann S, Prior K, Szczepanowski R, Harks I, Ehmke B, Goesmann A, Stoye J, and Harmsen D. 2012. Bacterial community shift in treated periodontitis patients revealed by Ion Torrent 16S rRNA gene amplicon sequencing. PLoS One 7:e41606.CrossrefPubMedGoogle Scholar5.Fichot EB and Norman RS. 2013. Microbial phylogenetic profiling with the Pacific Biosciences sequencing platform. Microbiome 1:10.PubMedGoogle Scholar6.Gloor GB, Hummelen R, Macklaim JM, Dickson RJ, Fernandes AD, MacPhee R, and Reid G. 2010. Microbiome profiling by Illumina sequencing of combinatorial sequence-tagged PCR products. PLoS One 5:e15406.CrossrefPubMedGoogle Scholar7.Huse SM, Welch DM, Morrison HG, and Sogin ML. 2010. Ironing out the wrinkles in the rare biosphere through improved OTU clustering. Environ. Microbiol. 12:1889–1898.CrossrefPubMedGoogle Scholar8.Kunin V, Engelbrektson A, Ochman H, and Hugenholtz P. 2010. Wrinkles in the rare biosphere: pyrosequencing errors can lead to artificial inflation of diversity estimates. Environ. Microbiol. 12:118–123.CrossrefPubMedGoogle Scholar9.Quince C, Lanzen A, Davenport RJ, and Turnbaugh PJ. 2011. Removing noise from pyrosequenced amplicons. BMC Bioinformatics 12:38.CrossrefPubMedGoogle Scholar10.Schloss PD, Gevers D, and Westcott SL. 2011. Reducing the effects of PCR amplification and sequencing artifacts on 16S rRNA-based studies. PLoS One 6:e27310.CrossrefPubMedGoogle Scholar11.Wang Q, Garrity GM, Tiedje JM, and Cole JR. 2007. Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl. Environ. Microbiol. 73:5261–5267.CrossrefPubMedGoogle Scholar12.Bentley DR, Balasubramanian S, Swerdlow HP, Smith GP, Milton J, Brown CG, Hall KP, Evers DJ, Barnes CL, Bignell HR, Boutell JM, Bryant J, Carter RJ, Keira Cheetham R, Cox AJ, Ellis DJ, Flatbush MR, Gormley NA, Humphray SJ, Irving LJ, Karbelashvili MS, Kirk SM, Li H, Liu X, Maisinger KS, Murray LJ, Obradovic B, Ost T, Parkinson ML, Pratt MR, Rasolonjatovo IM, Reed MT, Rigatti R, Rodighiero C, Ross MT, Sabot A, Sankar SV, Scally A, Schroth GP, Smith ME, Smith VP, Spiridou A, Torrance PE, Tzonev SS, et al. 2008. Accurate whole human genome sequencing using reversible terminator chemistry. Nature 456:53–59.CrossrefPubMedGoogle Scholar13.Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Huntley J, Fierer N, Owens SM, Betley J, Fraser L, Bauer M, Gormley N, Gilbert JA, Smith G, and Knight R. 2012. Ultra-high-throughput microbial community analysis on the Illumina HiSeq and MiSeq platforms. ISME J. 6:1621–1624.CrossrefPubMedGoogle Scholar14.Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Lozupone CA, Turnbaugh PJ, Fierer N, and Knight R. 2011. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proc. Natl. Acad. Sci. U. S. A. 108(Suppl. 1):4516–4522.CrossrefPubMedGoogle Scholar15.Werner JJ, Zhou D, Caporaso JG, Knight R, and Angenent LT. 2012. Comparison of Illumina paired-end and single-direction sequencing for microbial 16S rRNA gene amplicon surveys. ISME J. 6:1273–1276.PubMedGoogle Scholar16.Bokulich NA, Subramanian S, Faith JJ, Gevers D, Gordon JI, Knight R, Mills DA, and Caporaso JG. 2013. Quality-filtering vastly improves diversity estimates from Illumina amplicon sequencing. Nat. Methods 10:57–59.PubMedGoogle Scholar17.Masella AP, Bartram AK, Truszkowski JM, Brown DG, and Neufeld JD. 2012. PANDAseq: paired-end assembler for Illumina sequences. BMC Bioinformatics 13:31.CrossrefPubMedGoogle Scholar18.Schloss PD, Schubert AM, Zackular JP, Iverson KD, Young VB, and Petrosino JF. 2012. Stabilization of the murine gut microbiome following weaning. Gut Microbes 3:383–393.CrossrefPubMedGoogle Scholar19.Pruesse E, Quast C, Knittel K, Fuchs BM, Ludwig W, Peplies J, and Glockner FO. 2007. SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Res. 35:7188–7196.CrossrefPubMedGoogle Scholar20.Haas BJ, Gevers D, Earl AM, Feldgarden M, Ward DV, Giannoukos G, Ciulla D, Tabbaa D, Highlander SK, Sodergren E, Methé B, DeSantis TZ, Human Microbiome Consortium, Petrosino JF, Knight R, and Birren BW. 2011. Chimeric 16S rRNA sequence formation and detection in Sanger and 454-pyrosequenced PCR amplicons. Genome Res. 21:494–504.CrossrefPubMedGoogle Scholar21.Simpson JT, Wong K, Jackman SD, Schein JE, Jones SJ, and Birol I. 2009. ABySS: a parallel assembler for short read sequence data. Genome Res. 19:1117–1123.CrossrefPubMedGoogle Scholar22.Schloss PD, Westcott SL, Ryabin T, Hall JR, Hartmann M, Hollister EB, Lesniewski RA, Oakley BB, Parks DH, Robinson CJ, Sahl JW, Stres B, Thallinger GG, Van Horn DJ, and Weber CF. 2009. Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl. Environ. Microbiol. 75:7537–7541.CrossrefPubMedGoogle Scholar23.Schloss PD. 2010. The effects of alignment quality, distance calculation method, sequence filtering, and region on the analysis of 16S rRNA gene-based studies. PLoS Comput. Biol. 6:e1000844.CrossrefPubMedGoogle Scholar24.Schloss PD. 2009. A high-throughput DNA sequence aligner for microbial ecology studies. PLoS One 4:e8230.CrossrefPubMedGoogle Scholar25.Schloss PD. 2013. Secondary structure improves OTU assignments of 16S rRNA gene sequences. ISME J. 7:457–460.PubMedGoogle Scholar26.Edgar RC, Haas BJ, Clemente JC, Quince C, and Knight R. 2011. UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 27:2194–2200.CrossrefPubMedGoogle Scholar27.Schloss PD and Westcott SL. 2011. Assessing and improving methods used in operational taxonomic unit-based approaches for 16S rRNA gene sequence analysis. Appl. Environ. Microbiol. 77:3219–3226.CrossrefPubMedGoogle Scholar28.Yue JC and Clayton MK. 2005. A similarity measure based on species proportions. Commun. Stat. Theory Methods 34:2123–2131.Google ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 79 • Number 17 • 1 September 2013Pages: 5112 - 5120HistoryReceived: 1 April 2013Accepted: 13 June 2013Published online: 7 August 2013Copyright© 2013 American Society for Microbiology.PermissionsRequest permissions for this article.Request PermissionsContributorsAuthorsJames J. KozichDepartment of Microbiology and Immunology, University of Michigan, Ann Arbor, Michigan, USAView all articles by this authorSarah L. WestcottDepartment of Microbiology and Immunology, University of Michigan, Ann Arbor, Michigan, USAView all articles by this authorNielson T. BaxterDepartment of Microbiology and Immunology, University of Michigan, Ann Arbor, Michigan, USAView all articles by this authorSarah K. HighlanderDepartment of Molecular Virology and Microbiology, Baylor College of Medicine, Houston, Texas, USAView all articles by this authorPatrick D. SchlossDepartment of Microbiology and Immunology, University of Michigan, Ann Arbor, Michigan, USAView all articles by this authorNotesAddress correspondence to Patrick D. Schloss, [email protected].Metrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleMarch 2021Time Evolution of Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) in Wastewater during the First Pandemic Wave of COVID-19 in the Metropolitan Area of Barcelona, Spain Gemma Chavarria-Miró, Eduard Anfruns-Estrada, Adán Martínez-Velázquez, Mario Vázquez-Portero, Susana Guix, Miquel Paraira, Belén Galofré, Gloria Sánchez, Rosa M. Pintóand Albert BoschTime Evolution of Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) in Wastewater during the First Pandemic Wave of COVID-19 in the Metropolitan Area of Barcelona, SpainAuthors: Gemma Chavarria-Miró, Eduard Anfruns-Estrada, Adán Martínez-Velázquez, Mario Vázquez-Portero, Susana Guix https://orcid.org/0000-0002-1588-3198, Miquel Paraira, Belén Galofré, Gloria Sánchez, Rosa M. Pintó https://orcid.org/0000-0003-1382-6648 [email protected], and Albert Bosch https://orcid.org/0000-0002-8111-9059 [email protected]DOI: https://doi.org/10.1128/AEM.02750-20Volume 87, Number 711 March 2021ABSTRACTREFERENCESABSTRACTTwo large wastewater treatment plants (WWTP), covering around 2.7 million inhabitants, which represents around 85% of the metropolitan area of Barcelona, were sampled before, during, and after the implementation of a complete lockdown. Five one-step reverse transcriptase quantitative PCR (RT-qPCR) assays, targeting the polymerase (IP2 and IP4), the envelope (E), and the nucleoprotein (N1 and N2) genome regions, were employed for severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) RNA detection in 24-h composite wastewater samples concentrated by polyethylene glycol (PEG) precipitation. SARS-CoV-2 was detected in a sewage sample collected 41 days ahead of the declaration of the first COVID-19 case. The evolution of SARS-CoV-2 genome copies in wastewater evidenced the validity of water-based epidemiology (WBE) to anticipate COVID-19 outbreaks, to evaluate the impact of control measures, and even to estimate the burden of shedders, including presymptomatic, asymptomatic, symptomatic, and undiagnosed cases. For the latter objective, a model was applied for the estimation of the total number of shedders, evidencing a high proportion of asymptomatic infected individuals. In this way, an infection prevalence of 2.0 to 6.5% was figured. On the other hand, proportions of around 0.12% and 0.09% of the total population were determined to be required for positive detection in the two WWTPs. At the end of the lockdown, SARS-CoV-2 RNA apparently disappeared in the WWTPs but could still be detected in grab samples from four urban sewers. Sewer monitoring allowed for location of specific hot spots of COVID-19, enabling the rapid adoption of appropriate mitigation measures.IMPORTANCE Water-based epidemiology (WBE) is a valuable early warning tool for tracking the circulation of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) among the population, including not only symptomatic patients but also asymptomatic, presymptomatic, and misdiagnosed carriers, which represent a high proportion of the infected population. In the specific case of Barcelona, wastewater surveillance anticipated by several weeks not only the original COVID-19 pandemic wave but also the onset of the second wave. In addition, SARS-CoV-2 occurrence in wastewater evidenced the efficacy of the adopted lockdown measures on the circulation of the virus. Health authorities profited from WBE to complement other inputs and adopt rapid and adequate measures to mitigate the effects of the pandemic. For example, sentinel surveillance of specific sewers helped to locate COVID-19 hot spots and to conduct massive numbers of RT-PCR tests among the population.REFERENCES1.Pan Y, Zhang D, Yang P, Poon LLM, Wang Q. 2020. Viral load of SARS-CoV-2 in clinical samples. Lancet Infect Dis 20:411–412.CrossrefPubMedGoogle Scholar2.Zhang J, Wang S, Xue Y. 2020. Fecal specimen diagnosis 2019 novel coronavirus-infected pneumonia. J Med Virol 92:680–682.CrossrefPubMedGoogle Scholar3.Medema G, Heijnen L, Elsinga G, Italiaander R, Brouwer A. 2020. Presence of SARS-Coronavirus-2 in sewage. medRxiv 2020.03.29.20045880.Google Scholar4.Lodder W, de Roda Husman AM. 2020. SARS-CoV-2 in wastewater: potential health risk, but also data source. Lancet Gastroenterol Hepatol 5:533–534.CrossrefPubMedGoogle Scholar5.Ahmed W, Angel N, Edson J, Bibby K, Bivins A, O\'Brien JW, Choi PM, Kitajima M, Simpson SL, Li J, Tscharke B, Verhagen R, Smith WJM, Zaugg J, Dierens L, Hugenholtz P, Thomas KV, Mueller JF. 2020. First confirmed detection of SARS-CoV-2 in untreated wastewater in Australia: a proof of concept for the wastewater surveillance of COVID-19 in the community. Sci Total Environ 728:138764.CrossrefPubMedGoogle Scholar6.La Rosa G, Iaconelli M, Mancini P, Bonanno Ferraro G, Veneri C, Bonadonna L, Lucentini L, Suffredini E. 2020. First detection of SARS-CoV-2 in untreated wastewaters in Italy. Sci Total Environ 736:139652.CrossrefPubMedGoogle Scholar7.Randazzo W, Truchado P, Cuevas-Ferrando E, Simón P, Allende A, Sánchez G. 2020. SARS-CoV-2 RNA in wastewater anticipated COVID-19 occurrence in a low prevalence area. Water Res 181:115942–115942.CrossrefPubMedGoogle Scholar8.Xiao F, Sun J, Xu Y, Li F, Huang X, Li H, Zhao J, Huang J, Zhao J. 2020. Infectious SARS-CoV-2 in feces of patient with severe COVID-19. Emerg Infect Dis 26:1920–1922.CrossrefPubMedGoogle Scholar9.Zang R, Castro MFG, McCune BT, Zeng Q, Rothlauf PW, Sonnek NM, Liu Z, Brulois KF, Wang X, Greenberg HB, Diamond MS, Ciorba MA, Whelan SPJ, Ding S. 2020. TMPRSS2 and TMPRSS4 promote SARS-CoV-2 infection of human small intestinal enterocytes. Sci Immunol 5:eabc3582.CrossrefPubMedGoogle Scholar10.Bullard J, Dust K, Funk D, Strong JE, Alexander D, Garnett L, Boodman C, Bello A, Hedley A, Schiffman Z, Doan K, Bastien N, Li Y, Van Caeseele PG, Poliquin G. 2020. Predicting infectious severe acute respiratory syndrome coronavirus 2 from diagnostic samples. Clin Infect Dis 71:2663–2666.CrossrefPubMedGoogle Scholar11.Wölfel R, Corman VM, Guggemos W, Seilmaier M, Zange S, Muller MA, Niemeyer D, Jones TC, Vollmar P, Rothe C, Hoelscher M, Bleicker T, Brunink S, Schneider J, Ehmann R, Zwirglmaier K, Drosten C, Wendtner C. 2020. Virological assessment of hospitalized patients with COVID-2019. Nature 581:465–469.CrossrefPubMedGoogle Scholar12.Heaton KW, Radvan J, Cripps H, Mountford RA, Braddon FE, Hughes AO. 1992. Defecation frequency and timing, and stool form in the general population: a prospective study. Gut 33:818–824.CrossrefPubMedGoogle Scholar13.Artesi M, Bontems S, Göbbels P, Franckh M, Maes P, Boreux R, Meex C, Melin P, Hayette M-P, Bours V, Durkin K. 2020. A recurrent mutation at position 26340 of SARS-CoV-2 is associated with failure of the E gene quantitative reverse transcription-PCR utilized in a commercial dual-target diagnostic assay. J Clin Microbiol 58:e01598-20.CrossrefPubMedGoogle Scholar14.Sims N, Kasprzyk-Hordern B. 2020. Future perspectives of wastewater-based epidemiology: monitoring infectious disease spread and resistance to the community level. Environ Int 139:105689.CrossrefPubMedGoogle Scholar15.Barras C. 2018. Going to waste. Nat Med 24:1484–1487.CrossrefPubMedGoogle Scholar16.Corman VM, Landt O, Kaiser M, Molenkamp R, Meijer A, Chu DK, Bleicker T, Brunink S, Schneider J, Schmidt ML, Mulders DG, Haagmans BL, van der Veer B, van den Brink S, Wijsman L, Goderski G, Romette JL, Ellis J, Zambon M, Peiris M, Goossens H, Reusken C, Koopmans MP, Drosten C. 2020. Detection of 2019 novel coronavirus (2019-nCoV) by real-time RT-PCR. Euro Surveill 25:2000045. https://www.eurosurveillance.org/content/10.2807/1560-7917.ES.2020.25.3.2000045.CrossrefGoogle Scholar17.Tokars JI, Olsen SJ, Reed C. 2018. Seasonal incidence of symptomatic influenza in the United States. Clin Infect Dis 66:1511–1518.CrossrefPubMedGoogle Scholar18.Pollán M, Pérez-Gómez B, Pastor-Barriuso R, Oteo J, Hernán MA, Pérez-Olmeda M, Sanmartín JL, Fernández-García A, Cruz I, Fernández de Larrea N, Molina M, Rodríguez-Cabrera F, Martín M, Merino-Amador P, León Paniagua J, Muñoz-Montalvo JF, Blanco F, Yotti R, Blanco F, Gutiérrez Fernández R, Martín M, Mezcua Navarro S, Molina M, Muñoz-Montalvo JF, Salinero Hernández M, Sanmartín JL, Cuenca-Estrella M, Yotti R, León Paniagua J, Fernández de Larrea N, Fernández-Navarro P, Pastor-Barriuso R, Pérez-Gómez B, Pollán M, Avellón A, Fedele G, Fernández-García A, Oteo Iglesias J, Pérez Olmeda MT, Cruz I, Fernandez Martinez ME, Rodríguez-Cabrera FD, Hernán MA, Padrones Fernández S, Rumbao Aguirre JM, Navarro Marí JM, Palop Borrás B, Pérez Jiménez AB, Rodríguez-Iglesias M, Calvo Gascón AM, et al. 2020. Prevalence of SARS-CoV-2 in Spain (ENE-COVID): a nationwide, population-based seroepidemiological study. Lancet 396:535–544.CrossrefPubMedGoogle Scholar19.Havers FP, Reed C, Lim T, Montgomery JM, Klena JD, Hall AJ, Fry AM, Cannon DL, Chiang CF, Gibbons A, Krapiunaya I, Morales-Betoulle M, Roguski K, Rasheed MAU, Freeman B, Lester S, Mills L, Carroll DS, Owen SM, Johnson JA, Semenova V, Blackmore C, Blog D, Chai SJ, Dunn A, Hand J, Jain S, Lindquist S, Lynfield R, Pritchard S, Sokol T, Sosa L, Turabelidze G, Watkins SM, Wiesman J, Williams RW, Yendell S, Schiffer J, Thornburg NJ. 2020. Seroprevalence of Antibodies to SARS-CoV-2 in 10 Sites in the United States, March 23-May 12, 2020. JAMA Intern Med 180:1576.CrossrefGoogle Scholar20.Hart OE, Halden RU. 2020. Computational analysis of SARS-CoV-2/COVID-19 surveillance by wastewater-based epidemiology locally and globally: feasibility, economy, opportunities and challenges. Sci Total Environ 730:138875.CrossrefPubMedGoogle Scholar21.Coma E, Mora N, Prats-Uribe A, Fina F, Prieto-Alhambra D, Medina-Peralta M. 2020. Excess cases of influenza suggest an earlier start to the coronavirus epidemic in Spain than official figures tell us: an analysis of primary care electronic medical records from over 6 million people from Catalonia. medRxiv 2020.04.09.20056259.Google Scholar22.Young BE, Ong SWX, Kalimuddin S, Low JG, Tan SY, Loh J, Ng OT, Marimuthu K, Ang LW, Mak TM, Lau SK, Anderson DE, Chan KS, Tan TY, Ng TY, Cui L, Said Z, Kurupatham L, Chen MI, Chan M, Vasoo S, Wang LF, Tan BH, Lin RTP, Lee VJM, Leo YS, Lye DC, Singapore 2019 Novel Coronavirus Outbreak Research Team. 2020. Epidemiologic features and clinical course of patients infected with SARS-CoV-2 in Singapore. JAMA 323:1488–1494.CrossrefPubMedGoogle Scholar23.Lauer SA, Grantz KH, Bi Q, Jones FK, Zheng Q, Meredith HR, Azman AS, Reich NG, Lessler J. 2020. The incubation period of coronavirus disease 2019 (COVID-19) from publicly reported confirmed cases: estimation and application. Ann Intern Med 172:577–582.CrossrefPubMedGoogle Scholar24.Wang Y, Kang H, Liu X, Tong Z. 2020. Asymptomatic cases with SARS-CoV-2 infection. J Med Virol 92:1401–1403.CrossrefPubMedGoogle Scholar25.Blanco A, Abid I, Al-Otaibi N, Perez-Rodriguez FJ, Fuentes C, Guix S, Pinto RM, Bosch A. 2019. Glass wool concentration optimization for the detection of enveloped and non-enveloped waterborne viruses. Food Environ Virol 11:184–192.CrossrefPubMedGoogle Scholar26.Sanchez CM, Jimenez G, Laviada MD, Correa I, Sune C, Bullido M, Gebauer F, Smerdou C, Callebaut P, Escribano JM, Enjuanes L. 1990. Antigenic homology among coronaviruses related to transmissible gastroenteritis virus. Virology 174:410–417.CrossrefPubMedGoogle Scholar27.Wu Y, Guo C, Tang L, Hong Z, Zhou J, Dong X, Yin H, Xiao Q, Tang Y, Qu X, Kuang L, Fang X, Mishra N, Lu J, Shan H, Jiang G, Huang X. 2020. Prolonged presence of SARS-CoV-2 viral RNA in faecal samples. Lancet Gastroenterol Hepatol 5:434–435.CrossrefPubMedGoogle Scholar28.Penn R, Ward BJ, Strande L, Maurer M. 2018. Review of synthetic human faeces and faecal sludge for sanitation and wastewater research. Water Res 132:222–240.CrossrefPubMedGoogle Scholar29.Rose C, Parker A, Jefferson B, Cartmell E. 2015. The characterization of feces and urine: a review of the literature to inform advanced treatment technology. Crit Rev Environ Sci Technol 45:1827–1879.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 87 • Number 7 • 11 March 2021eLocator: e02750-20Editor: Christopher A. ElkinsCenters for Disease Control and PreventionHistoryReceived: 9 November 2020Accepted: 15 January 2021Published online: 22 January 2021Copyright© 2021 American Society for Microbiology. All Rights Reserved.PermissionsRequest permissions for this article.Request PermissionsKEYWORDSSARS-CoV-2COVID-19epidemiologysurveillanceearly warningsewageContributorsAuthorsGemma Chavarria-MiróEnteric Virus Laboratory, Department of Genetics, Microbiology, and Statistics, Section of Microbiology, Virology, and Biotechnology, School of Biology, University of Barcelona, Barcelona, SpainInstitute of Nutrition and Food Safety, University of Barcelona, Barcelona, SpainView all articles by this authorEduard Anfruns-EstradaEnteric Virus Laboratory, Department of Genetics, Microbiology, and Statistics, Section of Microbiology, Virology, and Biotechnology, School of Biology, University of Barcelona, Barcelona, SpainInstitute of Nutrition and Food Safety, University of Barcelona, Barcelona, SpainView all articles by this authorAdán Martínez-VelázquezEnteric Virus Laboratory, Department of Genetics, Microbiology, and Statistics, Section of Microbiology, Virology, and Biotechnology, School of Biology, University of Barcelona, Barcelona, SpainInstitute of Nutrition and Food Safety, University of Barcelona, Barcelona, SpainView all articles by this authorMario Vázquez-PorteroEnteric Virus Laboratory, Department of Genetics, Microbiology, and Statistics, Section of Microbiology, Virology, and Biotechnology, School of Biology, University of Barcelona, Barcelona, SpainInstitute of Nutrition and Food Safety, University of Barcelona, Barcelona, SpainView all articles by this authorSusana Guix https://orcid.org/0000-0002-1588-3198Enteric Virus Laboratory, Department of Genetics, Microbiology, and Statistics, Section of Microbiology, Virology, and Biotechnology, School of Biology, University of Barcelona, Barcelona, SpainInstitute of Nutrition and Food Safety, University of Barcelona, Barcelona, SpainView all articles by this authorMiquel ParairaAigües de Barcelona, Barcelona, SpainView all articles by this authorBelén GalofréAigües de Barcelona, Barcelona, SpainView all articles by this authorGloria SánchezDepartment of Preservation and Food Safety Technologies, Institute of Agrochemistry and Food Technology, IATA-CSIC, Paterna, Valencia, SpainView all articles by this authorRosa M. Pintó https://orcid.org/0000-0003-1382-6648 [email protected]Enteric Virus Laboratory, Department of Genetics, Microbiology, and Statistics, Section of Microbiology, Virology, and Biotechnology, School of Biology, University of Barcelona, Barcelona, SpainInstitute of Nutrition and Food Safety, University of Barcelona, Barcelona, SpainView all articles by this authorAlbert Bosch https://orcid.org/0000-0002-8111-9059 [email protected]Enteric Virus Laboratory, Department of Genetics, Microbiology, and Statistics, Section of Microbiology, Virology, and Biotechnology, School of Biology, University of Barcelona, Barcelona, SpainInstitute of Nutrition and Food Safety, University of Barcelona, Barcelona, SpainView all articles by this authorEditorChristopher A. ElkinsEditorCenters for Disease Control and PreventionMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleJuly 2021A Bioengineered Nisin Derivative To Control Streptococcus uberis Biofilms Mariana Pérez-Ibarreche, Des Field, R. Paul Rossand Colin HillA Bioengineered Nisin Derivative To Control Streptococcus uberis BiofilmsAuthors: Mariana Pérez-Ibarreche, Des Field [email protected], R. Paul Ross, and Colin Hill https://orcid.org/0000-0002-8527-1445 [email protected]DOI: https://doi.org/10.1128/AEM.00391-21Volume 87, Number 1627 July 2021ABSTRACTREFERENCESABSTRACTAntimicrobial peptides are evolving as novel therapeutic options against the increasing problem of multidrug-resistant microorganisms, and nisin is one such avenue. However, some bacteria possess a specific nisin resistance system (NSR), which cleaves the peptide reducing its bactericidal efficacy. NSR-based resistance was identified in strains of Streptococcus uberis, a ubiquitous pathogen that causes mastitis in dairy cattle. Previous studies have demonstrated that a nisin A derivative termed nisin PV, featuring S29P and I30V, exhibits enhanced resistance to proteolytic cleavage by NSR. Our objective was to investigate the ability of this nisin derivative to eradicate and inhibit biofilms of S. uberis DPC 5344 and S. uberis ATCC 700407 (nsr+) using crystal violet (biomass), 2,3-bis-(2-methoxy-4-nitro-5-sulfophenyl)-2H-tetrazolium-5-carboxanilide (XTT) (viability) assays, and confocal microscopy (viability and architecture). When preestablished biofilms were assessed, both peptides reduced biofilm biomass by over 60% compared to that of the untreated controls. However, a 42% higher reduction in viability was observed following treatment with nisin PV compared to that of nisin A. Accordingly, confocal microscopy analysis revealed significantly more dead cells on the biofilm upper surface and a reduced thickness following treatment with nisin PV. When biofilm inhibition was assessed, nisin PV inhibited biofilm formation and decreased viability up to 56% and 85% more than nisin A, respectively. Confocal microscopy analysis revealed a lack of biofilm for S. uberis ATCC 700407 and only dead cells for S. uberis DPC 5344. These results suggest that nisin PV is a promising alternative to effectively reduce the biofilm formation of S. uberis strains carrying NSR.IMPORTANCE One of the four most prevalent species of bovine mastitis-causing pathogens is S. uberis. Its ability to form biofilms confers on the bacteria greater resistance to antibiotics, requiring higher doses to be more effective. In a bid to limit antibiotic resistance development, the need for alternative antimicrobials is paramount. Bacteriocins such as nisin represent one such alternative that could alleviate the impact of mastitis caused by S. uberis. However, many strains of S. uberis have been shown to possess nisin resistance determinants, such as the nisin resistance protein (NSR). In this study, we demonstrate the ability of nisin and a nisin derivative termed PV that is insensitive to NSR to prevent and remove biofilms of NSR-producing S. uberis strains. These findings will add new information to the antimicrobial bacteriocins and control of S. uberis research fields specifically in relation to biofilms and nsr+ mastitis-associated strains.REFERENCES1.Dufour S, Labrie J, Jacques M. 2019. The mastitis pathogens culture collection. Microbiol Resour Announc 8:e00133-19.CrossrefPubMedGoogle Scholar2.Guimarães JLB, Brito MAVP, Lange CC, Silva MR, Ribeiro JB, Mendonça LC, Mendonça JFM, Souza GN. 2017. Estimate of the economic impact of mastitis: a case study in a Holstein dairy herd under tropical conditions. Prev Vet Med 142:46–50.CrossrefPubMedGoogle Scholar3.Rollin E, Dhuyvetter KC, Overton MW. 2015. The cost of clinical mastitis in the first 30 days of lactation: an economic modeling tool. Prev Vet Med 122:257–264.CrossrefPubMedGoogle Scholar4.Donlan RM, Costerton JW. 2002. Biofilms: survival mechanisms of clinically relevant microorganisms. Clin Microbiol Rev 15:167–193.CrossrefPubMedGoogle Scholar5.Flemming H-C, Wingender J. 2010. The biofilm matrix. Nat Rev Microbiol 8:623–633.CrossrefPubMedGoogle Scholar6.Dieser SA, Fessia AS, Ferrari MP, Raspanti CG, Odierno LM. 2017. Streptococcus uberis: In vitro biofilm production in response to carbohydrates and skim milk. Rev Argent Microbiol 49:305–310.CrossrefPubMedGoogle Scholar7.Schönborn S, Wente N, Paduch J-H, Krömker V. 2017. In vitro ability of mastitis causing pathogens to form biofilms. J Dairy Res 84:198–201.CrossrefPubMedGoogle Scholar8.Royster E, Wagner S. 2015. Treatment of mastitis in cattle. Vet Clin North Am Food Anim Pract 31:17–46.CrossrefPubMedGoogle Scholar9.Prince A, Sandhu P, Ror P, Dash E, Sharma S, Arakha M, Jha S, Akhter Y, Saleem M. 2016. Lipid-II independent antimicrobial mechanism of nisin depends on its crowding and degree of oligomerization. Sci Rep 6:37908.CrossrefPubMedGoogle Scholar10.Williams GC, Delves-Broughton J. 2003. Nisin, p 4128–4135. In Caballero B (ed), Encyclopedia of food sciences and nutrition, 2nd ed. Academic Press, Oxford, United Kingdom.CrossrefGoogle Scholar11.Khosa S, AlKhatib Z, Smits SHJ. 2013. NSR from Streptococcus agalactiae confers resistance against nisin and is encoded by a conserved nsr operon. Biol Chem 394:1543–1549.CrossrefPubMedGoogle Scholar12.Assoni L, Milani B, Carvalho MR, Nepomuceno LN, Waz NT, Guerra MES, Converso TR, Darrieux M. 2020. Resistance mechanisms to antimicrobial peptides in Gram-positive bacteria. Front Microbiol 11:593215.CrossrefPubMedGoogle Scholar13.Sun Z, Zhong J, Liang X, Liu J, Chen X, Huan L. 2009. Novel mechanism for nisin resistance via proteolytic degradation of nisin by the nisin resistance protein NSR. Antimicrob Agents Chemother 53:1964–1973.CrossrefPubMedGoogle Scholar14.Boakes S, Ayala T, Herman M, Appleyard AN, Dawson MJ, Cortés J. 2012. Generation of an actagardine A variant library through saturation mutagenesis. Appl Microbiol Biotechnol 95:1509–1517.CrossrefPubMedGoogle Scholar15.Chen S, Wilson-Stanford S, Cromwell W, Hillman JD, Guerrero A, Allen CA, Sorg JA, Smith L. 2013. Site-directed mutations in the lanthipeptide mutacin 1140. Appl Environ Microbiol 79:4015–4023.CrossrefPubMedGoogle Scholar16.Field D, Connor PMO, Cotter PD, Hill C, Ross RP. 2008. The generation of nisin variants with enhanced activity against specific Gram-positive pathogens. Mol Microbiol 69:218–230.CrossrefPubMedGoogle Scholar17.Islam MR, Shioya K, Nagao J, Nishie M, Jikuya H, Zendo T, Nakayama J, Sonomoto K. 2009. Evaluation of essential and variable residues of nukacin ISK-1 by NNK scanning. Mol Microbiol 72:1438–1447.CrossrefPubMedGoogle Scholar18.Field D, Blake T, Mathur H, O\' Connor PM, Cotter PD, Ross RP, Hill C. 2019. Bioengineering nisin to overcome the nisin resistance protein. Mol Microbiol 111:717–731.CrossrefPubMedGoogle Scholar19.Stepanović S, Vuković D, Dakić I, Savić B, Švabić-Vlahović M. 2000. A modified microtiter-plate test for quantification of staphylococcal biofilm formation. J Microbiol Methods 40:175–179.CrossrefPubMedGoogle Scholar20.Barrett DJ, Healy AM, Leonard FC, Doherty ML. 2005. Prevalence of pathogens causing subclinical mastitis in 15 dairy herds in the Republic of Ireland. Ir Vet J 58:333.CrossrefPubMedGoogle Scholar21.Klaas IC, Zadoks RN. 2018. An update on environmental mastitis: challenging perceptions. Transbound Emerg Dis 65(Suppl 1):166–185.CrossrefPubMedGoogle Scholar22.Blowey RW, Edmondson P. 2010. Mastitis control in dairy herds. CABI, Oxfordshire, United Kingdom.CrossrefGoogle Scholar23.Pyörälä S. 2009. Treatment of mastitis during lactation. Ir Vet J 62:S40–S44.CrossrefPubMedGoogle Scholar24.Haenni M, Lupo A, Madec J-Y. 2018. Antimicrobial resistance in Streptococcus spp. Microbiol Spectr 6:6.2.09.CrossrefGoogle Scholar25.McDougall S, Hussein H, Petrovski K. 2014. Antimicrobial resistance in Staphylococcus aureus, Streptococcus uberis and Streptococcus dysgalactiae from dairy cows with mastitis. N Z Vet J 62:68–76.CrossrefPubMedGoogle Scholar26.Petrovski KR, Grinberg A, Williamson NB, Abdalla ME, Lopez-Villalobos N, Parkinson TJ, Tucker IG, Rapnicki P. 2015. Susceptibility to antimicrobials of mastitis-causing Staphylococcus aureus, Streptococcus uberis and Str. dysgalactiae from New Zealand and the USA as assessed by the disk diffusion test. Aust Vet J 93:227–233.CrossrefPubMedGoogle Scholar27.Tomazi T, Freu G, Alves BG, de Souza Filho AF, Heinemann MB, Veiga Dos Santos M. 2019. Genotyping and antimicrobial resistance of Streptococcus uberis isolated from bovine clinical mastitis. PLoS One 14:e0223719.CrossrefPubMedGoogle Scholar28.McDougall S, Clausen L, Ha H-J, Gibson I, Bryan M, Hadjirin N, Lay E, Raisen C, Ba X, Restif O, Parkhill J, Holmes MA. 2020. Mechanisms of β-lactam resistance of Streptococcus uberis isolated from bovine mastitis cases. Vet Microbiol 242:108592.CrossrefPubMedGoogle Scholar29.Cotter PD, Ross RP, Hill C. 2013. Bacteriocins - a viable alternative to antibiotics? Nat Rev Microbiol 11:95–105.CrossrefPubMedGoogle Scholar30.Maurer JJ, Mattingly SJ. 1991. Molecular analysis of lipoteichoic acid from Streptococcus agalactiae. J Bacteriol 173:487–494.CrossrefPubMedGoogle Scholar31.Rea MC, Ross RP, Cotter PD, Hill C. 2011. Classification of bacteriocins from Gram-positive bacteria, p 29–53. In Drider D, Rebuffat S (ed), Prokaryotic antimicrobial peptides: from genes to applications. Springer, New York, NY.CrossrefGoogle Scholar32.Khosa S, Lagedroste M, Smits SHJ. 2016. Protein defense systems against the lantibiotic nisin: function of the immunity protein NisI and the resistance protein NSR. Front Microbiol 7:504.CrossrefPubMedGoogle Scholar33.Field D, Cotter PD, Ross RP, Hill C. 2015. Bioengineering of the model lantibiotic nisin. Bioengineered 6:187–192.CrossrefPubMedGoogle Scholar34.Li Q, Montalban-Lopez M, Kuipers OP. 2018. Increasing the antimicrobial activity of nisin-based lantibiotics against Gram-negative pathogens. Appl Environ Microbiol 84:e00052-18.CrossrefPubMedGoogle Scholar35.Reinoso EB. 2017. Bovine mastitis caused by Streptococcus uberis: virulence factors and biofilm. J Microb Biochem Technol 9:237–243.CrossrefGoogle Scholar36.Singh S, Singh SK, Chowdhury I, Singh R. 2017. Understanding the mechanism of bacterial biofilms resistance to antimicrobial agents. Open Microbiol J 11:53–62.CrossrefPubMedGoogle Scholar37.Verderosa AD, Totsika M, Fairfull-Smith KE. 2019. Bacterial biofilm eradication agents: a current review. Front Chem 7:824.CrossrefPubMedGoogle Scholar38.Moliva MV, Cerioli F, Reinoso EB. 2017. Evaluation of environmental and nutritional factors and sua gene on in vitro biofilm formation of Streptococcus uberis isolates. Microb Pathog 107:144–148.CrossrefPubMedGoogle Scholar39.Hayes K, Field D, Hill C, O\'Halloran F, Cotter L. 2019. A novel bioengineered derivative of nisin displays enhanced antimicrobial activity against clinical Streptococcus agalactiae isolates. J Glob Antimicrob Resist 19:14–21.CrossrefPubMedGoogle Scholar40.Reiners J, Lagedroste M, Ehlen K, Leusch S, Zaschke-Kriesche J, Smits SHJ. 2017. The N-terminal region of nisin is important for the BceAB-type ABC transporter NsrFP from Streptococcus agalactiae COH1. Front Microbiol 8:1643.CrossrefPubMedGoogle Scholar41.Lagedroste M, Reiners J, Smits SHJ, Schmitt L. 2019. Systematic characterization of position one variants within the lantibiotic nisin. Sci Rep 9:935.CrossrefPubMedGoogle Scholar42.Castelani L, Arcaro JRP, Braga JEP, Bosso AS, Moura Q, Esposito F, Sauter IP, Cortez M, Lincopan N. 2019. Short communication: activity of nisin, lipid bilayer fragments and cationic nisin-lipid nanoparticles against multidrug-resistant Staphylococcus spp. isolated from bovine mastitis. J Dairy Sci 102:678–683.CrossrefPubMedGoogle Scholar43.Ceotto-Vigoder H, Marques SLS, Santos INS, Alves MDB, Barrias ES, Potter A, Alviano DS, Bastos MCF. 2016. Nisin and lysostaphin activity against preformed biofilm of Staphylococcus aureus involved in bovine mastitis. J Appl Microbiol 121:101–114.CrossrefPubMedGoogle Scholar44.Field D, O\' Connor R, Cotter PD, Ross RP, Hill C. 2016. In vitro activities of nisin and nisin derivatives alone and in combination with antibiotics against Staphylococcus biofilms. Front Microbiol 7:508.CrossrefPubMedGoogle Scholar45.Kitazaki K, Koga S, Nagatoshi K, Kuwano K, Zendo T, Nakayama J, Sonomoto K, Ano H, Katamoto H. 2017. In vitro synergistic activities of cefazolin and nisin A against mastitis pathogens. J Vet Med Sci 79:1472–1479.CrossrefPubMedGoogle Scholar46.Cerioli MF, Moliva MV, Cariddi LN, Reinoso EB. 2018. Effect of the essential oil of Minthostachys verticillata (Griseb.) Epling and limonene on biofilm production in pathogens causing bovine mastitis. Front Vet Sci 5:146.CrossrefPubMedGoogle Scholar47.Montironi ID, Cariddi LN, Reinoso EB. 2016. Evaluation of the antimicrobial efficacy of Minthostachys verticillata essential oil and limonene against Streptococcus uberis strains isolated from bovine mastitis. Rev Argent Microbiol 48:210–216.CrossrefPubMedGoogle Scholar48.Corbin A, Pitts B, Parker A, Stewart PS. 2011. Antimicrobial penetration and efficacy in an in vitro oral biofilm model. Antimicrob Agents Chemother 55:3338–3344.CrossrefPubMedGoogle Scholar49.Angelopoulou A, Field D, Pérez-Ibarreche M, Warda AK, Hill C, Ross RP. 2020. Vancomycin and nisin A are effective against biofilms of multi-drug resistant Staphylococcus aureus isolates from human milk. PLoS One 15:e0233284.CrossrefPubMedGoogle Scholar50.Shin JM, Ateia I, Paulus JR, Liu H, Fenno JC, Rickard AH, Kapila YL. 2015. Antimicrobial nisin acts against saliva derived multi-species biofilms without cytotoxicity to human oral cells. Front Microbiol 6:617.CrossrefPubMedGoogle Scholar51.Zhao M, Qu Y, Liu J, Mai S, Gu L. 2020. A universal adhesive incorporating antimicrobial peptide nisin: effects on Streptococcus mutans and saliva-derived multispecies biofilms. Odontology 108:376–385.CrossrefPubMedGoogle Scholar52.Melchior MB, Fink‐Gremmels J, Gaastra W. 2006. Comparative assessment of the antimicrobial susceptibility of Staphylococcus aureus isolates from bovine mastitis in biofilm versus planktonic culture. J Vet Med Series B 53:326–332.CrossrefPubMedGoogle Scholar53.Olson ME, Ceri H, Morck DW, Buret AG, Read RR. 2002. Biofilm bacteria: formation and comparative susceptibility to antibiotics. Can J Vet Res 66:86–92.PubMedGoogle Scholar54.Tremblay YDN, Caron V, Blondeau A, Messier S, Jacques M. 2014. Biofilm formation by coagulase-negative staphylococci: impact on the efficacy of antimicrobials and disinfectants commonly used on dairy farms. Vet Microbiol 172:511–518.CrossrefPubMedGoogle Scholar55.D\'Urzo N, Martinelli M, Pezzicoli A, De Cesare V, Pinto V, Margarit I, Telford JL, Maione D, Members of the DEVANI Study Group. 2014. Acidic pH strongly enhances in vitro biofilm formation by a subset of hypervirulent ST-17 Streptococcus agalactiae strains. Appl Environ Microbiol 80:2176–2185.CrossrefPubMedGoogle Scholar56.Seil JT, Webster TJ. 2012. Antimicrobial applications of nanotechnology: methods and literature. Int J Nanomedicine 7:2767–2781.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 87 • Number 16 • 27 July 2021eLocator: e00391-21Editor: M. Julia PettinariUniversity of Buenos AiresHistoryReceived: 25 February 2021Accepted: 1 June 2021Published online: 9 June 2021Copyright© 2021 Pérez-Ibarreche et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license.PermissionsRequest permissions for this article.Request PermissionsKEYWORDSStreptococcus uberisbiofilmsNSRantimicrobial agentsnisin Anisin PVContributorsAuthorsMariana Pérez-IbarrecheAPC Microbiome Ireland, University College Cork, Cork, IrelandView all articles by this authorDes Field [email protected]APC Microbiome Ireland, University College Cork, Cork, IrelandView all articles by this authorR. Paul RossAPC Microbiome Ireland, University College Cork, Cork, IrelandView all articles by this authorColin Hill https://orcid.org/0000-0002-8527-1445 [email protected]APC Microbiome Ireland, University College Cork, Cork, IrelandSchool of Microbiology, University College Cork, Cork, IrelandView all articles by this authorEditorM. Julia PettinariEditorUniversity of Buenos AiresMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleJanuary 2021Airborne Disinfection by Dry Fogging Efficiently Inactivates Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2), Mycobacteria, and Bacterial Spores and Shows Limitations of Commercial Spore Carriers Jan Schinköthe, Hendrik A. Scheinemann, Sandra Diederich, Holger Freese, Michael Eschbaumer, Jens P. Teifkeand Sven ReicheAirborne Disinfection by Dry Fogging Efficiently Inactivates Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2), Mycobacteria, and Bacterial Spores and Shows Limitations of Commercial Spore CarriersAuthors: Jan Schinköthe, Hendrik A. Scheinemann https://orcid.org/0000-0002-4275-7672, Sandra Diederich, Holger Freese, Michael Eschbaumer, Jens P. Teifke https://orcid.org/0000-0001-7168-4970, and Sven Reiche https://orcid.org/0000-0001-7575-2483DOI: https://doi.org/10.1128/AEM.02019-20Volume 87, Number 315 January 2021ABSTRACTREFERENCESABSTRACTAirborne disinfection of high-containment facilities before maintenance or between animal studies is crucial. Commercial spore carriers (CSC) coated with 106 spores of Geobacillus stearothermophilus are often used to assess the efficacy of disinfection. We used quantitative carrier testing (QCT) procedures to compare the sensitivity of CSC with that of surrogates for nonenveloped and enveloped viruses, including severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), mycobacteria, and spores, to an aerosolized mixture of peroxyacetic acid and hydrogen peroxide (aPAA-HP). We then used the QCT methodology to determine relevant process parameters to develop and validate effective disinfection protocols (≥4-log10 reduction) in various large and complex facilities. Our results demonstrate that aPAA-HP is a highly efficient procedure for airborne room disinfection. Relevant process parameters such as temperature and relative humidity can be wirelessly monitored. Furthermore, we found striking differences in inactivation efficacies against some of the tested microorganisms. Overall, we conclude that dry fogging a mixture of aPAA-HP is highly effective against a broad range of microorganisms as well as material compatible with relevant concentrations. Furthermore, CSC are artificial bioindicators with lower resistance and thus should not be used for validating airborne disinfection when microorganisms other than viruses have to be inactivated.IMPORTANCE Airborne disinfection is not only of crucial importance for the safe operation of laboratories and animal rooms where infectious agents are handled but also can be used in public health emergencies such as the current severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) pandemic. We show that dry fogging an aerosolized mixture of peroxyacetic acid and hydrogen peroxide (aPAA-HP) is highly microbicidal, efficient, fast, robust, environmentally neutral, and a suitable airborne disinfection method. In addition, the low concentration of dispersed disinfectant, particularly for enveloped viral pathogens such as SARS-CoV-2, entails high material compatibility. For these reasons and due to the relative simplicity of the procedure, it is an ideal disinfection method for hospital wards, ambulances, public conveyances, and indoor community areas. Thus, we conclude that this method is an excellent choice for control of the current SARS-CoV-2 pandemic.REFERENCES1.Ackland NR, Hinton MR, Denmeade KR. 1980. Controlled formaldehyde fumigation system. Appl Environ Microbiol 39:480–487.CrossrefPubMedGoogle Scholar2.Lach VH. 1990. A study of conventional formaldehyde fumigation methods. J Appl Bacteriol 68:471–477.CrossrefPubMedGoogle Scholar3.Krause J, McDonnell G, Riedesel H. 2001. Biodecontamination of animal rooms and heat-sensitive equipment with vaporized hydrogen peroxide. Contemp Top Lab Anim Sci 40:18–21.PubMedGoogle Scholar4.Krishnan J, Berry J, Fey G, Wagener S. 2006. Vaporized hydrogen peroxide-based biodecontamination of a high-containment laboratory under negative pressure. Appl Biosaf 11:74–80.CrossrefGoogle Scholar5.Vannier M, Chewins J. 2019. Hydrogen peroxide vapour is an effective replacement for formaldehyde in a BSL4 foot and mouth disease vaccine manufacturing facility. Lett Appl Microbiol 69:237–245.CrossrefPubMedGoogle Scholar6.Gregersen J-P, Roth B. 2012. Inactivation of stable viruses in cell culture facilities by peracetic acid fogging. Biologicals 40:282–287.CrossrefPubMedGoogle Scholar7.Krishnan J, Fey G, Stansfield C, Landry L, Nguy H, Klassen S, Robertson C. 2012. Evaluation of a dry fogging system for laboratory decontamination. Appl Biosaf 17:132–141.CrossrefGoogle Scholar8.McDonnell G, Russell AD. 1999. Antiseptics and disinfectants. activity, action, and resistance. Clin Microbiol Rev 12:147–179.CrossrefPubMedGoogle Scholar9.Kitis M. 2004. Disinfection of wastewater with peracetic acid: a review. Environ Int 30:47–55.CrossrefPubMedGoogle Scholar10.Jones LA, Hoffman RK, Phillips CR. 1967. Sporicidal activity of peracetic acid and β-propiolactone at subzero temperatures. Appl Microbiol 15:357–362.CrossrefPubMedGoogle Scholar11.Mücke H, Sprössig M. 1967. On the antimicrobial effect of peracetic acid. 1. Preparation, analysis and properties of peracetic acid. Pharmazie 22:444–445.PubMedGoogle Scholar12.Pottage T, Macken S, Walker JT, Bennett AM. 2012. Meticillin-resistant Staphylococcus aureus is more resistant to vaporized hydrogen peroxide than commercial Geobacillus stearothermophilus biological indicators. J Hosp Infect 80:41–45.CrossrefPubMedGoogle Scholar13.Anonymous. 2020. EN 17272:2020. Chemical disinfectants and antiseptics. Methods of airborne room disinfection by automated process: determination of bactericidal, mycobactericidal, sporicidal, fungicidal, yeasticidal, virucidal and phagocidal activities. Deutsche Industrienorm, Berlin, Germany.Google Scholar14.Springthorpe VS, Sattar SA. 2005. Carrier tests to assess microbicidal activities of chemical disinfectants for use on medical devices and environmental surfaces. J AOAC Int 88:182–201.CrossrefPubMedGoogle Scholar15.Springthorpe VS, Sattar SA. 2007. Application of a quantitative carrier test to evaluate microbicides against mycobacteria. J AOAC Int 90:817–824.CrossrefPubMedGoogle Scholar16.Anonymous. 2013. EN 14349:2013–02. Chemical disinfectants and antiseptics. Quantitative surface test for the evaluation of bactericidal activity of chemical disinfectants and antiseptics used in the veterinary area on non-porous surfaces without mechanical action: test method and requirements (phase 2, step 2). Deutsche Industrienorm, Berlin, Germnay.Google Scholar17.Kaspari O, Lemmer K, Becker S, Lochau P, Howaldt S, Nattermann H, Grunow R. 2014. Decontamination of a BSL3 laboratory by hydrogen peroxide fumigation using three different surrogates for Bacillus anthracis spores. J Appl Microbiol 117:1095–1103.CrossrefPubMedGoogle Scholar18.Raguse M, Fiebrandt M, Stapelmann K, Madela K, Laue M, Lackmann J-W, Thwaite JE, Setlow P, Awakowicz P, Moeller R. 2016. Improvement of biological indicators by uniformly distributing Bacillus subtilis spores in monolayers to evaluate enhanced spore decontamination technologies. Appl Environ Microbiol 82:2031–2038.CrossrefPubMedGoogle Scholar19.Ijaz MK, Rubino J. 2008. Should test methods for disinfectants use vertebrate viruses dried on carriers to advance virucidal claims? Infect Control Hosp Epidemiol 29:192–194.CrossrefPubMedGoogle Scholar20.Chosewood LC, Wilson DE. 2011. Biosafety in microbiological and biomedical laboratories. US Department of Health and Human Services, Washington, DC.Google Scholar21.Tellier R. 2006. Review of aerosol transmission of influenza A virus. Emerg Infect Dis 12:1657–1662.CrossrefPubMedGoogle Scholar22.Hinds WC. 1999. Aerosol Technology. properties, behavior, and measurement of airborne particles, 2nd ed. John Wiley Sons, Inc., Hoboken, NJ.Google Scholar23.Andersen BM, Syversen G, Thoresen H, Rasch M, Hochlin K, Seljordslia B, Snevold I, Berg E. 2010. Failure of dry mist of hydrogen peroxide 5% to kill Mycobacterium tuberculosis. J Hosp Infect 76:80–83.CrossrefPubMedGoogle Scholar24.Deegan RD, Bakajin O, Dupont TF, Huber G, Nagel SR, Witten TA. 1997. Capillary flow as the cause of ring stains from dried liquid drops. Nature 389:827–829.CrossrefGoogle Scholar25.Anonymous. 1994. Guideline of the Federal Health Office for testing the effectiveness of surface disinfectants for the disinfection of tuberculosis and commentary thereon. Bundesgesundheitsblatt 37:274–275.Google Scholar26.Gregory PH. 1961. The microbiology of the atmosphere. Leonard Hill (Books) Ltd., London, United Kingdom.CrossrefGoogle Scholar27.Falkinham JO. 2016. Nontuberculous mycobacteria: community and nosocomial waterborne opportunistic pathogens. Clin Microbiol Newsl 38:1–7.CrossrefGoogle Scholar28.Wood J. 2014. Evaluation of chlorine dioxide gas and peracetic acid fog for the decontamination of a mock heating, ventilation, and air conditioning duct system. Assessment and evaluation report (EPA 600-R-14–014). EPA, Washington, DC.Google Scholar29.Colwell RR. 1979. Enumeration of specific populations by the most-probable-number (MPN) method, p 56–61, In Costerton JW, Colwell RR (ed), Native aquatic bacteria: enumeration, activity, and ecology. American Society for Testing and Materials International, West Conshohocken, PA.CrossrefGoogle Scholar30.Kärber G. 1931. Contribution to the collective treatment of pharmacological series experiments. Naunyn Schmiedebergs Arch Pharmakol Exp Pathol 162:480–483.CrossrefGoogle Scholar31.Guizelini BP, Vandenberghe LPS, Sella SRBR, Soccol CR. 2012. Study of the influence of sporulation conditions on heat resistance of Geobacillus stearothermophilus used in the development of biological indicators for steam sterilization. Arch Microbiol 194:991–999.CrossrefPubMedGoogle Scholar32.Rogers JV, Choi YW, Richter WR, Rudnicki DC, Joseph DW, Sabourin CLK, Taylor ML, Chang JCS. 2007. Formaldehyde gas inactivation of Bacillus anthracis, Bacillus subtilis, and Geobacillus stearothermophilus spores on indoor surface materials. J Appl Microbiol 103:1104–1112.CrossrefPubMedGoogle Scholar33.Rogers JV, Sabourin CLK, Choi YW, Richter WR, Rudnicki DC, Riggs KB, Taylor ML, Chang J. 2005. Decontamination assessment of Bacillus anthracis, Bacillus subtilis, and Geobacillus stearothermophilus spores on indoor surfaces using a hydrogen peroxide gas generator. J Appl Microbiol 99:739–748.CrossrefPubMedGoogle Scholar34.Akbar Velayati A, Farnia P, Mozafari M, Malekshahian D, Seif S, Rahideh S, Mirsaeidi M. 2014. Molecular epidemiology of nontuberculous Mycobacteria Isolates from clinical and environmental sources of a metropolitan city. PLoS One 9:e114428.CrossrefPubMedGoogle Scholar35.Aghajani J, Rajaei E, Farnia P, Malekshahian D, Seif S. 2018. Mycobacterium farcinogenes and Mycobacterium senegalense as new environmental threats. Biomed Biotechnol Res J 2:184–190.CrossrefGoogle Scholar36.Chamoiseau G. 1979. Etiology of farcy in African bovines: nomenclature of the causal organisms Mycobacterium farcinogenes Chamoiseau and Mycobacterium senegalense (Chamoiseau) comb. nov. Int J Syst Evol Microbiol 29:407–410.CrossrefGoogle Scholar37.Griffiths PA. 1997. The resistance of Mycobacterium tuberculosis and other mycobacteria of increasing clinical importance to chemical agents. DSC:DXN011372. PhD thesis. Aston University, Birmingham, United Kingdom.Google Scholar38.Beswick AJ, Farrant J, Makison C, Gawn J, Frost G, Crook B, Pride J. 2011. Comparison of multiple systems for laboratory whole room fumigation. Appl Biosaf 16:139–157.CrossrefGoogle Scholar39.Carlin JB, Doyle LW. 2001. Basic concepts of statistical reasoning: hypothesis tests and the t-test. J Paediatr Child Health 37:72–77.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 87 • Number 3 • 15 January 2021eLocator: e02019-20Editor: Harold L. DrakeUniversity of BayreuthHistoryReceived: 18 August 2020Accepted: 5 November 2020Published online: 6 November 2020Copyright© 2021 Schinköthe et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license.PermissionsRequest permissions for this article.Request PermissionsKEYWORDSairborne disinfectiondry fogroom disinfectionperoxyacetic acidsevere acute respiratory syndrome coronavirus 2 (SARS-CoV-2)Indiana vesiculovirus (VSIV)murine norovirus (MNV)Mycobacterium senegalenseGeobacillus stearothermophilusBacillus subtilisContributorsAuthorsJan SchinkötheDepartment of Experimental Animal Facilities and Biorisk Management, Friedrich-Loeffler-Institut, Greifswald-Insel Riems, GermanyPresent address: Jan Schinköthe, Institute of Veterinary Pathology, Faculty of Veterinary Medicine, Leipzig University, Leipzig, Germany.View all articles by this authorHendrik A. Scheinemann https://orcid.org/0000-0002-4275-7672Department of Experimental Animal Facilities and Biorisk Management, Friedrich-Loeffler-Institut, Greifswald-Insel Riems, GermanyView all articles by this authorSandra DiederichInstitute of Novel and Emerging Infectious Diseases, Friedrich-Loeffler-Institut, Greifswald-Insel Riems, GermanyView all articles by this authorHolger FreeseDepartment of Experimental Animal Facilities and Biorisk Management, Friedrich-Loeffler-Institut, Greifswald-Insel Riems, GermanyView all articles by this authorMichael EschbaumerInstitute of Diagnostic Virology, Friedrich-Loeffler-Institut, Greifswald-Insel Riems, GermanyView all articles by this authorJens P. Teifke https://orcid.org/0000-0001-7168-4970Department of Experimental Animal Facilities and Biorisk Management, Friedrich-Loeffler-Institut, Greifswald-Insel Riems, GermanyView all articles by this authorSven Reiche https://orcid.org/0000-0001-7575-2483Department of Experimental Animal Facilities and Biorisk Management, Friedrich-Loeffler-Institut, Greifswald-Insel Riems, GermanyView all articles by this authorEditorHarold L. DrakeEditorUniversity of BayreuthNotesAddress correspondence to Sven Reiche, [email protected].Jan Schinköthe and Hendrik A. Scheinemann contributed equally to this work. Author order was determined on the basis of seniority in the project.Metrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleJuly 2021Articles of Significant Interest in This IssueArticles of Significant Interest in This IssueDOI: https://doi.org/10.1128/AEM.01182-21Volume 87, Number 1627 July 2021Information ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 87 • Number 16 • 27 July 2021eLocator: e01182-21HistoryPublished online: 27 July 2021Copyright© 2021 American Society for Microbiology. All Rights Reserved.PermissionsRequest permissions for this article.Request PermissionsContributorsMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShareApplied and Environmental MicrobiologyArticleMay 2021Saccharomyces cerevisiae Gene Expression during Fermentation of Pinot Noir Wines at an Industrially Relevant Scale Taylor Reiter, Rachel Montpetit, Shelby Byer, Isadora Frias, Esmeralda Leon, Robert Viano, Michael Mcloughlin, Thomas Halligan, Desmon Hernandez, Ron Runnebaumand Ben MontpetitSaccharomyces cerevisiae Gene Expression during Fermentation of Pinot Noir Wines at an Industrially Relevant ScaleAuthors: Taylor Reiter https://orcid.org/0000-0002-7388-421X, Rachel Montpetit, Shelby Byer, Isadora Frias, Esmeralda Leon, Robert Viano, Michael Mcloughlin, Thomas Halligan, Desmon Hernandez, Ron Runnebaum, and Ben Montpetit https://orcid.org/0000-0002-8317-983X [email protected]DOI: https://doi.org/10.1128/AEM.00036-21Volume 87, Number 1111 May 2021ABSTRACTREFERENCESABSTRACTSaccharomyces cerevisiae metabolism produces ethanol and other compounds during the fermentation of grape must into wine. Thousands of genes change expression over the course of a wine fermentation, allowing S. cerevisiae to adapt to and dominate the fermentation environment. Investigations into these gene expression patterns previously revealed genes that underlie cellular adaptation to the grape must and wine environments, involving metabolic specialization and ethanol tolerance. However, the majority of studies detailing gene expression patterns have occurred in controlled environments that may not recapitulate the biological and chemical complexity of fermentations performed at production scale. Here, an analysis of the S. cerevisiae RC212 gene expression program is presented, drawing from 40 pilot-scale fermentations (150 liters) using Pinot noir grapes from 10 California vineyards across two vintages. A core gene expression program was observed across all fermentations irrespective of vintage, similar to that of laboratory fermentations, in addition to novel gene expression patterns likely related to the presence of non-Saccharomyces microorganisms and oxygen availability during fermentation. These gene expression patterns, both common and diverse, provide insight into Saccharomyces cerevisiae biology critical to fermentation outcomes under industry-relevant conditions.IMPORTANCE This study characterized Saccharomyces cerevisiae RC212 gene expression during Pinot noir fermentation at pilot scale (150 liters) using industry-relevant conditions. The reported gene expression patterns of RC212 are generally similar to those observed under laboratory fermentation conditions but also contain gene expression signatures related to yeast-environment interactions found in a production setting (e.g., the presence of non-Saccharomyces microorganisms). Key genes and pathways highlighted by this work remain undercharacterized, indicating the need for further research to understand the roles of these genes and their impact on industrial wine fermentation outcomes.REFERENCES1.Querol A, Fernández-Espinar MT, Lí del Olmo M, Barrio E. 2003. Adaptive evolution of wine yeast. Int J Food Microbiol 86:3–10.CrossrefPubMedGoogle Scholar2.Rossignol T, Dulau L, Julien A, Blondin B. 2003. Genome-wide monitoring of wine yeast gene expression during alcoholic fermentation. Yeast 20:1369–1385.CrossrefPubMedGoogle Scholar3.Marks VD, Ho Sui SJ, Erasmus D, Van Der Merwe GK, Brumm J, Wasserman WW, Bryan J, Van Vuuren HJ. 2008. Dynamics of the yeast transcriptome during wine fermentation reveals a novel fermentation stress response. FEMS Yeast Res 8:35–52.CrossrefPubMedGoogle Scholar4.Rossouw D, Jolly N, Jacobson D, Bauer FF. 2012. The effect of scale on gene expression: commercial versus laboratory wine fermentations. Appl Microbiol Biotechnol 93:1207–1219.CrossrefPubMedGoogle Scholar5.Backhus LE, DeRisi J, Brown PO, Bisson LF. 2001. Functional genomic analysis of a commercial wine strain of Saccharomyces cerevisiae under differing nitrogen conditions. FEMS Yeast Res 1:111–125.CrossrefPubMedGoogle Scholar6.Mendes-Ferreira A, Del Olmo M, García-Martínez J, Jiménez-Martí E, Mendes-Faia A, Pérez-Ortín JE, Leao C. 2007. Transcriptional response of Saccharomyces cerevisiae to different nitrogen concentrations during alcoholic fermentation. Appl Environ Microbiol 73:3049–3060.CrossrefPubMedGoogle Scholar7.Gasch AP, Spellman PT, Kao CM, Carmel-Harel O, Eisen MB, Storz G, Botstein D, Brown PO. 2000. Genomic expression programs in the response of yeast cells to environmental changes. Mol Biol Cell 11:4241–4257.CrossrefPubMedGoogle Scholar8.Puig S, Pérez-Ortín JE. 2000. Stress response and expression patterns in wine fermentations of yeast genes induced at the diauxic shift. Yeast 16:139–148.CrossrefPubMedGoogle Scholar9.Bisson LF. 1999. Stuck and sluggish fermentations. Am J Enol Viticult 50:107–119.Google Scholar10.Gutiérrez A, Beltran G, Warringer J, Guillamón JM. 2013. Genetic basis of variations in nitrogen source utilization in four wine commercial yeast strains. PLoS One 8:e67166.CrossrefPubMedGoogle Scholar11.Patel S, Shibamoto T. 2002. Effect of different strains of Saccharomyces cerevisiae on production of volatiles in Napa Gamay wine and Petite Sirah wine. J Agric Food Chem 50:5649–5653.CrossrefPubMedGoogle Scholar12.Rossouw D, Naes T, Bauer FF. 2008. Linking gene regulation and the exo-metabolome: a comparative transcriptomics approach to identify genes that impact on the production of volatile aroma compounds in yeast. BMC Genomics 9:530.CrossrefPubMedGoogle Scholar13.Riou C, Nicaud J-M, Barre P, Gaillardin C. 1997. Stationary-phase gene expression in Saccharomyces cerevisiae during wine fermentation. Yeast 13:903–915.CrossrefPubMedGoogle Scholar14.Rossouw D, Bauer FF. 2009. Comparing the transcriptomes of wine yeast strains: toward understanding the interaction between environment and transcriptome during fermentation. Appl Microbiol Biotechnol 84:937–954.CrossrefPubMedGoogle Scholar15.Casalta E, Aguera E, Picou C, Rodriguez-Bencomo J-J, Salmon J-M, Sablayrolles J-M. 2010. A comparison of laboratory and pilot-scale fermentations in winemaking conditions. Appl Microbiol Biotechnol 87:1665–1673.CrossrefPubMedGoogle Scholar16.Walker ME, Nguyen TD, Liccioli T, Schmid F, Kalatzis N, Sundstrom JF, Gardner JM, Jiranek V. 2014. Genome-wide identification of the fermentome; genes required for successful and timely completion of wine-like fermentation by Saccharomyces cerevisiae. BMC Genomics 15:552.CrossrefPubMedGoogle Scholar17.Cadière A, Aguera E, Caillé S, Ortiz-Julien A, Dequin S. 2012. Pilot-scale evaluation the enological traits of a novel, aromatic wine yeast strain obtained by adaptive evolution. Food Microbiol 32:332–337.CrossrefPubMedGoogle Scholar18.Vila I. 1998. Les levures aromatiques en vinification: évaluation de ce caractère par l’analyse sensorielle et l’analyse chimique: déterminisme biochimique des facteurs responsables. PhD thesis. Montpellier 2 University, Montpellier, France.Google Scholar19.Du Toit W, Marais J, Pretorius I, Du Toit M. 2006. Oxygen in must and wine: a review. S Afr J Enol Viticult 27:76–94.CrossrefGoogle Scholar20.Orellana M, Aceituno FF, Slater AW, Almonacid LI, Melo F, Agosin E. 2014. Metabolic and transcriptomic response of the wine yeast Saccharomyces cerevisiae strain EC1118 after an oxygen impulse under carbon-sufficient, nitrogen-limited fermentative conditions. FEMS Yeast Res 14:412–424.CrossrefPubMedGoogle Scholar21.Varela C, Cárdenas J, Melo F, Agosin E. 2005. Quantitative analysis of wine yeast gene expression profiles under winemaking conditions. Yeast 22:369–383.CrossrefPubMedGoogle Scholar22.Roullier-Gall C, Boutegrabet L, Gougeon RD, Schmitt-Kopplin P. 2014. A grape and wine chemodiversity comparison of different appellations in Burgundy: vintage vs terroir effects. Food Chem 152:100–107.CrossrefPubMedGoogle Scholar23.Vilanova M, Rodríguez I, Canosa P, Otero I, Gamero E, Moreno D, Talaverano I, Valdés E. 2015. Variability in chemical composition of Vitis vinifera cv Mencía from different geographic areas and vintages in Ribeira Sacra (NW Spain). Food Chem 169:187–196.CrossrefPubMedGoogle Scholar24.Ramos MC, Martínez de Toda F. 2019. Variability of Tempranillo grape composition in the Rioja DOCa (Spain) related to soil and climatic characteristics. J Sci Food Agric 99:1153–1165.CrossrefPubMedGoogle Scholar25.Grainger C, Yeh A, Byer S, Hjelmeland A, Lima MM, Runnebaum RC. 2021. Vineyard site impact on the elemental composition of Pinot noir wines. Food Chem 334:127386.CrossrefPubMedGoogle Scholar26.Bokulich NA, Thorngate JH, Richardson PM, Mills DA. 2014. Microbial biogeography of wine grapes is conditioned by cultivar, vintage, and climate. Proc Natl Acad Sci U S A 111:E139–E148.CrossrefPubMedGoogle Scholar27.Bokulich NA, Collins TS, Masarweh C, Allen G, Heymann H, Ebeler SE, Mills DA. 2016. Associations among wine grape microbiome, metabolome, and fermentation behavior suggest microbial contribution to regional wine characteristics. mBio 7:e00631-16.CrossrefPubMedGoogle Scholar28.David V, Terrat S, Herzine K, Claisse O, Rousseaux S, Tourdot-Maréchal R, Masneuf-Pomarede I, Ranjard L, Alexandre H. 2014. High-throughput sequencing of amplicons for monitoring yeast biodiversity in must and during alcoholic fermentation. J Ind Microbiol Biotechnol 41:811–821.CrossrefPubMedGoogle Scholar29.Pinto C, Pinho D, Cardoso R, Custódio V, Fernandes J, Sousa S, Pinheiro M, Egas C, Gomes AC. 2015. Wine fermentation microbiome: a landscape from different Portuguese wine appellations. Front Microbiol 6:905.CrossrefPubMedGoogle Scholar30.Wang C, García-Fernández D, Mas A, Esteve-Zarzoso B. 2015. Fungal diversity in grape must and wine fermentation assessed by massive sequencing, quantitative PCR and DGGE. Front Microbiol 6:1156.CrossrefPubMedGoogle Scholar31.Garofalo C, Russo P, Beneduce L, Massa S, Spano G, Capozzi V. 2016. Non-Saccharomyces biodiversity in wine and the \"microbial terroir”: a survey on Nero di Troia wine from the Apulian region, Italy. Ann Microbiol 66:143–150.CrossrefGoogle Scholar32.Mezzasalma V, Sandionigi A, Bruni I, Bruno A, Lovicu G, Casiraghi M, Labra M. 2017. Grape microbiome as a reliable and persistent signature of field origin and environmental conditions in Cannonau wine production. PLoS One 12:e0184615.CrossrefPubMedGoogle Scholar33.Mezzasalma V, Sandionigi A, Guzzetti L, Galimberti A, Grando MS, Tardaguila J, Labra M. 2018. Geographical and cultivar features differentiate grape microbiota in northern Italy and Spain vineyards. Front Microbiol 9:946.CrossrefPubMedGoogle Scholar34.Singh P, Santoni S, This P, Péros J-P. 2018. Genotype-environment interaction shapes the microbial assemblage in grapevine’s phyllosphere and carposphere: an NGS approach. Microorganisms 6:96.CrossrefGoogle Scholar35.Liu D, Chen Q, Zhang P, Chen D, Howell KS. 2020. The fungal microbiome is an important component of vineyard ecosystems and correlates with regional distinctiveness of wine. mSphere 5:e00534-20.CrossrefPubMedGoogle Scholar36.Jolly N, Augustyn O, Pretorius I. 2003. The occurrence of non-Saccharomyces cerevisiae yeast species over three vintages in four vineyards and grape musts from four production regions of the Western Cape, South Africa. S Afr J Enol Viticult 24:35–42.CrossrefGoogle Scholar37.Ghosh S, Bagheri B, Morgan HH, Divol B, Setati ME. 2015. Assessment of wine microbial diversity using ARISA and cultivation-based methods. Ann Microbiol 65:1833–1840.CrossrefGoogle Scholar38.Wang C, Esteve-Zarzoso B, Cocolin L, Mas A, Rantsiou K. 2015. Viable and culturable populations of Saccharomyces cerevisiae, Hanseniaspora uvarum and Starmerella bacillaris (synonym Candida zemplinina) during Barbera must fermentation. Food Res Int 78:195–200.CrossrefPubMedGoogle Scholar39.Bagheri B, Bauer F, Setati M. 2016. The diversity and dynamics of indigenous yeast communities in grape must from vineyards employing different agronomic practices and their influence on wine fermentation. S Afr J Enol Viticult 36:243–251.CrossrefGoogle Scholar40.Bagheri B, Bauer FF, Cardinali G, Setati ME. 2020. Ecological interactions are a primary driver of population dynamics in wine yeast microbiota during fermentation. Sci Rep 10: 4911.CrossrefPubMedGoogle Scholar41.Brou P, Taillandier P, Beaufort S, Brandam C. 2018. Mixed culture fermentation using Torulaspora delbrueckii and Saccharomyces cerevisiae with direct and indirect contact: impact of anaerobic growth factors. Eur Food Res Technol 244:1699–1710.CrossrefGoogle Scholar42.Curiel JA, Morales P, Gonzalez R, Tronchoni J. 2017. Different non-Saccharomyces yeast species stimulate nutrient consumption in S. cerevisiae mixed cultures. Front Microbiol 8:2121.CrossrefPubMedGoogle Scholar43.Alonso-del-Real J, Pérez-Torrado R, Querol A, Barrio E. 2019. Dominance of wine Saccharomyces cerevisiae strains over S. kudriavzevii in industrial fermentation competitions is related to an acceleration of nutrient uptake and utilization. Environ Microbiol 21:1627–1644.CrossrefPubMedGoogle Scholar44.Tronchoni J, Curiel JA, Morales P, Torres-Pérez R, Gonzalez R. 2017. Early transcriptional response to biotic stress in mixed starter fermentations involving Saccharomyces cerevisiae and Torulaspora delbrueckii. Int J Food Microbiol 241:60–68.CrossrefPubMedGoogle Scholar45.Shekhawat K, Patterton H, Bauer FF, Setati ME. 2019. RNA-seq based transcriptional analysis of Saccharomyces cerevisiae and Lachancea thermotolerans in mixed-culture fermentations under anaerobic conditions. BMC Genomics 20:145.CrossrefPubMedGoogle Scholar46.Conacher CG, Rossouw D, Bauer F. 2019. Peer pressure: evolutionary responses to biotic pressures in wine yeasts. FEMS Yeast Res 19:foz072.CrossrefPubMedGoogle Scholar47.Bordet F, Joran A, Klein G, Roullier-Gall C, Alexandre H. 2020. Yeast-yeast interactions: mechanisms, methodologies and impact on composition. Microorganisms 8:600.CrossrefGoogle Scholar48.Egli C, Edinger W, Mitrakul C, Henick-Kling T. 1998. Dynamics of indigenous and inoculated yeast populations and their effect on the sensory character of Riesling and Chardonnay wines. J Appl Microbiol 85:779–789.CrossrefPubMedGoogle Scholar49.Bartowsky EJ. 2009. Bacterial spoilage of wine and approaches to minimize it. Lett Appl Microbiol 48:149–156.CrossrefPubMedGoogle Scholar50.Cantu A, Lafontaine S, Frias I, Sokolowsky M, Yeh A, Lestringant P, Hjelmeland A, Byer S, Heymann H, Runnebaum RC. 2021. Investigating the impact of regionality on the sensorial and chemical aging characteristics of Pinot noir grown throughout the US West coast. Food Chem 337:127720.CrossrefPubMedGoogle Scholar51.Bisson LF. 2019. Gene expression in yeasts during wine fermentation, p 165–209. In Romano P, Ciani M, Fleet GH (ed), Yeasts in the production of wine. Springer, New York, NY.CrossrefGoogle Scholar52.Rodicio R, Heinisch JJ. 2017. Carbohydrate metabolism in wine yeasts, p 189–213. In König H, Unden G, Fröhlich J (ed), Biology of microorganisms on grapes, in must and in wine, 2nd ed. Springer International Publishing, Cham, Switzerland.CrossrefGoogle Scholar53.Ozcan S, Johnston M. 1996. Two different repressors collaborate to restrict expression of the yeast glucose transporter genes HXT2 and HXT4 to low levels of glucose. Mol Cell Biol 16:5536–5545.CrossrefPubMedGoogle Scholar54.Ogawa N, DeRisi J, Brown PO. 2000. New components of a system for phosphate accumulation and polyphosphate metabolism in Saccharomyces cerevisiae revealed by genomic expression analysis. Mol Biol Cell 11:4309–4321.CrossrefPubMedGoogle Scholar55.Ljungdahl PO, Daignan-Fornier B. 2012. Regulation of amino acid, nucleotide, and phosphate metabolism in Saccharomyces cerevisiae. Genetics 190:885–929.CrossrefPubMedGoogle Scholar56.Barbosa C, Mendes-Faia A, Lage P, Mira NP, Mendes-Ferreira A. 2015. Genomic expression program of Saccharomyces cerevisiae along a mixed-culture wine fermentation with Hanseniaspora guilliermondii. Microb Cell Fact 14:124.CrossrefPubMedGoogle Scholar57.Kosel J, Čadež N, Schuller D, Carreto L, Franco-Duarte R, Raspor P. 2017. The influence of Dekkera bruxellensis on the transcriptome of Saccharomyces cerevisiae and on the aromatic profile of synthetic wine must. FEMS Yeast Res 17:fox018.CrossrefGoogle Scholar58.Reiter T, Montpetit R, Byer S, Frias I, Leon E, Viano R, Mcloughlin M, Halligan T, Hernandez D, Figueroa-Balderas R, Cantu D, Steenwerth K, Runnebaum R, Montpetit B. 2021. Transcriptomics provides a genetic signature of vineyard site and offers insight into vintage-independent inoculated fermentation outcomes. mSystems 6:e00033-21.CrossrefPubMedGoogle Scholar59.Devatine A, Chiciuc I, Mietton-Peuchot M. 2011. The protective role of dissolved carbon dioxide against wine oxidation: a simple and rational approach. OENO One 45:189–197.CrossrefGoogle Scholar60.Moenne MI, Saa P, Laurie VF, Pérez-Correa JR, Agosin E. 2014. Oxygen incorporation and dissolution during industrial-scale red wine fermentations. Food Bioprocess Technol 7:2627–2636.CrossrefGoogle Scholar61.Abramova N, Sertil O, Mehta S, Lowry CV. 2001. Reciprocal regulation of anaerobic and aerobic cell wall mannoprotein gene expression in Saccharomyces cerevisiae. J Bacteriol 183:2881–2887.CrossrefPubMedGoogle Scholar62.Sertil O, Kapoor R, Cohen BD, Abramova N, Lowry CV. 2003. Synergistic repression of anaerobic genes by Mot3 and Rox1 in Saccharomyces cerevisiae. Nucleic Acids Res 31:5831–5837.CrossrefPubMedGoogle Scholar63.Kwast KE, Burke PV, Poyton RO. 1998. Oxygen sensing and the transcriptional regulation of oxygen-responsive genes in yeast. J Exp Biol 201:1177–1195.CrossrefPubMedGoogle Scholar64.Wöhl T, Klier H, Ammer H, Lottspeich F, Magdolen V. 1993. The HYP2 gene of Saccharomyces cerevisiae is essential for aerobic growth: characterization of different isoforms of the hypusine-containing protein Hyp2p and analysis of gene disruption mutants. Mol Gen Genet 241:305–311.CrossrefPubMedGoogle Scholar65.Burke PV, Poyton RO. 1998. Structure/function of oxygen-regulated isoforms in cytochrome c oxidase. J Exp Biol 201:1163–1175.CrossrefPubMedGoogle Scholar66.Remize F, Cambon B, Barnavon L, Dequin S. 2003. Glycerol formation during wine fermentation is mainly linked to Gpd1p and is only partially controlled by the HOG pathway. Yeast 20:1243–1253.CrossrefPubMedGoogle Scholar67.Jung J-Y, Kim T-Y, Ng C-Y, Oh M-K. 2012. Characterization of GCY1 in Saccharomyces cerevisiae by metabolic profiling. J Appl Microbiol 113:1468–1478.CrossrefPubMedGoogle Scholar68.Kliewer WM. 1970. Free amino acids and other nitrogenous fractions in wine grapes. J Food Sci 35:17–21.CrossrefGoogle Scholar69.ter Schure EG, van Riel NAW, Verrips CT. 2000. The role of ammonia metabolism in nitrogen catabolite repression in Saccharomyces cerevisiae. FEMS Microbiol Rev 24:67–83.CrossrefPubMedGoogle Scholar70.Huang HL, Brandriss MC. 2000. The regulator of the yeast proline utilization pathway is differentially phosphorylated in response to the quality of the nitrogen source. Mol Cell Biol 20:892–899.CrossrefPubMedGoogle Scholar71.Takagi H, Taguchi J, Kaino T. 2016. Proline accumulation protects Saccharomyces cerevisiae cells in stationary phase from ethanol stress by reducing reactive oxygen species levels. Yeast 33:355–363.CrossrefPubMedGoogle Scholar72.Rosenfeld E, Beauvoit B, Blondin B, Salmon J-M. 2003. Oxygen consumption by anaerobic Saccharomyces cerevisiae under enological conditions: effect on fermentation kinetics. Appl Environ Microbiol 69:113–121.CrossrefPubMedGoogle Scholar73.Tarko T, Duda-Chodak A, Sroka P, Siuta M. 2020. The impact of oxygen at various stages of vinification on the chemical composition and the antioxidant and sensory properties of white and red wines. Int J Food Sci 2020:7902974.CrossrefPubMedGoogle Scholar74.O\'Connor-Cox ESC, Lodolo EJ, Axcell BC. 1996. Mitochondrial relevance to yeast fermentative performance: a review. J Inst Brewing 102:19–25.CrossrefGoogle Scholar75.Kitagaki H, Takagi H. 2014. Mitochondrial metabolism and stress response of yeast: applications in fermentation technologies. J Biosci Bioeng 117:383–393.CrossrefPubMedGoogle Scholar76.Zara G, van Vuuren HJ, Mannazzu I, Zara S, Budroni M. 2019. Transcriptomic response of Saccharomyces cerevisiae during fermentation under oleic acid and ergosterol depletion. Fermentation 5:57.CrossrefGoogle Scholar77.Zeng J, Smith KE, Chong P. 1993. Effects of alcohol-induced lipid interdigitation on proton permeability in l-α-dipalmitoylphosphatidylcholine vesicles. Biophys J 65:1404–1414.CrossrefPubMedGoogle Scholar78.Landolfo S, Politi H, Angelozzi D, Mannazzu I. 2008. ROS accumulation and oxidative damage to cell structures in Saccharomyces cerevisiae wine strains during fermentation of high-sugar-containing medium. Biochim Biophys Acta 1780:892–898.CrossrefPubMedGoogle Scholar79.Molenaar D, Van Berlo R, De Ridder D, Teusink B. 2009. Shifts in growth strategies reflect tradeoffs in cellular economics. Mol Syst Biol 5:323.CrossrefPubMedGoogle Scholar80.Wimpenny J. 1969. The effect of Eh on regulatory processes in facultative anaerobes. Biotechnol Bioeng 11:623–629.CrossrefPubMedGoogle Scholar81.Somlo M, Fukuhara H. 1965. On the necessity of molecular oxygen for the synthesis of respiratory enzymes in yeast. Biochem Biophys Res Commun 19:587–591.CrossrefPubMedGoogle Scholar82.Pammer M, Briza P, Ellinger A, Schuster T, Stucka R, Feldmann H, Breitenbach M. 1992. DIT101 (CSD2, CAL1), a cell cycle-regulated yeast gene required for synthesis of chitin in cell walls and chitosan in spore walls. Yeast 8:1089–1099.CrossrefPubMedGoogle Scholar83.Argüello-Miranda O, Liu Y, Wood NE, Kositangool P, Doncic A. 2018. Integration of multiple metabolic signals determines cell fate prior to commitment. Mol Cell 71:733–744.CrossrefPubMedGoogle Scholar84.Zhao H, Wang Q, Liu C, Shang Y, Wen F, Wang F, Liu W, Xiao W, Li W. 2018. A role for the respiratory chain in regulating meiosis initiation in Saccharomyces cerevisiae. Genetics 208:1181–1194.CrossrefPubMedGoogle Scholar85.Sipiczki M. 2011. Diversity, variability and fast adaptive evolution of the wine yeast (Saccharomyces cerevisiae) genome: a review. Ann Microbiol 61:85–93.CrossrefGoogle Scholar86.Reiner S, Micolod D, Zellnig G, Schneiter R. 2006. A genomewide screen reveals a role of mitochondria in anaerobic uptake of sterols in yeast. Mol Biol Cell 17:90–103.CrossrefPubMedGoogle Scholar87.Perez-Gallardo RV, Briones LS, Díaz-Pérez AL, Gutiérrez S, Rodríguez-Zavala JS, Campos-García J. 2013. Reactive oxygen species production induced by ethanol in Saccharomyces cerevisiae increases because of a dysfunctional mitochondrial iron-sulfur cluster assembly system. FEMS Yeast Res 13:804–819.CrossrefPubMedGoogle Scholar88.Carmel-Harel O, Storz G. 2000. Roles of the glutathione- and thioredoxin-dependent reduction systems in the Escherichia coli and Saccharomyces cerevisiae responses to oxidative stress. Annu Rev Microbiol 54:439–461.CrossrefPubMedGoogle Scholar89.Toledano MB, Delaunay-Moisan A, Outten CE, Igbaria A. 2013. Functions and cellular compartmentation of the thioredoxin and glutathione pathways in yeast. Antioxid Redox Signal 18:1699–1711.CrossrefPubMedGoogle Scholar90.Bridi R, González A, Bordeu E, López-Alarcón C, Aspée A, Diethelm B, Lissi E, Parpinello GP, Versari A. 2015. Monitoring peroxides generation during model wine fermentation by FOX-1 assay. Food Chem 175:25–28.CrossrefPubMedGoogle Scholar91.Maslanka R, Zadrag-Tecza R, Kwolek-Mirek M. 2020. Linkage between carbon metabolism, redox status and cellular physiology in the yeast Saccharomyces cerevisiae devoid of SOD1 or SOD2 gene. Genes 11:780.CrossrefGoogle Scholar92.Matsufuji Y, Yamamoto K, Yamauchi K, Mitsunaga T, Hayakawa T, Nakagawa T. 2013. Novel physiological roles for glutathione in sequestering acetaldehyde to confer acetaldehyde tolerance in Saccharomyces cerevisiae. Appl Microbiol Biotechnol 97:297–303.CrossrefPubMedGoogle Scholar93.Koc A, Mathews CK, Wheeler LJ, Gross MK, Merrill GF. 2006. Thioredoxin is required for deoxyribonucleotide pool maintenance during S phase. J Biol Chem 281:15058–15063.CrossrefPubMedGoogle Scholar94.Muller E. 1991. Thioredoxin deficiency in yeast prolongs S phase and shortens the G1 interval of the cell cycle. J Biol Chem 266:9194–9202.CrossrefPubMedGoogle Scholar95.Shenton D, Grant CM. 2003. Protein S-thiolation targets glycolysis and protein synthesis in response to oxidative stress in the yeast Saccharomyces cerevisiae. Biochem J 374:513–519.CrossrefPubMedGoogle Scholar96.Ralser M, Wamelink MM, Kowald A, Gerisch B, Heeren G, Struys EA, Klipp E, Jakobs C, Breitenbach M, Lehrach H, Krobitsch S. 2007. Dynamic rerouting of the carbohydrate flux is key to counteracting oxidative stress. J Biol 6:10.CrossrefPubMedGoogle Scholar97.Deponte M. 2017. The incomplete glutathione puzzle: just guessing at numbers and figures? Antioxid Redox Signal 27:1130–1161.CrossrefPubMedGoogle Scholar98.Saint-Prix F, Bönquist L, Dequin S. 2004. Functional analysis of the ALD gene family of Saccharomyces cerevisiae during anaerobic growth on glucose: the NADP+-dependent Ald6p and Ald5p isoforms play a major role in acetate formation. Microbiology (Reading) 150:2209–2220.CrossrefPubMedGoogle Scholar99.Greetham D, Vickerstaff J, Shenton D, Perrone GG, Dawes IW, Grant CM. 2010. Thioredoxins function as deglutathionylase enzymes in the yeast Saccharomyces cerevisiae. BMC Biochem 11:3.CrossrefPubMedGoogle Scholar100.Collinson EJ, Wheeler GL, Garrido EO, Avery AM, Avery SV, Grant CM. 2002. The yeast glutaredoxins are active as glutathione peroxidases. J Biol Chem 277:16712–16717.CrossrefPubMedGoogle Scholar101.Ohdate T, Kita K, Inoue Y. 2010. Kinetics and redox regulation of Gpx1, an atypical 2-Cys peroxiredoxin, in Saccharomyces cerevisiae. FEMS Yeast Res 10:787–790.CrossrefPubMedGoogle Scholar102.Ohdate T, Izawa S, Kita K, Inoue Y. 2010. Regulatory mechanism for expression of GPX1 in response to glucose starvation and Ca2+ in Saccharomyces cerevisiae: involvement of Snf1 and Ras/cAMP pathway in Ca2+ signaling. Genes Cells 15:59–75.CrossrefPubMedGoogle Scholar103.Greetham D, Grant CM. 2009. Antioxidant activity of the yeast mitochondrial one-Cys peroxiredoxin is dependent on thioredoxin reductase and glutathione in vivo. Mol Cell Biol 29:3229–3240.CrossrefPubMedGoogle Scholar104.Martins AMT, Cordeiro CAA, Freire AMJP. 2001. In situ analysis of methylglyoxal metabolism in Saccharomyces cerevisiae. FEBS Lett 499:41–44.CrossrefPubMedGoogle Scholar105.Dhaoui M, Auchère F, Blaiseau P-L, Lesuisse E, Landoulsi A, Camadro J-M, Haguenauer-Tsapis R, Belgareh-Touzé N. 2011. Gex1 is a yeast glutathione exchanger that interferes with pH and redox homeostasis. Mol Biol Cell 22:2054–2067.CrossrefPubMedGoogle Scholar106.Cordente AG, Capone DL, Curtin CD. 2015. Unravelling glutathione conjugate catabolism in Saccharomyces cerevisiae: the role of glutathione/dipeptide transporters and vacuolar function in the release of volatile sulfur compounds 3-mercaptohexan-1-ol and 4-mercapto-4-methylpentan-2-one. Appl Microbiol Biotechnol 99:9709–9722.CrossrefPubMedGoogle Scholar107.Kumar A, Tikoo S, Maity S, Sengupta S, Sengupta S, Kaur A, Kumar Bachhawat A. 2012. Mammalian proapoptotic factor ChaC1 and its homologues function as γ-glutamyl cyclotransferases acting specifically on glutathione. EMBO Rep 13:1095–1101.CrossrefPubMedGoogle Scholar108.Chen X, Li S, Liu L. 2014. Engineering redox balance through cofactor systems. Trends Biotechnol 32:337–343.CrossrefPubMedGoogle Scholar109.Fariña L, Medina K, Urruty M, Boido E, Dellacassa E, Carrau F. 2012. Redox effect on volatile compound formation in wine during fermentation by Saccharomyces cerevisiae. Food Chem 134:933–939.CrossrefPubMedGoogle Scholar110.Bloem A, Sanchez I, Dequin S, Camarasa C. 2016. Metabolic impact of redox cofactor perturbations on the formation of aroma compounds in Saccharomyces cerevisiae. Appl Environ Microbiol 82:174–183.CrossrefPubMedGoogle Scholar111.Xu X, Bao M, Niu C, Wang J, Liu C, Zheng F, Li Y, Li Q. 2019. Engineering the cytosolic NADH availability in lager yeast to improve the aroma profile of beer. Biotechnol Lett 41:363–369.CrossrefPubMedGoogle Scholar112.Xu X, Song Y, Guo L, Cheng W, Niu C, Wang J, Liu C, Zheng F, Zhou Y, Li X, Mu Y, Li Q. 2020. Higher NADH availability of lager yeast increases the flavor stability of beer. J Agric Food Chem 68:584–590.CrossrefPubMedGoogle Scholar113.Causton HC, Ren B, Koh SS, Harbison CT, Kanin E, Jennings EG, Lee TI, True HL, Lander ES, Young RA. 2001. Remodeling of yeast genome expression in response to environmental changes. Mol Biol Cell 12:323–337.CrossrefPubMedGoogle Scholar114.Mbuyane LL, de Kock M, Bauer FF, Divol B. 2018. Torulaspora delbrueckii produces high levels of C5 and C6 polyols during wine fermentations. FEMS Yeast Res 18:foy084.CrossrefGoogle Scholar115.Endo A, Futagawa-Endo Y, Dicks LM. 2009. Isolation and characterization of fructophilic lactic acid bacteria from fructose-rich niches. Syst Appl Microbiol 32:593–600.CrossrefPubMedGoogle Scholar116.Du Toit M, Pretorius IS. 2000. Microbial spoilage and preservation of wine: using weapons for nature’s own arsenal. S Afr J Enol Viticult 21:74–96.CrossrefGoogle Scholar117.Quain DE, Boulton CA. 1987. Growth and metabolism of mannitol by strains of Saccharomyces cerevisiae. J Gen Microbiol 133:1675–1684.CrossrefPubMedGoogle Scholar118.Jordan P, Choe J-Y, Boles E, Oreb M. 2016. Hxt13, Hxt15, Hxt16 and Hxt17 from Saccharomyces cerevisiae represent a novel type of polyol transporters. Sci Rep 6:23502.CrossrefPubMedGoogle Scholar119.Ramakrishnan V, Walker GA, Fan Q, Ogawa M, Luo Y, Luong P, Joseph C, Bisson LF. 2016. Inter-kingdom modification of metabolic behavior: [GAR+] prion induction in Saccharomyces cerevisiae mediated by wine ecosystem bacteria. Front Ecol Evol 4:137.CrossrefGoogle Scholar120.Brown JC, Lindquist S. 2009. A heritable switch in carbon source utilization driven by an unusual yeast prion. Genes Dev 23:2320–2332.CrossrefPubMedGoogle Scholar121.Jarosz DF, Brown JCS, Walker GA, Datta MS, Ung WL, Lancaster AK, Rotem A, Chang A, Newby GA, Weitz DA, Bisson LF, Lindquist S. 2014. Cross-kingdom chemical communication drives a heritable, mutually beneficial prion-based transformation of metabolism. Cell 158:1083–1093.CrossrefPubMedGoogle Scholar122.Nurgel C, Pickering G. 2005. Contribution of glycerol, ethanol and sugar to the perception of viscosity and density elicited by model white wines. J Texture Stud 36:303–323.CrossrefGoogle Scholar123.Ansell R, Granath K, Hohmann S, Thevelein JM, Adler L. 1997. The two isoenzymes for yeast NAD+-dependent glycerol 3-phosphate dehydrogenase encoded by GPD1 and GPD2 have distinct roles in osmoadaptation and redox regulation. EMBO J 16:2179–2187.CrossrefPubMedGoogle Scholar124.Remize F, Barnavon L, Dequin S. 2001. Glycerol export and glycerol-3-phosphate dehydrogenase, but not glycerol phosphatase, are rate limiting for glycerol production in Saccharomyces cerevisiae. Metab Eng 3:301–312.CrossrefPubMedGoogle Scholar125.Smith KD, Gordon PB, Rivetta A, Allen KE, Berbasova T, Slayman C, Strobel SA. 2015. Yeast Fex1p is a constitutively expressed fluoride channel with functional asymmetry of its two homologous domains. J Biol Chem 290:19874–19887.CrossrefPubMedGoogle Scholar126.Clayton MG. 1997. Fluoride inhibition of wine yeasts: a thesis presented in partial fulfilment of the requirements for the degree of Master of Science in Microbiology at Massey University. Master’s thesis. Massey University, Palmerston North, New Zealand.Google Scholar127.Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, Gingeras TR. 2013. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29:15–21.CrossrefPubMedGoogle Scholar128.Smith T, Heger A, Sudbery I. 2017. UMI-tools: modeling sequencing errors in Unique Molecular Identifiers to improve quantification accuracy. Genome Res 27:491–499.CrossrefPubMedGoogle Scholar129.Anders S, Pyl PT, Huber W. 2015. HTSeq: a Python framework to work with high-throughput sequencing data. Bioinformatics 31:166–169.CrossrefPubMedGoogle Scholar130.Robinson MD, McCarthy DJ, Smyth GK. 2010. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26:139–140.CrossrefPubMedGoogle Scholar131.Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W, Smyth GK. 2015. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res 43:e47.CrossrefPubMedGoogle Scholar132.Liebermeister W, Noor E, Flamholz A, Davidi D, Bernhardt J, Milo R. 2014. Visual account of protein investment in cellular functions. Proc Natl Acad Sci U S A 111:8488–8493.CrossrefPubMedGoogle Scholar133.Yu G, Wang L-G, Han Y, He Q-Y. 2012. clusterProfiler: an R package for comparing biological themes among gene clusters. Omics 16:284–287.CrossrefPubMedGoogle Scholar134.Rorbach J, Bobrowicz A, Pearce S, Minczuk M. 2014. Polyadenylation in bacteria and organelles. Methods Mol Biol 1125:211–227.CrossrefPubMedGoogle Scholar135.Brown CT, Irber L. 2016. sourmash: a library for MinHash sketching of DNA. J Open Source Software 1:27.CrossrefGoogle Scholar136.Pierce NT, Irber L, Reiter T, Brooks P, Brown CT. 2019. Large-scale sequence comparisons with sourmash. F1000Res 8:1006.CrossrefPubMedGoogle Scholar137.Li H. 2013. Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. arXiv 1303.3997.Google Scholar138.Ciriacy M. 1975. Genetics of alcohol dehydrogenase in Saccharomyces cerevisiae. Mol Gen Genet 138:157–164.CrossrefPubMedGoogle Scholar139.Zuzuarregui A, Monteoliva L, Gil C, del Olmo M.l. 2006. Transcriptomic and proteomic approach for understanding the molecular basis of adaptation of Saccharomyces cerevisiae to wine fermentation. Appl Environ Microbiol 72:836–847.CrossrefPubMedGoogle Scholar140.Curiel JA, Salvadó Z, Tronchoni J, Morales P, Rodrigues AJ, Quirós M, Gonzalez R. 2016. Identification of target genes to control acetate yield during aerobic fermentation with Saccharomyces cerevisiae. Microb Cell Fact 15:156.CrossrefPubMedGoogle Scholar141.Piper P, Mahé Y, Thompson S, Pandjaitan R, Holyoak C, Egner R, Mühlbauer M, Coote P, Kuchler K. 1998. The Pdr12 ABC transporter is required for the development of weak organic acid resistance in yeast. EMBO J 17:4257–4265.CrossrefPubMedGoogle Scholar142.Caro LHP, Smits GJ, van Egmond P, Chapman JW, Klis FM. 1998. Transcription of multiple cell wall protein-encoding genes in Saccharomyces cerevisiae is differentially regulated during the cell cycle. FEMS Microbiol Lett 161:345–349.CrossrefPubMedGoogle Scholar143.Crépin L, Truong NM, Bloem A, Sanchez I, Dequin S, Camarasa C. 2017. Management of multiple nitrogen sources during wine fermentation by Saccharomyces cerevisiae. Appl Environ Microbiol 83:e02617-16.CrossrefPubMedGoogle Scholar144.Hazelwood LA, Daran J-M, Van Maris AJ, Pronk JT, Dickinson JR. 2008. The Ehrlich pathway for fusel alcohol production: a century of research on Saccharomyces cerevisiae metabolism. Appl Environ Microbiol 74:2259–2266.CrossrefPubMedGoogle Scholar145.Cebollero E, Reggiori F. 2009. Regulation of autophagy in yeast Saccharomyces cerevisiae. Biochim Biophys Acta 1793:1413–1421.CrossrefPubMedGoogle Scholar146.Alexandre H, Ansanay-Galeote V, Dequin S, Blondin B. 2001. Global gene expression during short-term ethanol stress in Saccharomyces cerevisiae. FEBS Lett 498:98–103.CrossrefPubMedGoogle Scholar147.Lillie SH, Pringle JR. 1980. Reserve carbohydrate metabolism in Saccharomyces cerevisiae: responses to nutrient limitation. J Bacteriol 143:1384–1394.CrossrefPubMedGoogle Scholar148.François J, Parrou JL. 2001. Reserve carbohydrates metabolism in the yeast Saccharomyces cerevisiae. FEMS Microbiol Rev 25:125–145.CrossrefPubMedGoogle Scholar149.Singer MA, Lindquist S. 1998. Multiple effects of trehalose on protein folding in vitro and in vivo. Mol Cell 1:639–648.CrossrefPubMedGoogle Scholar150.Parrou JL, Teste M-A, Francois J. 1997. Effects of various types of stress on the metabolism of reserve carbohydrates in Saccharomyces cerevisiae: genetic evidence for a stress-induced recycling of glycogen and trehalose. Microbiology 143:1891–1900.CrossrefPubMedGoogle Scholar151.Udom N, Chansongkrow P, Charoensawan V, Auesukaree C. 2019. Coordination of the cell wall integrity and high-osmolarity glycerol pathways in response to ethanol stress in Saccharomyces cerevisiae. Appl Environ Microbiol 85:e00551-19.CrossrefPubMedGoogle ScholarInformation ContributorsInformationPublished InApplied and Environmental MicrobiologyVolume 87 • Number 11 • 11 May 2021eLocator: e00036-21Editor: Edward G. DudleyThe Pennsylvania State UniversityHistoryReceived: 8 January 2021Accepted: 15 March 2021Published online: 19 March 2021Copyright© 2021 Reiter et al. This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license.PermissionsRequest permissions for this article.Request PermissionsKEYWORDSSaccharomyces cerevisiaefermentationgene expressionContributorsAuthorsTaylor Reiter https://orcid.org/0000-0002-7388-421XFood Science Graduate Group, University of California, Davis, Davis, California, USADepartment of Viticulture and Enology, University of California, Davis, Davis, California, USADepartment of Population Health and Reproduction, University of California, Davis, Davis, California, USAView all articles by this authorRachel MontpetitDepartment of Viticulture and Enology, University of California, Davis, Davis, California, USAView all articles by this authorShelby ByerDepartment of Viticulture and Enology, University of California, Davis, Davis, California, USAView all articles by this authorIsadora FriasDepartment of Viticulture and Enology, University of California, Davis, Davis, California, USAView all articles by this authorEsmeralda LeonDepartment of Chemical Engineering, University of California, Davis, Davis, California, USAView all articles by this authorRobert VianoDepartment of Chemical Engineering, University of California, Davis, Davis, California, USAView all articles by this authorMichael McloughlinDepartment of Chemical Engineering, University of California, Davis, Davis, California, USAView all articles by this authorThomas HalliganDepartment of Chemical Engineering, University of California, Davis, Davis, California, USAView all articles by this authorDesmon HernandezDepartment of Chemical Engineering, University of California, Davis, Davis, California, USAView all articles by this authorRon RunnebaumDepartment of Viticulture and Enology, University of California, Davis, Davis, California, USADepartment of Chemical Engineering, University of California, Davis, Davis, California, USAView all articles by this authorBen Montpetit https://orcid.org/0000-0002-8317-983X [email protected]Food Science Graduate Group, University of California, Davis, Davis, California, USADepartment of Viticulture and Enology, University of California, Davis, Davis, California, USAView all articles by this authorEditorEdward G. DudleyEditorThe Pennsylvania State UniversityMetrics CitationsMetricsCitationsView OptionsMediaFiguresOtherTablesShare American Society for Microbiology (\"ASM\") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM (\"ASM Web Sites\") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.ACCEPT the cookie policy and & CONTINUEFIND OUT MORE about the privacy policy