Skip to main content
Log in

The galaxy of the non-Linnaean nomenclature

  • Original Paper
  • Published:
History and Philosophy of the Life Sciences Aims and scope Submit manuscript

Abstract

Contrary to the traditional claim that needs for unambiguous communication about animal and plant species are best served by a single set of names (Linnaean nomenclature) ruled by international Codes, I suggest that a more diversified system is required, especially to cope with problems emerging from aggregation of biodiversity data in large databases. Departures from Linnaean nomenclature are sometimes intentional, but there are also other, less obvious but widespread forms of not Code-compliant grey nomenclature. A first problem is due to the circumstance that the Codes are intended to rule over the way names are applied to species and other taxonomic units, whereas users of taxonomy need names to be applied to specimens. For different reasons, it is often impossible to refer a specimen with certainty to a named species, and in those cases an open nomenclature is employed. Second, molecular taxonomy leads to the discovery of clusters of gene sequence diversity not necessarily equivalent to the species recognized and named by taxonomists. Those clusters are mostly indicated with informal names or formulas that challenge comparison between different publications or databases. In several instances, it is not even clear if a formula refers to an individual voucher specimen, or is a provisional species name. The use of non-Linnaean names and formulas must be revised and strengthened by fixing standard formats for the different kinds of objects or hypotheses and providing permanent association of ‘grey names’ with standardized source information such as author and year. In the context of a broad-scope revisitation of aims and scope of scientific nomenclature, it may be worth rethinking if natural objects like plant galls and lichens, although other than the ‘single-entity’ objects traditionally covered by biological classifications, may nevertheless deserve taxonomic names.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Unfortunately, nomenclatural curation does not seem to be high priority in otherwise well-managed data bases. For example, in the list of taxa to which are referred the entries of the BOLD database (about which more in Sect. 3.2.), several species are listed twice, under alternative combinations of the same specific epithet with two different generic names, but the taxonomic identity of these names is not flagged. The following examples where the first of the two Latin names corresponding to the same species is the one in current use according to the Roskov et al. (2019) show that duplicate names are found in that database also for popularly known animals, such as yellow-cheeked gibbon (Nomascus gabriellae = Hylobates gabriellae), pygmy marmoset (Cebuella pygmaea = Callithrix pygmaea), Malayan tapir (Acrocodia indica = Tapirus indicus), Bahia porcupine (Sphiggurus insidiosus = Coendou insidiosus), pygmy hippopotamus (Hexaprotodon liberiensis = Choeropsis liberiensis), alpaca (Vicugna pacos = Lama pacos), harp seal (Pagophilus groenlandicus = Phoca groenlandica), American mink (Neovison vison = Mustela vison), eyra (Puma yagouaroundi = Herpailurus yagouaroundi), ground pangolin (Manis temminckii = Smutsia temminckii), long-tailed pangolin (Manis tetradactyla = Phataginus tetradactyla), tree pangolin (Manis tricuspis = Phataginus tricuspis).

  2. And still currently accepted (Wilson and Reeder 2005; Roskov et al. 2019).

  3. The original (valid) spelling of this genus name, introduced by Linnaeus (1758) is Raja; Raia is an incorrect subsequent spelling without separate status in zoological nomenclature, in the sense of Art. 33.3 of the Code (International Commission on Zoological Nomenclature 1999). Specifically, on Raja versus Raia, see Fricke et al. (2019), ad vocem.

  4. Little more than a curiosity is the survival of distinct names for the larva and the adult of the same animal. This double nomenclature was less exceptional in the XIX century, when knowledge about life cycles and metamorphoses was much less advanced than today. Thus Phyllosoma (described by Leach 1818), Nectochaeta (proposed by von Marenzeller 1892) and Pelagosphaera (introduced by Mingazzini 1905) were considered for a while to be independent genera (of crustaceans, polychaetes and peanut worms respectively), but were eventually recognized as larval stages of animals for which valid names had been previously established, based on adults. Things are admittedly less obvious when the name introduced for a larva is older than the name introduced for what eventually turns out to be the corresponding adult. Strict application of the Principle of Priority, one of the cornerstones of the Code, would fix the ‘larval’ name as valid for the adult too; but a ruling by the International Commission on Zoological Nomenclature might determine a reversal of precedence, as in the case of Phoronis Wright, 1856 (described on the adult animal) versus Actinotrocha Müller, 1846 (described on larvae), eventually suppressed (International Commission on Zoological Nomenclature 2015). The latter action should have terminated the paradoxical situation of both names being deliberately kept in use, one for the larva, one for the adult of the same animal, but this is practiced by some authors until now, e.g. with Actinotrocha branchiata and Phoronis muelleri, or Actinotrocha sabatieri and Phoronis psammophila (Conway 2015; Roskov et al. 2019).

  5. Unfortunately, this acronym, suggested by Heppell (1991) for the ill-fated International Standard Taxonomic Code, is actually in use for a different thing, the International Standard Text Code (see http://www.istc-international.org/, accessed August 1st, 2019). Ironic fate for the name of concepts intended to establish unique standards but also, as suggested by a reviewer of this article, an example of the fate of inadequately contextualized names.

References

  • Acharius, E. (1810). Lichenographia universalis. Göttingen: Danckwerts.

    Google Scholar 

  • Bely, A. E., & Weisblat, D. A. (2006). Lessons from leeches: a call for DNA barcoding in the lab. Evolution and Development, 8, 491–501.

    Google Scholar 

  • Bengtson, P. (1988). Open nomenclature. Palaeontology, 31, 223–227.

    Google Scholar 

  • Berendsohn, W. G. (1995). The concept of “potential taxa” in databases. Taxon, 44, 207–212.

    Google Scholar 

  • Berendsohn, W. G., Döring, M., Geoffroy, M., Glück, K., Güntsch, A., Hahn, A., et al. (2003). The Berlin Model: a concept-based taxonomic information model. Schriftenreihe für Vegetationskunde, 39, 15–42.

    Google Scholar 

  • Blaxter, M., Mann, J., Chapman, T., Thomas, F., Whitton, C., Robin Floyd, R., et al. (2005). Defining operational taxonomic units using DNA barcode data. Philosophical Transactions of the Royal Society of London B Biological Sciences, 360, 1935–1943.

    Google Scholar 

  • Boykin, L. M., Kinene, T., Wainaina, J. M., Savill, A., Seal, S., Mugerwa, H., et al. (2018). Review and guide to a future naming system of African Bemisia tabaci species. Systematic Entomology, 43, 427–433.

    Google Scholar 

  • Brunetti, R., Gissi, C., Pennati, R., Caicci, F., Gasparini, F., & Manni, L. (2015). Morphological evidence that the molecularly determined Ciona intestinalis type A and type B are different species: Ciona robusta and Ciona intestinalis. Journal of Zoological Systematics and Evolutionary Research, 53, 186–193.

    Google Scholar 

  • Brünnich, M. T. (1772). Zoologiae fundamenta praelectionibus academicis accommodata = Grunde i dyrelaeren. Hafniae et Lipsiae: Pelt.

  • Cantino, P. D., & de Queiroz, K. (2010). PhyloCode: International Code of Phylogenetic Nomenclature (Version 4c). http://www.ohio.edu/phylocode. Accessed July 31, 2019.

  • Collins, R. A., & Cruickshank, R. H. (2012). The seven deadly sins of DNA barcoding. Molecular Ecology Resources, 13, 969–975.

    Google Scholar 

  • Conway, D. V. P. (2015). Marine zooplankton of southern Britain. Part 3: Ostracoda, Stomatopoda, Nebaliacea, Mysida, Amphipoda, Isopoda, Cumacea, Euphausiacea, Decapoda, Annelida, Tardigrada, Nematoda, Phoronida, Bryozoa, Entoprocta, Brachiopoda, Echinodermata, Chaetognatha, Hemichordata and Chordata. (Edited by A. W. G. John). Occasional Publications. Marine Biological Association of the United Kingdom, No. 27, Plymouth.

  • de Bary, A. (1866). Morphologie und Physiologie der Pilze, Flechten und Myxomyceten. Leipzig: Engelmann.

    Google Scholar 

  • de Queiroz, K. (1988). Systematics and the Darwinian revolution. Philosophy of Science, 55, 238–259.

    Google Scholar 

  • de Queiroz, K., & Gauthier, J. (1990). Phylogeny as a central principle in taxonomy: Phylogenetic definitions of taxon names. Systematic Zoology, 39, 307–322.

    Google Scholar 

  • De Smet, W. M. A. (1991). Meeting user needs by an alternative nomenclature. In D. L. Hawksworth (Ed.), Improving the stability of names: Needs and options (pp. 179–181). Königstein: Koeltz Scientific Books.

    Google Scholar 

  • Dehal, P., Satou, Y., Campbell, R. K., Chapman, J., Degnan, B., De Tomaso, A., et al. (2002). The draft genome of Ciona intestinalis: insights into chordate and vertebrate origins. Science, 298, 2157–2167.

    Google Scholar 

  • Delaroche, F. E. (1809). Suite du mémoire sur les espèces de poissons observées à Iviça. Observations sur quelques-uns des poissons indiqués dans le précédent tableau et descriptions des espèces nouvelles ou peu connues. Annales du Muséum d’Histoire Naturelle, Paris, 13, 313–361, pls. 20–25.

  • Erxleben, J. C. P. (1777) Systema regni animalis per classes, ordines, genera, species, varietates: cum synonymia et historia animalium. Classis I: Mammalia. Lipsiae: Impensis Weygandianis.

  • Fišer, C., Alther, R., Zakšek, V., Borko, S., Fuchs, A., & Altermatt, F. (2018). Translating Niphargus barcodes from Switzerland into taxonomy with a description of two new species (Amphipoda, Niphargidae). ZooKeys, 760, 113–141.

    Google Scholar 

  • Fišer, C., Konec, M., Alther, R., Švara, V., & Altermatt, F. (2017). Taxonomic, phylogenetic and ecological diversity of Niphargus (Amphipoda: Crustacea) in the Hölloch cave system (Switzerland). Systematics and Biodiversity, 15, 218–237.

    Google Scholar 

  • Floyd, R., Eyualem, A., Papert, A., & Blaxter, M. (2002). Molecular barcodes for soil nematode identification. Molecular Ecology, 11, 839–850.

    Google Scholar 

  • Franz, N. M., Chen, M., Kianmajd, P., Yu, S., Bowers, S., Weakley, A. S., et al. (2016). Names are not good enough: reasoning over taxonomic change in the Andropogon complex. Semantic Web, 7, 645–667.

    Google Scholar 

  • Franz, N. M., & Peet, R. K. (2009). Towards a language for mapping relationships among taxonomic concepts. Systematics and Biodiversity, 7, 5–20.

    Google Scholar 

  • Franz, N. M., Peet, R. K., & Weakley, A. S. (2008). On the use of taxonomic concepts in support of biodiversity research and taxonomy. In Q. D. Wheeler (Ed.), The new taxonomy (pp. 63–86). Boca Raton: CRC Press.

    Google Scholar 

  • Fricke, R., Eschmeyer, W. N. & van der Laan, R. (Eds.). (2019). Eschmeyer’s Catalog of fishes: Genera, species, references. http://researcharchive.calacademy.org/research/ichthyology/catalog/fishcatmain.asp. Accessed July 30, 2019.

  • Groves, C., & Grubb, P. (2011). Ungulate taxonomy. Baltimore: The Johns Hopkins University Press.

    Google Scholar 

  • Hawksworth, D. L., Hibbett, D. S., Kirk, P. M., & Lücking, R. (2016). (308–310) Proposals to permit DNA sequence data to serve as types of names of fungi. Taxon, 65, 899–900.

    Google Scholar 

  • Hebert, P. D. N. (2003). Biological identifications through DNA barcodes. Proceedings of the Royal Society of London, Series B: Biological Sciences, 270, 313–321.

    Google Scholar 

  • Heppell, D. (1991). Names without number? In D. L. Hawksworth (Ed.), Improving the stability of names: Needs and options (pp. 191–196). Königstein: Koeltz.

    Google Scholar 

  • Horton, T., Gofas, S., Kroh, A., Poore, G. C. B., Read, G., Rosenberg, G., et al. (2017). Improving nomenclatural consistency: A decade of experience in the World Register of Marine Species. European Journal of Taxonomy, 389, 1–24.

    Google Scholar 

  • Huemer, P., & Karsholt, O. (1995). Gelechiidae. In A. Minelli, S. Ruffo, & S. La Posta (Eds.), Checklist delle specie della fauna italiana (Vol. 83, pp. 28–41). Bologna: Calderini.

    Google Scholar 

  • Iannelli, F., Pesole, G., Sordino, P., & Gissi, C. (2007). Mitogenomics reveals two cryptic species in Ciona intestinalis. Trends in Genetics, 23, 419–422.

    Google Scholar 

  • International Commission on Zoological Nomenclature. (1999). International Code of Zoological Nomenclature (4th ed.). London: The International Trust for Zoological Nomenclature.

    Google Scholar 

  • International Commission on Zoological Nomenclature. (2015). Opinion 2373 (Case 3626): Phoronis Wright, 1856 (Phoronida) and P. muelleri de Selys Longchamps, 1903: both names conserved. Bulletin of Zoological Nomenclature, 72, 327–328.

    Google Scholar 

  • Jolivet, P. (1998). Interrelationship between insects and plants. London: CRC.

    Google Scholar 

  • Jörger, K. M., Norenburg, J. L., Wilson, N. G., & Schrödl, M. (2012). Barcoding against a paradox? Combined molecular species delineations reveal multiple cryptic lineages in elusive meiofaunal sea slugs. BMC Evolutionary Biology, 12, 245.

    Google Scholar 

  • Jörger, K. M., & Schrödl, M. (2013). How to describe a cryptic species? Practical challenges of molecular taxonomy. Frontiers in Zoology, 10, 59.

    Google Scholar 

  • Kõljalg, U., Tedersoo, L., Nilsson, R. H., & Abarenkov, K. (2016). Digital identifiers for fungal species. Science, 352, 1182–1183.

    Google Scholar 

  • Kutschera, U., Langguth, H., Kuo, D.-H., Weisblat, D. A., & Shankland, M. (2013). Description of a new leech species from North America, Helobdella austinensis n. sp. (Hirudinea: Glossiphoniidae), with observations on its feeding behaviour. Zoosystematics and Evolution, 89, 239–246.

    Google Scholar 

  • Lapage, S. P., Sneath, P. H. A., Lessel, E. F., Skerman, V. B. D., Seeliger, H. P. R., & Clark, W. A. (1990). International code of nomenclature of bacteria. Washington, DC: ASM Press.

    Google Scholar 

  • Leach, W. E. (1818). Sur quelques genres nouveaux de Crustacés. Journal de Physique, 88, 304–307.

    Google Scholar 

  • Legg, J. P., French, R., Rogan, D., Okao-Okuja, G., & Brown, J. K. (2002). A distinct Bemisia tabaci (Gennadius) (Hemiptera: Sternorrhyncha: Aleyrodidae) genotype cluster is associated with the epidemic of severe cassava mosaic virus disease in Uganda. Molecular Ecology, 11, 1219–1229.

    Google Scholar 

  • Leonelli, S. (2016). Data-centric biology. Chicago, IL: University of Chicago Press.

    Google Scholar 

  • Lepage, D. (2019). Avibasethe World Bird Database. http://avibase.bsc-eoc.org. Accessed July 30, 2019.

  • Lepage, D., Vaidya, G., & Guralnick, R. (2014). Avibase—A database system for managing and organizing taxonomic concepts. ZooKeys, 420, 117–135.

    Google Scholar 

  • Linnaeus, C. (1751). Philosophia botanica in qua explicantur fundamenta botanica cum definitionibus partium, exemplis terminorum, observationibus rariorum. Stockholm: Kiesewetter.

  • Linnaeus, C. (1753). Species plantarum: exhibentes plantas rite cognitas, ad genera relatas, cum differentiis specificis, nominibus trivialibus, synonymis selectis, locis natalibus, secundum systema sexuale digestas. Holmiae: Laurentius Salvius.

  • Linnaeus, C. (1758). Systema naturæ per regna tria naturæ, secundum classes, ordines, genera, species, cum characteribus, differentiis, synonymis, locis, Edition X. 1. Stockholm: Laurentius Salvius.

  • Marks, E. N. (1983). Mosquitoes of the Purari River lowlands. In T. Petr (Ed.), The Purari: Tropical environment of a high rainfall river basin (pp. 531–550). The Hague: Junk.

    Google Scholar 

  • Matthews, S. C. (1973). Notes on open nomenclature and on synonymy lists. Palaeontology, 16, 713–719.

    Google Scholar 

  • Minelli, A. (2000). The ranks and the names of species and higher taxa, or, a dangerous inertia of the language of natural history. In M. T. Ghiselin & A. E. Leviton (Eds.), Cultures and institutions of natural history. Essays in the history and philosophy of science (pp. 339–351). San Francisco: California Academy of Sciences.

    Google Scholar 

  • Minelli, A. (2017a). Grey nomenclature needs rules. Ecologica Montenegrina, 7, 656–666.

    Google Scholar 

  • Minelli, A. (2017b). Updating taxonomic practice to cope with challenges from within and without the discipline. Biodiversity Journal, 8, 671–674.

    Google Scholar 

  • Minelli, A. (2017c). Lichens and galls—two families of chimeras in the space of form. Azafea, 19, 91–105.

    Google Scholar 

  • Mingazzini, P. (1905). Un Gefireo pelagico. Pelagosphaera Aloysii n. gen., n. sp. Rendiconti delle sedute solenni della R. Accademia Nazionale dei Lincei, 14, 713–720.

    Google Scholar 

  • Morard, R., Escarguel, G., Weiner, A. K., André, A., Douady, C. J., Wade, C. M., et al. (2016). Nomenclature for the nameless: A proposal for an integrative molecular taxonomy of cryptic diversity exemplified by planktonic Foraminifera. Systematic Biology, 65, 925–940.

    Google Scholar 

  • Mugerwa, H., Rey, M. E., Alicai, T., Ateka, E., Atuncha, H., Ndunguru, J., et al. (2012). Genetic diversity and geographic distribution of Bemisia tabaci (Gennadius) (Hemiptera: Aleyrodidae) genotypes associated with cassava in East Africa. Ecology and Evolution, 2, 2749–2762.

    Google Scholar 

  • Packer, L., Monckton, S. K., Onuferko, T. M., & Ferrari, R. R. (2018). Validating taxonomic identifications in entomological research. Insect Conservation and Diversity, 11, 1–12.

    Google Scholar 

  • Page, R. D. (2016). DNA barcoding and taxonomy: Dark taxa and dark texts. Philosophical Transactions of the Royal Society B, 371, 20150334.

    Google Scholar 

  • Pante, E., Schoelinck, C., & Puillandre, N. (2015). From integrative taxonomy to species description: one step beyond. Systematic Biology, 64, 152–160.

    Google Scholar 

  • Patterson, D., Mozzherin, D., Shorthouse, D., & Thessen, A. (2016). Challenges with using names to link digital biodiversity information. Biodiversity Data Journal, 4, e8080.

    Google Scholar 

  • Pérez-Ponce de León, G., & Nadler, S. A. (2010). What we don’t recognize can hurt us: A plea for awareness about cryptic species. Journal of Parasitology, 96, 453–464.

    Google Scholar 

  • Piantadosi, S. T., Tily, H., & Gibson, E. (2012). The communicative function of ambiguity in language. Cognition, 122, 280–291.

    Google Scholar 

  • Pinacho-Pinacho, C. D., García-Varela, M., Sereno-Uribe, A. L., & Pérez-Ponce de León, G. (2018). A hyper-diverse genus of acanthocephalans revealed by tree-based and nontree-based species delimitation methods: ten cryptic species of Neoechinorhynchus in Middle American freshwater fishes. Molecular Phylogenetics and Evolution, 127, 30–45.

    Google Scholar 

  • Pleijel, F. (2000). Phylogenetic taxonomy, a farewell to species, and a revision of Heteropodarke (Hesionidae, Polychaeta, Annelida). Systematic Biology, 48, 755–789.

    Google Scholar 

  • Pleijel, F., & Rouse, G. W. (1999). Least-inclusive taxonomic unit: a new taxonomic concept for biology. Proceedings of the Royal Society of London B, 267, 627–630.

    Google Scholar 

  • Pleijel, F., & Rouse, G. W. (2000). A new taxon, capricornia (Hesionidae, Polychaeta), illustrating the LITU (‘least-inclusive taxonomic unit’) concept. Zoologica Scripta, 29, 157–168.

    Google Scholar 

  • Pyle, R., & Michel, E. (2008). Zoobank: developing a nomenclatural tool for unifying 250 years of biological information. Zootaxa, 1950, 39–50.

    Google Scholar 

  • Ratnasingham, S., & Hebert, P. D. N. (2007). BARCODING, BOLD: The barcode of life data system (www.barcodinglife.org). Molecular Ecology Notes, 7, 355–364.

    Google Scholar 

  • Richter, R. (1948). Einführung in die Zoologische Nomenklatur durch Erläuterung der Internationalen Regeln (2nd ed.). Frankfurt: Waldemar Kramer.

    Google Scholar 

  • Risso, A. (1810). Ichthyologie de Nice, ou, Histoire naturelle des poissons du département des Alpes Maritimes. Paris: Schoell.

    Google Scholar 

  • Roskov Y., Ower G., Orrell T., Nicolson D., Bailly N., & Kirk P. M., et al. (Eds.) (2019). Species 2000 and ITIS catalogue of life, 2019 annual checklist. Leiden: Species 2000 (Naturalis). www.catalogueoflife.org/annual-checklist/2019. Accessed July 30, 2019.

  • Ryberg, M., & Nilsson, R. H. (2018). New light on names and naming of dark taxa. MycoKeys, 30, 31–39.

    Google Scholar 

  • Samyn, Y., & De Clerck, O. (2012). No name, no game. European Journal of Taxonomy, 10, 1–3.

    Google Scholar 

  • Schindel, D. E., & Miller, S. E. (2010). Provisional nomenclature: The on-ramp to taxonomic names. In A. Polaszek (Ed.), Systema naturae 250: The Linnaean Ark (pp. 109–115). Boca Raton: CRC.

    Google Scholar 

  • Schwendener, S. (1868). Über die Beziehungen zwischen Algen und Flechtengonidien. Botanische Zeitung, 26, 289–292.

    Google Scholar 

  • Siddall, M. E., & Borda, E. (2003). Phylogeny and revision of the leech genus Helobdella (Glossiphoniidae) based on mitochondrial gene sequences and morphological data and a special consideration of the triserialis complex. Zoologica Scripta, 32, 23–33.

    Google Scholar 

  • Sigovini, M., Keppel, E., & Tagliapietra, D. (2016). Open nomenclature in the biodiversity era. Methods in Ecology and Evolution, 7, 1217–1225.

    Google Scholar 

  • Sseruwagi, P., Legg, J. P., Maruthi, M. N., Colvin, J., Rey, M. E. C., & Brown, J. K. (2005). Genetic diversity of Bemisia tabaci (Gennadius) (Hemiptera: Aleyrodidae) populations and presence of the B biotype and non-B biotype that can induce silverleaf symptoms in squash, in Uganda. Annals of Applied Biology, 147, 253–265.

    Google Scholar 

  • Sterner, B., & Franz, N. M. (2017). Taxonomy for humans or computers? Cognitive pragmatics for big data. Biological Theory, 12, 99–111.

    Google Scholar 

  • Trontelj, P., & Fišer, C. (2009). Cryptic species diversity should not be trivialized. Systematics and Biodiversity, 7, 1–3.

    Google Scholar 

  • Turland, N. J., Wiersema, J. H., Barrie, F. R., Greuter, W., Hawksworth, D. L., Herendeen, P. S., et al. (Eds.). (2018). International code of nomenclature for algae, fungi, and plants (Shenzhen Code) adopted by the nineteenth international botanical congress Shenzhen, China. Glashutten: Koeltz Botanical Books.

    Google Scholar 

  • von Marenzeller, E. (1892). Sur une Polynoïde pelagique (Nectochaeta grimaldii, nov. gen., nov. sp.) recueillie par l’Hirondelle en 1888. Bulletin de la Société Zoologique de France, 17, 183–185.

    Google Scholar 

  • Wilson, D. E., & Reeder, D. M. (Eds.) (2005). Mammal species of the world: A taxonomic and geographic reference (Vols. 1, 2) 3rd Ed. Baltimore: Johns Hopkins University Press.

  • Zachos, F. E., Apollonio, M., Barmann, E. V., Festa-Bianchet, M., Gohlich, U., Habel, J. C., et al. (2013). Species inflation and taxonomic artefacts. A critical comment on recent trends in mammalian classification. Mammalian Biology, 78, 1–6.

    Google Scholar 

Download references

Acknowledgements

I am grateful to Catherine Kendig and Joeri Witteveen for inviting me to contribute to this special issue and to both of them, two anonymous referees and the journal’s Editor Sabina Leonelli for their helpful suggestions on previous versions of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alessandro Minelli.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Minelli, A. The galaxy of the non-Linnaean nomenclature. HPLS 41, 31 (2019). https://doi.org/10.1007/s40656-019-0271-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s40656-019-0271-0

Keywords

Navigation