Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Review Article
  • Published: September 2007

Mechanistic approaches to the study of evolution: the functional synthesis

  • Antony M. Dean 1 &
  • Joseph W. Thornton 2  

Nature Reviews Genetics volume  8 ,  pages 675–688 ( 2007 ) Cite this article

5468 Accesses

266 Citations

32 Altmetric

Metrics details

By combining evolutionary sequence analyses and manipulative molecular experiments, the functional synthesis of molecular evolution provides a powerful framework to elucidate the mechanisms by which historical mutations have altered biochemical processes and produced novel phenotypes. By using this approach, inferred ancestral sequences can be resurrected and their phenotypes and fitness effects assessed experimentally.

The functional synthesis of molecular evolution provides independent corroboration of statistical inferences that have been drawn from sequence analyses.

The functional synthesis of molecular evolution explicitly connects genotype with phenotype to allow mechanistic insights into the causes of adaptive change and evolutionary constraint.

The functional synthesis of molecular evolution provides decisive tests of recent adaptations where genetic variation still segregates in present-day species, and of ancient adaptations where genetic variation is fixed in present-day species.

The functional synthesis of molecular evolution can resolve long-standing questions about evolutionary processes and important evolutionary questions about metabolic, cellular, developmental and behavioural systems.

The functional synthesis of molecular evolution can be used to characterize adaptive landscapes and explore the evolution of complexity.

The functional synthesis of molecular evolution is poised to move beyond studies of single genes to allow the analysis of the evolution of pathways and networks that are made up of multiple genes.

The functional synthesis of molecular evolution should become routine in studies of molecular evolution.

An emerging synthesis of evolutionary biology and experimental molecular biology is providing much stronger and deeper inferences about the dynamics and mechanisms of evolution than were possible in the past. The new approach combines statistical analyses of gene sequences with manipulative molecular experiments to reveal how ancient mutations altered biochemical processes and produced novel phenotypes. This functional synthesis has set the stage for major advances in our understanding of fundamental questions in evolutionary biology. Here we describe this emerging approach, highlight important new insights that it has made possible, and suggest future directions for the field.

This is a preview of subscription content, access via your institution

Access options

Subscribe to this journal

Receive 12 print issues and online access

$189.00 per year

only $15.75 per issue

Buy this article

  • Purchase on Springer Link
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

mechanistic hypothesis biology

Similar content being viewed by others

mechanistic hypothesis biology

Phylogenomics and the rise of the angiosperms

mechanistic hypothesis biology

Genome-wide association studies

mechanistic hypothesis biology

Aging clocks based on accumulating stochastic variation

Dykhuizen, D. E. & Hartl, D. L. Selection in chemostats. Microbiol. Rev. 47 , 150–168 (1983).

CAS   PubMed   PubMed Central   Google Scholar  

Elena, S. F. & Lenski, R. E. Evolution experiments with microorganisms: the dynamics and genetic bases of adaptation. Nature Rev. Genet. 4 , 457–469 (2003).

Article   CAS   PubMed   Google Scholar  

Losos, J. B., Jackman, T. R., Larson, A., Queiroz, K. & Rodriguez-Schettino, L. Contingency and determinism in replicated adaptive radiations of island lizards. Science 279 , 2115–2118 (1998).

Powers, D. A., Lauerman, T., Crawford, D. & DiMichele, L. Genetic mechanisms for adapting to a changing environment. Annu. Rev. Genet. 25 , 629–659 (1991).

Wourms, M. K. & Wasserman, F. E. Butterfly wing markings are more advantageous during handling than during the initial strike of an avian predator. Evolution 39 , 845–851 (1985).

Article   PubMed   Google Scholar  

Moller, A. P. Female choice selects for male sexual tail ornaments in the monogamous swallow. Nature 332 , 640–642 (1988).

Article   Google Scholar  

Rainey, P. B. & Rainey, K. Evolution of cooperation and conflict in experimental bacterial populations. Nature 425 , 72–74 (2003).

Denver, D. R. et al. The transcriptional consequences of mutation and natural selection in Caenorhabditis elegans . Nature Genet. 37 , 544–548 (2005).

Endler, J. A. Natural selection on color patterns in Poecilia reticulata . Evolution 34 , 76–91 (1980).

Thornton, J. W. Resurrecting ancient genes: experimental analysis of extinct molecules. Nature Rev. Genet. 5 , 366–375 (2004). An introduction to gene 'resurrection' — that is, phylogenetic reconstruction, biochemical synthesis and functional characterization of ancient sequences — as a strategy for testing evolutionary hypotheses.

Fitzpatrick, M. J., Feder, E., Rowe, L. & Sokolowski, M. B. Maintaining a behaviour polymorphism by frequency-dependent selection on a single gene. Nature 447 , 210–212 (2007). Polymorphism in activity of a cGMP-dependent protein kinase that elicits different foraging behaviours in Drosophila melanogaster larva is maintained by negative frequency-dependent selection during starvation.

Douglas, S. J., Dawson-Scully, K. & Sokolowski, M. B. The neurogenetics and evolution of food-related behaviour. Trends Neurosci. 28 , 644–652 (2005).

Toth, A. L. & Robinson, G. E. Evo–devo and the evolution of social behavior. Trends Genet. 23 , 334–341 (2007).

Colosimo, P. F. et al. Widespread parallel evolution in sticklebacks by repeated fixation of Ectodysplasin alleles. Science 307 , 1928–1933 (2005). Using transgenic, phylogenetic and quantitative genetic analysis, the authors identify an allele that is involved in the loss of external armour during the evolution of numerous freshwater stickleback populations.

Gompel, N., Prud'homme, B., Wittkopp, P. J., Kassner, V. A. & Carroll, S. B. Chance caught on the wing: cis -regulatory evolution and the origin of pigment patterns in Drosophila . Nature 433 , 481–487 (2005). This pioneering study in the evolution of development (along with reference 16) used transgenic techniques to identify decisively specific regulatory elements that underlie evolutionary differences in the expression of genes that drive pigmentation patterns between fruitfly species.

Prud'homme, B. et al. Repeated morphological evolution through cis -regulatory changes in a pleiotropic gene. Nature 440 , 1050–1053 (2006).

Konishi, S. et al. An SNP caused loss of seed shattering during rice domestication. Science 312 , 1392–1396 (2006). A beautiful example of the functional synthesis to study the evolution of development: fine quantitative genetic mapping identified a single substitution associated with the loss of seed shattering that occurs during rice domestication; transgenic and functional analysis established the specific effects of the historical mutation on gene expression, seed development and the shattering phenotype.

Wang, H. et al. The origin of the naked grains of maize. Nature 436 , 714–719 (2005).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Daborn, P. J. et al. A single p450 allele associated with insecticide resistance in Drosophila . Science 297 , 2253–2256 (2002).

Chung, H. et al. Cis -regulatory elements in the accord retrotransposon result in tissue-specific expression of the Drosophila melanogaster insecticide resistance gene Cyp6g1 . Genetics 175 , 1071–1077 (2007). Using genetic manipulation in fruitflies, the authors show decisively that a transposon insertion in the regulatory region of a gene that metabolizes insecticides is sufficient to recapitulate the evolution of DDT (dichloro-diphenyl-trichloroethane) resistance.

Bantinaki, E. et al. Adaptive divergence in experimental populations of Pseudomonas fluorescens . III. Mutational origins of wrinkly spreader diversity. Genetics 176 , 441–453 (2007).

Gilbert, S. F., Opitz, J. M. & Raff, R. A. Resynthesizing evolutionary and developmental biology. Dev. Biol. 173 , 357–372 (1996).

Wilson, E. O. Sociobiology: The New Synthesis (Belknap, Cambridge, 1975).

Google Scholar  

Hubby, J. L. & Lewontin, R. C. A molecular approach to the study of genic heterozygosity in natural populations. I. The number of alleles at different loci in Drosophila pseudoobscura . Genetics 54 , 577–594 (1966).

Kreitman, M. Nucleotide polymorphism at the alcohol dehydrogenase locus of Drosophila melanogaster . Nature 304 , 412–417 (1983).

Golding, G. B. & Dean, A. M. The structural basis of molecular adaptation. Mol. Biol. Evol. 15 , 355–369 (1998). A review summarizing classic early work in the functional synthesis.

McDonald, J. H. & Kreitman, M. Adaptive protein evolution at the adh locus in Drosophila . Nature 351 , 652–654 (1991).

Yang, Z. & Nielsen, R. Codon-substitution models for detecting molecular adaptation at individual sites along specific lineages. Mol. Biol. Evol. 19 , 908–917 (2002).

Tajima, F. Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics 123 , 585–596 (1989).

Nielsen, R. Molecular signatures of natural selection. Annu. Rev. Genet. 39 , 197–218 (2005).

Nielsen, R. Statistical tests of selective neutrality in the age of genomics. Heredity 86 , 641–647 (2001).

Eyre-Walker, A. Changing effective population size and the McDonald–Kreitman test. Genetics 62 , 2017–2024 (2002).

Koehn, R. K. & Hilbish, T. J. The adaptive importance of genetic variation. Am. Sci. 75 , 134–141 (1987).

Watt, W. B. in The Evolution of Population Biology (eds Singh, R. S. & Uyenoyama, M. K.) (Cambridge Univ. Press, Cambridge, 2004).

Wheat, C. W., Watt, W. B., Pollock, D. D. & Schulte, P. M. From DNA to fitness differences: sequences and structures of adaptive variants of Colias phosphoglucose isomerase (PGI). Mol. Biol. Evol. 23 , 499–512 (2006).

Brideau, N. J. et al. Two Dobzhansky–Muller genes interact to cause hybrid lethality in Drosophila . Science 314 , 1292–1295 (2006).

Geffeney, S. L., Fujimoto, E., Brodie, E. D. 3rd, Brodie, E. D. Jr & Ruben, P. C. Evolutionary diversification of TTX-resistant sodium channels in a predator–prey interaction. Nature 434 , 759–763 (2005). An elegant mechanistic-structural explanation of the repeated evolution of resistance to the tetrodotoxin of toxic newt prey, caused by amino-acid replacements in the voltage-gated sodium channels of garter snake muscles.

Zhang, J. Parallel adaptive origins of digestive RNases in Asian and African leaf monkeys. Nature Genet. 38 , 819–823 (2006). Evolution of foregut fermentation in Asian and African leaf-eating monkeys is characterized by parallel amino-acid replacements that produce similar functional shifts in digestive RNases.

Zhang, J. et al. The crystal structure of a high oxygen affinity species of haemoglobin (bar-headed goose haemoglobin in the oxy form). J. Mol. Biol. 255 , 484–493 (1996).

Zhang, J. & Rosenberg, H. F. Complementary advantageous substitutions in the evolution of an antiviral RNase of higher primates. Proc. Natl Acad. Sci. USA 99 , 5486–5491 (2002).

Jermann, T. M., Opitz, J. G., Stackhouse, J. & Benner, S. A. Reconstructing the evolutionary history of the artiodactyl ribonuclease superfamily. Nature 374 , 57–59 (1995).

Jessen, T. H., Weber, R. E., Fermi, G., Tame, J. & Braunitzer, G. Adaptation of bird hemoglobins to high altitudes: demonstration of molecular mechanism by protein engineering. Proc. Natl Acad. Sci. USA 88 , 6519–6522 (1991).

Gaucher, E. A., Thomson, J. M., Burgan, M. F. & Benner, S. A. Inferring the palaeoenvironment of ancient bacteria on the basis of resurrected proteins. Nature 425 , 285–288 (2003).

Chang, B. S., Jonsson, K., Kazmi, M. A., Donoghue, M. J. & Sakmar, T. P. Recreating a functional ancestral archosaur visual pigment. Mol. Biol. Evol. 19 , 1483–1489 (2002).

Thomson, J. M. et al. Resurrecting ancestral alcohol dehydrogenases from yeast. Nature Genet. 37 , 630–635 (2005). Ancestral yeast alcohol dehydrogenase (ADH) was reconstructed, expressed and shown to have the functional characteristics that are typical of extant ADH1, which is involved in ethanol production, rather than ADH2, which is involved in ethanol consumption.

Ugalde, J. A., Chang, B. S. & Matz, M. V. Evolution of coral pigments recreated. Science 305 , 1433 (2004).

Soong, T. W. & Venkatesh, B. Adaptive evolution of tetrodotoxin resistance in animals. Trends Genet. 22 , 621–626 (2006).

Doebley, J., Stec, A. & Hubbard, L. The evolution of apical dominance in maize. Nature 386 , 485–488 (1997).

Shapiro, M. D. et al. Genetic and developmental basis of evolutionary pelvic reduction in threespine sticklebacks. Nature 428 , 717–723 (2004).

Stern, D. L. Evolutionary developmental biology and the problem of variation. Evolution 54 , 1079–1091 (2000).

Sucena, E., Delon, I., Jones, I., Payre, F. & Stern, D. L. Regulatory evolution of shavenbaby / ovo underlies multiple cases of morphological parallelism. Nature 424 , 935–938 (2003).

Shimizu, K. K. et al. Darwinian selection on a selfing locus. Science 306 , 2081–2084 (2004).

de Meaux, J., Pop, A. & Mitchell-Olds, T. cis -regulatory evolution of chalcone-synthase expression in the genus Arabidopsis . Genetics 174 , 2181–2202 (2006).

Benderoth, M. et al. Positive selection driving diversification in plant secondary metabolism. Proc. Natl Acad. Sci. USA 103 , 9118–9123 (2006).

Protas, M. E. et al. Genetic analysis of cavefish reveals molecular convergence in the evolution of albinism. Nature Genet. 38 , 107–111 (2006).

Osborne, K. A. et al. Natural behavior polymorphism due to a cGMP-dependent protein kinase of Drosophila . Science 277 , 834–836 (1997).

Newcomb, R. D. et al. A single amino acid substitution converts a carboxylesterase to an organophosphorus hydrolase and confers insecticide resistance on a blowfly. Proc. Natl Acad. Sci. USA 94 , 7464–7468 (1997).

Newcomb, R. D., East, P. D., Russell, R. J. & Oakeshott, J. G. Isolation of α cluster esterase genes associated with organophosphate resistance in Lucilia cuprina. Insect Mol. Biol. 5 , 211–216 (1996).

Parker, A. G., Campbell, P. M., Spackman, M. E., Russell, R. J. & Oakeshott, J. G. Comparison of an esterase associated with organophosphate resistance in Lucilia cuprina with an orthologue not associated with resistance in Drosophila melanogaster . Pestic. Biochem. Physiol. 55 , 85–99 (1996).

Newcomb, R. D., Campbell, P. M., Russell, R. J. & Oakeshott, J. G. cDNA cloning, baculovirus-expression and kinetic properties of the esterase, E3, involved in organophosphorus resistance in Lucilia cuprina . Insect Biochem. Mol. Biol. 27 , 15–25 (1997).

Hartley, C. J. et al. Amplification of DNA from preserved specimens shows blowflies were preadapted for the rapid evolution of insecticide resistance. Proc. Natl Acad. Sci. USA 103 , 8757–8762 (2006).

Claudianos, C., Russell, R. J. & Oakeshott, J. G. The same amino acid substitution in orthologous esterases confers organophosphate resistance on the house fly and a blowfly. Insect Biochem. Mol. Biol. 29 , 675–686 (1999).

Hoekstra, H. E., Hirschmann, R. J., Bundey, R. A., Insel, P. A. & Crossland, J. P. A single amino acid mutation contributes to adaptive beach mouse color pattern. Science 313 , 101–104 (2006).

Steiner, C. C., Weber, J. N. & Hoekstra, H. E. Adaptive variation in beach mice produced by interacting pigmentation genes. PLoS Biol. 5 , e219 (2007).

Yokoyama, S., Zhang, H., Radlwimmer, F. B. & Blow, N. S. Adaptive evolution of color vision of the Comoran coelacanth ( Latimeria chalumnae ). Proc. Natl Acad. Sci. USA 96 , 6279–6284 (1999).

Yokoyama, S. & Tada, T. Adaptive evolution of the African and Indonesian coelacanths to deep-sea environments. Gene 261 , 35–42 (2000).

Yokoyama, S. Color vision of the coelacanth ( Latimeria chalumnae ) and adaptive evolution of rhodopsin ( RH1 ) and rhodopsin-like ( RH2 ) pigments. J. Hered. 91 , 215–220 (2000).

Shi, Y., Radlwimmer, F. B. & Yokoyama, S. Molecular genetics and the evolution of ultraviolet vision in vertebrates. Proc. Natl Acad. Sci. USA 98 , 11731–11736 (2001).

Shi, Y. & Yokoyama, S. Molecular analysis of the evolutionary significance of ultraviolet vision in vertebrates. Proc. Natl Acad. Sci. USA 100 , 8308–8313 (2003).

Yokoyama, S., Starmer, W. T., Takahashi, Y. & Tada, T. Tertiary structure and spectral tuning of UV and violet pigments in vertebrates. Gene 365 , 95–103 (2006).

Yokoyama, S. & Radlwimmer, F. B. The molecular genetics and evolution of red and green color vision in vertebrates. Genetics 158 , 1697–1710 (2001).

Yokoyama, S. & Radlwimmer, F. B. The molecular genetics of red and green color vision in mammals. Genetics 153 , 919–932 (1999).

Yokoyama, S. & Radlwimmer, F. B. The 'five-sites' rule and the evolution of red and green color vision in mammals. Mol. Biol. Evol. 15 , 560–567 (1998).

Wright, S. The roles of mutation, inbreeding, crossbreeding, and selection in evolution. Proc. 6th Int. Cong. Genet. 1 , 356–366 (1932).

Phillips, P. C. & Arnold, S. J. Visualizing multivariate selection. Evolution 43 , 1209–1222 (1989).

Gavrilets, S. A dynamical theory of speciation on holey adaptive landscapes. Am. Nat. 154 , 1–22 (1999).

Gillespie, J. H. Molecular evolution over the mutational landscape. Evolution 38 , 1116–1129 (1984).

Kauffman, S. A. The Origins of Order: Self-organization and Selection in Evolution (Oxford Univ. Press, Oxford, 1993).

Fisher, R. A. The Genetical Theory of Natural Selection (Oxford Univ. Press, Oxford, 1930).

Book   Google Scholar  

Maynard Smith, J. Natural selection and the concept of a protein space. Nature 225 , 563–564 (1970).

Weinreich, D. M., Delaney, N. F., Depristo, M. A. & Hartl, D. L. Darwinian evolution can follow only very few mutational paths to fitter proteins. Science 312 , 111–114 (2006).

Gillespie, J. H. A simple stochastic gene substitution model. Theor. Popul. Biol. 23 , 2020–2015 (1983).

Weinreich, D. M., Watson, R. A. & Chao, L. Perspective: sign epistasis and genetic constraint on evolutionary trajectories. Evolution 59 , 1165–1174 (2005).

CAS   PubMed   Google Scholar  

Lunzer, M., Miller, S. P., Felsheim, R. & Dean, A. M. The biochemical architecture of an ancient adaptive landscape. Science 310 , 499–501 (2005).

Zhu, G., Golding, G. B. & Dean, A. M. The selective cause of an ancient adaptation. Science 307 , 1279–1282 (2005).

Hurley, J. H. & Dean, A. M. Structure of 3-isopropylmalate dehydrogenase in complex with NAD + : ligand-induced loop closing and mechanism for cofactor specificity. Structure 2 , 1007–1016 (1994).

Hurley, J. H., Dean, A. M., Koshland, D. E. Jr & Stroud, R. M. Catalytic mechanism of NADP + -dependent isocitrate dehydrogenase: implications from the structures of magnesium-isocitrate and NADP + complexes. Biochemistry 30 , 8671–8678 (1991).

Miller, S. P., Lunzer, M. & Dean, A. M. Direct demonstration of an adaptive constraint. Science 314 , 458–461 (2006).

Dean, A. M. & Golding, G. B. Protein engineering reveals ancient adaptive replacements in isocitrate dehydrogenase. Proc. Natl Acad. Sci. USA 94 , 3104–3109 (1997).

Chen, R., Greer, A. & Dean, A. M. A highly active decarboxylating dehydrogenase with rationally inverted coenzyme specificity. Proc. Natl Acad. Sci. USA 92 , 11666–11670 (1995).

Gould, S. J. & Lewontin, R. C. The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme. Proc. R. Soc. Lond. B Biol. Sci. 205 , 581–598 (1979).

Bridgham, J. T., Carroll, S. M. & Thornton, J. W. Evolution of hormone-receptor complexity by molecular exploitation. Science 312 , 97–101 (2006).

Ortlund, E. A., Bridgham, J. T., Redinbo, M. R. & Thornton, J. W. Crystal structure of an ancient protein: evolution of a new function by conformational epistasis. Science 16 August 2007 (doi:101126/science.1142819).

Thornton, J. W., Need, E. & Crews, D. Resurrecting the ancestral steroid receptor: ancient origin of estrogen signaling. Science 301 , 1714–1717 (2003).

Thornton, J. W. Evolution of vertebrate steroid receptors from an ancestral estrogen receptor by ligand exploitation and serial genome expansions. Proc. Natl Acad. Sci. USA 98 , 5671–5676 (2001).

Paley, W. Natural Theology: or, Evidences of the Existence and Attributes of the Deity, Collected from the Appearances of Nature (E. Paulder, London, 1802).

Force, A. et al. Preservation of duplicate genes by complementary, degenerative mutations. Genetics 151 , 1531–1545 (1999).

Ohno, S. Evolution by Gene Duplication (Springer, New York, 1970).

Copley, S. D. Enzymes with extra talents: moonlighting functions and catalytic promiscuity. Curr. Opin. Chem. Biol. 7 , 265–272 (2003).

O'Brien, P. J. & Herschlag, D. Catalytic promiscuity and the evolution of new enzymatic activities. Chem. Biol. (London) 6 , R91–R105 (1999).

CAS   Google Scholar  

Khersonsky, O., Roodveldt, C. & Tawfik, D. S. Enzyme promiscuity: evolutionary and mechanistic aspects. Curr. Opin. Chem. Biol. 10 , 498–508 (2006).

Gerlt, J. A. & Babbitt, P. C. Divergent evolution of enzymatic function: mechanistically diverse superfamilies and functionally distinct suprafamilies. Annu. Rev. Biochem. 70 , 209–246 (2001).

Carroll, S. B. Endless forms: the evolution of gene regulation and morphological diversity. Cell 101 , 577–580 (2000).

Hoekstra, H. E. & Coyne, J. A. The locus of evolution: evo devo and the genetics of adaptation. Evolution 61 , 995–1016 (2007).

Leu, J. Y. & Murray, A. W. Experimental evolution of mating discrimination in budding yeast. Curr. Biol. 16 , 280–286 (2006).

Velicer, G. J. et al. Comprehensive mutation identification in an evolved bacterial cooperator and its cheating ancestor. Proc. Natl Acad. Sci. USA 103 , 8107–8112 (2006).

Woods, R., Schneider, D., Winkworth, C. L., Riley, M. A. & Lenski, R. E. Tests of parallel molecular evolution in a long-term experiment with Escherichia coli . Proc. Natl Acad. Sci. USA 103 , 9107–9112 (2006).

Zhong, S., Khodursky, A., Dykhuizen, D. E. & Dean, A. M. Evolutionary genomics of ecological specialization. Proc. Natl Acad. Sci. USA 101 , 11719–11724 (2004).

Davidson, E. H. The Regulatory Genome: Gene Regulatory Networks in Development and Evolution (Academic, Burlington, 2006).

Kuhn, T. The Structure of Scientific Revolutions (Univ. Chicago Press, Chicago, 1996).

Kornegay, J. R., Schilling, J. W. & Wilson, A. C. Molecular adaptation of a leaf-eating bird: stomach lysozyme of the hoatzin. Mol. Biol. Evol. 11 , 921–928 (1994).

Nachman, M. W., Hoekstra, H. E. & D'Agostino, S. L. The genetic basis of adaptive melanism in pocket mice. Proc. Natl Acad. Sci. USA 100 , 5268–5273 (2003).

Hiebl, I., Braunitzer, G. & Schneeganss, D. The primary structures of the major and minor hemoglobin-components of adult Andean goose ( Chloephaga melanoptera , Anatidae): the mutation Leu>Ser in position 55 of the β-chains. Biol. Chem. Hoppe-Seyler 368 , 1559–1569 (1987).

Bateson, W. Preface from Mendel's Principles of Heredity: A Defense (Cambridge Univ. Press, Cambridge, 1902).

Goldschmidt, R. The material basis of evolution (Yale, New Haven, 1940).

Bateson, W. Materials for the Study of Variation Treated with Especial Regard to Discontinuity in the Origin of Species (Macmillan, London, 1894).

Wilks, H. M. et al. A specific, highly active malate dehydrogenase by redesign of a lactate dehydrogenase framework. Science 242 , 1541–1544 (1988).

Cresko, W. A. et al. Parallel genetic basis for repeated evolution of armor loss in Alaskan threespine stickleback populations. Proc. Natl Acad. Sci. USA 101 , 6050–6055 (2004).

Spiller, B., Gershenson, A., Arnold, F. H. & Stevens, R. C. A structural view of evolutionary divergence. Proc. Natl Acad. Sci. USA 96 , 12305–12310 (1999).

Rothman, S. C., Voorhies, M. & Kirsch, J. F. Directed evolution relieves product inhibition and confers in vivo function to a rationally designed tyrosine aminotransferase. Protein Sci. 13 , 763–772 (2004).

Oue, S., Okamoto, A., Yano, T. & Kagamiyama, H. Redesigning the substrate specificity of an enzyme by cumulative effects of the mutations of non-active site residues. J. Biol. Chem. 274 , 2344–2349 (1999).

Hsu, C. C., Hong, Z., Wada, M., Franke, D. & Wong, C. H. Directed evolution of D-sialic acid aldolase to L-3-deoxy-manno-2-octulosonic acid (L-KDO) aldolase. Proc. Natl Acad. Sci. USA 102 , 9122–9126 (2005).

Whitlock, M. C., Phillips, P. C., Moore, F. B. G. & Tonsor, S. J. Multiple fitness peaks and epistasis. Annu. Rev. Ecol. Syst. 26 , 601–629 (1995).

Provine, W. B. The origins of theoretical population genetics (Univ. Chicago Press, Chicago, 1971).

Gavrilets, S. Evolution and speciation on holey adaptive landscapes. Trends Ecol. Evol. 12 , 307–312 (1997).

Yokoyama, S. Molecular evolution of color vision in vertebrates. Gene 300 , 69–78 (2002).

Download references

Acknowledgements

We thank M. Borello, M. Travisano, P. Phillips, B. Cresko, P. Rainey, A. Kondrashov, S. Yokoyama, R. Newcomb, D.Weinreich, B. Hall, an anonymous referee and members of the Thornton and Dean laboratories for comments. Supported by the US National Science Foundation (NSF IOB-0546906), the US National Institutes of Health (NIH R01-GM081592), and a Sloan Foundation Fellowship to J.W.T. and NIH R01-GM060,611 to A.M.D.

Author information

Authors and affiliations.

140 Gortner Laboratories, University of Minnesota, St Paul, 55108, Minnesota, USA

Antony M. Dean

5289 University of Oregon, Eugene, 97403, Oregon, USA

Joseph W. Thornton

You can also search for this author in PubMed   Google Scholar

Corresponding authors

Correspondence to Antony M. Dean or Joseph W. Thornton .

Ethics declarations

Competing interests.

The authors declare no competing financial interests.

Related links

Further information.

Anthony M. Dean's homepage

Joseph W. Thornton's homepage

The study of the origin and evolution of development, originally restricted to comparative methods, but increasingly using experimental approaches.

A mathematical framework, based on the genealogy of alleles, for estimating population genetic alleles.

A population genetic model in which beneficial mutations are fixed sequentially in the population through a series of selective sweeps, and in which neutral and deleterious mutations can be ignored as having low probabilities of fixation.

A precise assay of the relative growth rates (fitnesses) of competing strains can be obtained in the chemostat, a continuous culture device that is used to impose starvation for a specific resource in a constant environment.

A library of random mutants that have been generated by PCR amplification of a gene is ligated into a plasmid, transformed into a strain and screened for a desired function.

A complex of substrate bound to enzyme just before catalysis.

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Dean, A., Thornton, J. Mechanistic approaches to the study of evolution: the functional synthesis. Nat Rev Genet 8 , 675–688 (2007). https://doi.org/10.1038/nrg2160

Download citation

Issue Date : September 2007

DOI : https://doi.org/10.1038/nrg2160

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Evolutionary cell biology (ecb): lessons, challenges, and opportunities for the integrative study of cell evolution.

  • Parsifal Fidelio Islas-Morales
  • Luis F Jiménez-García
  • Christian R Voolstra

Journal of Biosciences (2021)

The Influence of Higher-Order Epistasis on Biological Fitness Landscape Topography

  • Daniel M. Weinreich
  • Yinghong Lan
  • Robert B. Heckendorn

Journal of Statistical Physics (2018)

Whole genome engineering by synthesis

  • Zhouqing Luo
  • Junbiao Dai

Science China Life Sciences (2018)

Molecular evolution between chemistry and biology

  • Peter Schuster

European Biophysics Journal (2018)

Strain-dependent mutational effects for Pepino mosaic virus in a natural host

  • Julia Minicka
  • Santiago F. Elena
  • Beata Hasiów-Jaroszewska

BMC Evolutionary Biology (2017)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

mechanistic hypothesis biology

SEP home page

  • Table of Contents
  • Random Entry
  • Chronological
  • Editorial Information
  • About the SEP
  • Editorial Board
  • How to Cite the SEP
  • Special Characters
  • Advanced Tools
  • Support the SEP
  • PDFs for SEP Friends
  • Make a Donation
  • SEPIA for Libraries
  • Entry Contents

Bibliography

Academic tools.

  • Friends PDF Preview
  • Author and Citation Info
  • Back to Top

Mechanisms in Science

Around the turn of the twenty-first century, what has come to be called the new mechanical philosophy (or, for brevity, the new mechanism ) emerged as a framework for thinking about the philosophical assumptions underlying many areas of science, especially in sciences such as biology, neuroscience, and psychology. In this entry, we introduce and summarize the distinctive features of this framework, and we discuss how it addresses a range of classic issues in the philosophy of science, including explanation, metaphysics, the relations between scientific disciplines, and the process of scientific discovery. For each of these issues, we show how the mechanistic framework has reoriented philosophical work, what the new mechanism has contributed to the discussion, and what remains to be done.

1. The Rise of the New Mechanism

2.1.1 producing, underlying, and maintaining, 2.1.2 regularity, 2.3.1 conserved quantity accounts, 2.3.2 mechanistic accounts, 2.3.3 activity-based accounts, 2.3.4 counterfactual accounts, 2.4.1 organization and aggregativity, 2.4.2 varieties of organization, 2.4.3 modularity, 2.4.4 jointness, 2.4.5 levels, 2.4.6 stable and ephemeral mechanisms, 2.5.1 what mechanisms are not, 2.5.2 what are not mechanisms, 2.6 philosophical work to be done, 3.1 etiological and constitutive explanations, 3.2 constitutive relevance, 3.3.1 characterizing completeness, 3.3.2 abstraction and idealization, 3.4 philosophical work to be done, 4.1 token mechanisms, type mechanisms, and laws, 4.2 mechanisms, levels, and emergence, 4.3 mechanisms and realization, 4.4 mechanisms and natural kinds, 4.5 mechanisms and functions, 4.6 philosophical work to be done, 5.1 theory reduction, 5.2.1 integrative pluralism, 5.3 philosophical work to be done, 6.1 discovery via a-ha moments, 6.2 discovery via strategies, 6.3 mechanistic evidence in medical discoveries, 6.4 philosophical work to be done, 7. conclusion, other internet resources, related entries.

Twentieth century philosophy of science was largely dominated by logical empiricism. More a framework for doing philosophy of science than any coherent set of doctrines, logical empiricism addressed a range of issues in philosophy of science through the lens of the logical and mathematical structures constitutive of scientific thought and practice (see the entry on logical empiricism ). Logical empiricism tended, by and large, to focus on abstract, epistemic features of science, with little attention to scientific practice. Physics was the dominant exemplar.

The new mechanical philosophy emerged around the turn of the twenty-first century as a new framework for thinking about the philosophy of science. The philosophers who developed this framework were, by comparison with the logical empiricists, practitioners as well of the history of science and tended, by and large, to focus on the biological, rather than physical, sciences. Many new mechanists developed their framework explicitly as a successor to logical empiricist treatments of causation, levels, explanation, laws of nature, reduction, and discovery.

As with logical empiricism, the new mechanical philosophy is less a systematic and coherent set of doctrines than it is an orientation to the subject matter of the philosophy of science. The approach emerged as philosophers and historians of science began to break from the once-standard practice of reconstructing scientific inference with the tools of logic and, instead, to embrace detailed investigation of actual episodes from the history of science. The main tenets of logical empiricism had been under intense criticism for decades, and a new era of historically informed philosophy of science had taken hold through the works of, e.g., Kuhn (1962), Laudan (1977), and Lakatos (1977). To many such scholars raised in this post-logical empiricist milieu, it appeared that much of the practice of contemporary science (both in the laboratory and in print) was driven by the search for mechanisms, that many of the grand achievements in the history of science were discoveries of mechanisms, and that more traditional philosophy of science, for whatever reason, had failed to appreciate this central feature of the scientific worldview.

Aspects of the new mechanical philosophy began to emerge in the late 1960s. Fodor (1968), for example, contrasted mechanistic explanations (dealing with parts and their law-like interactions) with functional explanations in psychology. Wimsatt (1972a, 1976), building on the work of Simon (1962) and Kaufman (1971), argued repeatedly that the abstract and idealized structures of logical empiricism were ill-suited to understanding how scientists discover and explain complex systems at multiple levels of organization. Cummins (1975) provided an account of functional analyses, characterizing a function as a contribution a component part makes to the overall capacity of some system that includes that component. Salmon (1984, 1989) argued that empiricist views of scientific explanation are fundamentally flawed because they neglect causal mechanisms. Cartwright (1989) argued that the logical empiricist conception of a law of nature is, in fact, a philosophical fiction used to describe the search for capacities and nomological machines.

These strands began to coalesce into an over-arching perspective in the 1990s. The earliest clear statement of the new mechanism was Bechtel and Richardson’s (2010 [1993]) Discovering Complexity . They self-consciously put aside logical empiricist concerns with theory reduction and focused instead on the process by which scientists discover mechanisms (see Section 6 below). Soon after, Glennan argued that mechanisms are the secret connexion Hume sought between cause and effect (1996), a thesis related to and partly inspired by Cartwright’s focus on capacities and nomological machines (Glennan 1997). Likewise, Thagard’s How Scientists Explain Disease centered the search for causes and mechanisms in medicine (Thagard 2000; see also Section 6 below). Machamer, Darden and Craver’s “Thinking about Mechanisms” (Machamer, Darden, and Craver 2000; familiarly known as “MDC”) drew these strands together and became for many the lightening rod of the new mechanist perspective. MDC suggested that the philosophy of biology, and perhaps the philosophy of science more generally, should be restructured around the fundamental idea that many scientists organize their work around the search for mechanisms.

2. The Concept of a Mechanism

The term “mechanism” emerged in the seventeenth century and derived from Greek and Latin terms for “machine” (Dijksterhuis 1961). Descartes understood mechanics as the basic building block of the physical world; in Le Monde , he proposed to explain diverse phenomena in the natural world (such as planetary motion, the tides, the motion of the blood, and the properties of light) in terms of the conservation of inertial motion through contact action (see the entry on René Descartes ). Subsequently, the idea of mechanism has been transformed many times to reflect an evolving understanding of the basic causal forces in the world (besides conserved motion): e.g., attraction and repulsion (du Bois Reymond), conservation of energy (Helmholtz), gravitational attraction (Newton) (Boas 1952; Westfall 1971; see also entries on Hermann von Helmholtz and Isaac Newton ). The concept of mechanism has had an almost separate evolution in the history of the life sciences (Allen 2005; Des Chene 2001, 2005; Nicholson 2012), at times eschewing the metaphysical austerity embraced by Descartes and many early mechanists.

The new mechanists inherit the word “mechanism” from these antecedents, but, in their effort to capture how the term is used in contemporary science, have distanced themselves both from the idea that mechanisms are machines and especially from the austere metaphysical world picture in which all real change involves only one or a limited set of fundamental activities or forces (cf. Andersen 2014a,b).

Mechanists have generally eschewed the effort to spell out necessary and sufficient conditions for something to be a mechanism. Instead, they offer qualitative descriptions designed to capture the way scientists use the term and deploy the concept in their experimental and inferential practices.

Three characterizations are most commonly cited:

  • MDC: “Mechanisms are entities and activities organized such that they are productive of regular changes from start or set-up to finish or termination conditions” (2000: 3).
  • Glennan: “A mechanism for a behavior is a complex system that produces that behavior by the interaction of a number of parts, where the interaction between parts can be characterized by direct, invariant, change-relating generalizations” (2002: S344).
  • Bechtel and Abrahamsen: “A mechanism is a structure performing a function in virtue of its component parts, component operations, and their organization. The orchestrated functioning of the mechanism is responsible for one or more phenomena” (2005: 423).

Each of these characterizations contains four basic features: (1) a phenomenon, (2) parts, (3) causings, and (4) organization. We consider each of these in detail below.

A useful canonical visual representation of a mechanism underlying a phenomenon is shown in Figure 1 (from Craver 2007). At the top is the phenomenon, some system S engaged in behavior ψ. This is the behavior of the mechanism as a whole. Beneath it are the parts (the X s) and their activities (the φs) organized together. The dotted roughly-vertical lines reflect the fact that the parts and activities are contained within, are components of, the mechanism engaged in this behavior. Thus represented, mechanisms are decompositional in the sense that the behavior of the system as a whole can be broken down into organized interactions among the activities of the parts.

missing text, please inform

Figure 1. A visual representation of a mechanism (adapted from Craver 2007).

In the early literature, these different characterizations were often treated as competitors. Tabery (2004) argued instead that they reflect different, and complementary, emphases and intellectual orientations. Many mechanists have adopted this ecumenical stance. For example, Illari and Williamson offer a “consensus concept” of mechanism:

A mechanism for a phenomenon consists of entities and activities organized in such a way that they are responsible for the phenomenon. (2012: 120)

Likewise, Glennan refers to “minimal mechanism”:

A mechanism for a phenomenon consists of entities (or parts) whose activities and interactions are organized in such a way that they produce the phenomenon. (Glennan forthcoming: Ch. 2)

These ecumenical characterizations each include the four basic elements and are designed to make the characterization more inclusive. MDC’s insistence on the regularity of mechanisms is abandoned, for example, to accommodate mechanisms that work only once or that work irregularly (Skipper and Milstein 2005; Bogen 2005; see also Section 2.1.2 below). Bechtel and Abrahamsen’s emphasis on the “functions” is abandoned to accommodate mechanisms that serve no end and to distance mechanism from this loaded term so often opposed to mechanism (although see Craver 2001a; Garson 2013; Maley and Piccinini forthcoming; see also Section 4.5 below).

These ecumenical characterizations intentionally downplay the fact that the term “mechanism” is used differently in different scientific and philosophical contexts (see Levy 2013 and Andersen 2014a,b for alternative overviews of the differences). Indeed, much of the progress in the early years involved learning to recognize the many ways that the term “mechanism” can be used and the many commitments that can be undertaken in its name. (For still other characterizations of mechanism, see Woodward (2002), Fagan (2012), Nicholson (2012), and Garson (2013)). Taking these ecumenical views as a starting point, we now consider the four basic components: 1) the phenomenon, 2) parts, 3) causings, and 4) organization.

2.1 Phenomenon

The phenomenon is the behavior of the mechanism as a whole. All mechanisms are mechanisms of some phenomenon (Kauffman 1971; Glennan 1996, 2002). The mechanism of protein synthesis synthesizes proteins. The mechanism of the action potential generates action potentials. The boundaries of a mechanism—what is in the mechanism and what is not—are fixed by reference to the phenomenon that the mechanism explains. The components in a mechanism are components in virtue of being relevant to the phenomenon.

MDC (2000) describe mechanisms as working from start- or set-up conditions to termination conditions. They insist that it is impoverished to describe the phenomenon as an input-output relation because there are often many such inputs and outputs from a mechanism and because central features of a phenomenon might be neither inputs nor outputs (but rather details about how the phenomenon unfolds over time). Darden, appealing to the example of protein synthesis, often associates the phenomenon with the end-state: the protein (Darden 2006). Craver (2007), following Cummins (1975) and Cartwright (1989), often speaks of the phenomenon roughly as a capacity or behavior of the mechanism as a whole.

New mechanists speak variously of the mechanism as producing, underlying, or maintaining the phenomenon (Craver and Darden 2013). The language of production is best applied to mechanisms conceived as a causal sequence terminating in some end-product: as when a virus produces symptoms via a disease mechanism or an enzyme phosphorylates a substrate. In such cases, the phenomenon might be an object (the production of a protein), a state of affairs (being phosphorylated), or an activity or event (such as digestion). For many physiological mechanisms, in contrast, it is more appropriate to say that the mechanism underlies the phenomenon. The mechanism of the action potential or of working memory, for example, underlies the phenomenon, here characteristically understood as a capacity or behavior of the mechanism as a whole. Finally, a mechanism might maintain a phenomenon, as when homeostatic mechanisms hold body temperature within tightly circumscribed boundaries. In such cases, the phenomenon is a state of affairs, or perhaps a range of states of affairs, that is held in place by the mechanism. These ways of talking can in many cases be inter-translated (e.g., the product is produced, the production has an underlying mechanism, and the state of affairs is maintained by an underlying mechanism). Yet clearly confusion can arise from mixing these ways of talking.

Must the relationship between the mechanism and the phenomenon be regular? This is an area of active discussion (DesAutels 2011; Andersen 2011, 2014a,b; Krickel 2014). MDC stipulate that mechanisms are regular in that they work “always or for the most part in the same ways under the same conditions” (2000: 3). Some have understood this (incorrectly in our view) as asserting that there are no mechanisms that work only once, or that a mechanism must work significantly more than once in order to count as a mechanism.

Some argue that mechanisms have to be regular in this factual sense (Andersen 2014a,b); i.e., repeated on many occasions (see Leuridan 2010). This view would seem to require a somewhat arbitrary cut-off point in degree of regularity between things that truly count as mechanisms and those that do not. Some mechanists (Bogen 2005; Glennan 2009) argue that there is no difficulty applying the term “mechanism” to one-off causal sequences, as when an historian speaks of the mechanism that gave rise to World War I. Other mechanists argue that the type-token distinction is too crude a dichotomy to capture the many levels of abstraction at which mechanism types and tokens might be characterized (Darden 1991).

It is possible, however, to read the MDC statement as asserting, not a factual kind of regularity, but as a counterfactual kind of near-determinism: were all the conditions the same, then the mechanism would likely produce the same phenomenon, where “likely” accommodates mechanisms with stochastic elements.

While the MDC account leaves open the possibility that some mechanisms are stochastic, it clearly rules out mechanisms that usually fail to produce their phenomena. Skipper and Millstein (2005) press this point to argue that the MDC account cannot accommodate the idea that natural selection is a mechanism. If, as Gould (1990) argued, one could not reproduce the history of life by rewinding the tapes and letting things play forward again, then natural selection would not be an MDC mechanism (see also Section 2.6 below). It is unclear why MDC would allow for the possibility of stochastic mechanisms and rule out, by definition, the possibility that they might fail more often than they work. Whether any biological mechanisms are truly irregular in this sense (i.e., all the causally relevant factors are the same but the product of the mechanism differs) is a separate question from whether they are mechanisms simpliciter (see Bogen 2005; Machamer 2004; Steel 2008 develops a stochastic account of mechanisms).

Krickel (2014) reviews the many different ways of unpacking the relevant notion of regularity (see also Andersen 2012). Her favored solution, “reverse regularity,” holds that there must be a generalization to the effect that, typically, when the phenomenon occurs, the mechanism was acting.

Mechanists have struggled to find a concise way to express the idea of parthood required of the components in a mechanism. The project is to develop an account that is both sufficiently permissive to include the paradigmatic mechanisms from diverse areas of science and yet not vacuous.

Formal mereologies are difficult to apply to the material parts of biological mechanisms. Axioms of mereology, such as reflexivity (everything is a part of itself) and unrestricted composition (any two things form a whole) do not apply in standard biological uses of the “part” concept.

Glennan (1996) recognized the difficulty of defining parthood very early on. His proposal:

The parts of mechanisms must have a kind of robustness and reality apart from their place within that mechanism. It should in principle be possible to take the part out of the mechanism and consider its properties in another context. (1996: 53)

Yet even this is perhaps too strong, given that some parts of a mechanism might become unstable when removed from their mechanistic context. Later, Glennan (2002: S345) says that the properties of a part must be stable in the absence of interventions, or that parts must be stable enough to be called objects. This notion is perhaps too strong to accommodate the more ephemeral parts of some biochemical mechanisms or of the mechanisms of natural selection (Skipper and Millstein 2005; but see Illari and Williamson 2010).

2.3 Causings

Mechanists have disagreed with one another about how to understand the cause in causal mechanism. New mechanists have in general been at pains both (1) to liberate the relevant causal notion from any overly austere view that restricts causation to only a small class of phenomena (such as collisions, attraction/repulsion, or energy conservation), and (2) to distance themselves from the Humean, regularist conception of causation common among logical empiricists (see also the entry on the the metaphysics of causation ). Four ways of unpacking the cause in causal mechanism have been discussed: conserved quantity accounts, mechanistic accounts, activities accounts, and counterfactual accounts. (It should be noted that some mechanists have evolved in their thinking about causation.)

According to transmission accounts, causation involves the transmission and propagation of marks or conserved quantities (Salmon 1984, 1994; Dowe 1992). The most influential form of this view holds that two causal processes causally interact when they intersect in space-time and exchange some amount of a conserved quantity, such as mass. On this view, causation is local (the processes must intersect) and singular (it is fully instantiated in particular causal processes), though the account relies upon laws of conservation (Hitchcock 1995). Although this view inspired many of the new mechanists, and although it shares their commitment to looking toward science for an account of causation, it has generally been rejected by new mechanists (though see Millstein 2006; Roe 2014).

This view has been unpopular in part because it has little direct application in nonfundamental sciences, such as biology. The causal claims biologists make usually don’t involve explicit reference to conserved quantities (even if they presuppose such notions fundamentally) (Glennan 2002; Craver 2007). Furthermore, biological mechanisms often involve causation by omission, prevention, and double prevention (that is, when a mechanism works by removing a cause, preventing a cause, or inhibiting an inhibitor) (Schaffer 2000, 2004). Such forms of causal disconnection are ubiquitous in the special sciences.

Glennan (1996, 2009) sees causation (at least non-fundamental causation) as derivative from the concept of mechanism: causal claims are claims about the existence of a mechanism. The truth-maker for a causal claim at one level of organization is a mechanism at a lower level. In short, mechanisms are the hidden connexion Hume sought between cause and effect. Like the Salmon-Dowe account, Glennan’s view is singular: particular mechanisms link particular causes and particular effects (Glennan forthcoming)

This view has been charged with circularity: the concept of mechanism ineliminably contains a causal element. However, Glennan replies that many accounts of causation (such as Woodward’s 2003 account, see Section 2.3.4 below) share this flaw. Furthermore, he argues that for at least all non-fundamental causes, a mechanisms clearly explains how a given cause produces its effect.

Whether the analysis succeeds depends on how one deals with the resulting regress (Craver 2007). As Glennan (2009) notes, the decomposition of causes into mechanisms might continue infinitely, in which case there is no point arguing about which notion is more fundamental, or the decomposition might ground out in some basic, lowest-level causal notion that is primitive and so not analyzable into other causal mechanisms. The latter option must confront the widely touted absence of causation in the theories of fundamental physics (Russell 1913); at very small size scales, classical conceptions of objects and properties no longer seem to apply, making it difficult to see what content is left to the idea that there are mechanisms at work (see also Teller 2010; Kuhlman and Glennan 2014).

Still other mechanists, such as Bogen (2005, 2008a) and Machamer (Machamer 2004), embrace an Anscombian, non-reductive view that causation should be understood in terms of productive activities (see also the entry on G.E.M. Anscombe ). Activities are kinds of causing, such as magnetic attraction and repulsion or hydrogen bonding. Defenders of activity-based accounts eschew the need to define the concept, relying on science to say what activities are and what features they might have. This view is a kind of causal minimalism (Godfrey-Smith 2010). Whether an activity occurs is not a matter of how frequently it occurs or whether it would occur always or for the most part in the same conditions (Bogen 2005).

This account has been criticized as vacuous because it fails to say what activities are (Psillos 2004), to account for the relationship of causal and explanatory relevance (Woodward 2002), and to mark an adequate distinction between activities and correlations (Psillos 2004), though see Bogen (2005, 2008a) for a response. Glennan (forthcoming) argues that these problems can be addressed by recognizing that activities in a mechanism at one level depend on lower-level mechanisms. (See also Persson 2010 for a criticism of activities based on their inability to handle cases of polygenic effects.)

Lastly, some new mechanists, particularly those interested in providing an account of scientific explanation, have gravitated toward a counterfactual view of causal relevance, and in particular, to the manipulationist view expressed in Woodward (2001, 2003) (see, e.g., Glennan 2002; Craver 2007). The central commitment of this view is that models of mechanisms describe variables that make a difference to the values of other variables in the model and to the phenomenon. Difference-making in this manipulationist sense is understood as a relationship between variables in which interventions on cause variables can be used to change the value of effect variables (see the entry on causation and manipulability ).

Unlike the views discussed above, this way of thinking about causation provides a ready analysis of explanatory relevance that comports well with the methods for testing causal claims. Roughly, one variable is causally relevant to a second when there exists an ideal intervention on the first that changes the value of the second via the change induced on the first. The view readily accommodates omissions, preventions, and double preventions—situations that have traditionally proven troublesome for production-type accounts of causation. In short, the claim that C causes E requires only that ideal interventions on C can be used to change the value of E , not that C and E are physically connected to one another. Finally, this view provides some tools for accommodating higher-level causal relations and the non-accidental laws of biology. On the other hand, the counterfactual account is non-reductive (like the mechanistic view), and it inherits challenges faced by other counterfactual views, such as pre-emption and over-determination which are common in biological mechanisms (see the entry on counterfactual theories of causation ).

2.4 Organization

The characteristic organization of mechanisms is itself the subject of considerable discussion.

Wimsatt (1997) contrasts mechanistic organization with aggregation, a distinction that mechanists have used to articulate how the parts of a mechanism are organized together to form a whole (see Craver 2001b). Aggregate properties are properties of wholes that are simple sums of the properties of their parts. In aggregates, the parts can be rearranged and intersubstituted for one another without changing the property or behavior of the whole, the whole can be taken apart and put back together without disrupting the property or behavior of the whole, and the property of the whole changes only linearly with the addition and removal of parts. These features of aggregates hold because organization is irrelevant to the property of the whole. Wimsatt thus conceives organization as non-aggregativity. He also describes it as a mechanistic form of emergence (see Section 4.2 below).

Mechanistic emergence is ubiquitous—truly aggregative properties are rare. Thus mechanists have tended to recognize a spectrum of organization, with aggregates at one end and highly organized mechanisms on the other. Indeed, many mechanisms studied by biologists involve parts and causings all across this spectrum. (For further discussion of mechanistic emergence in relationship to other varieties, see Richardson and Stephan 2007.)

Following Wimsatt, mechanists have detailed several kinds of organization characteristic of mechanisms. A canonical list includes both spatial and temporal organization. Spatial organization includes location, size, shape, position, and orientation; temporal organization includes the order, rate, and duration of the component activities. More recently, mechanists have emphasized organizational patterns in mechanisms as a whole. Bechtel, for example, discusses how mathematical models, and dynamical models in particular, are used to reveal complex temporal organization in interactive mechanisms (Bechtel 2006, 2011, 2013b). Some argue that dynamical models push beyond the limits of the mechanistic framework (e.g., Chemero and Silbestein 2008 and, at times, Bechtel himself; see Kaplan and Bechtel 2011). Others argue that dynamical models are, in fact, often merely descriptive (i.e., non-explanatory models) or, alternatively, that they are used to describe the temporal organization of mechanisms (Kaplan and Bechtel 2011; Kaplan 2012).

Mechanists have also recently borrowed from Alon’s (2006; Milo et al. 2002) work on network motifs, repeated patterns in causal networks, to expand the vocabulary for thinking about abstract patterns of organization (Levy 2014; Levy and Bechtel 2012). Understanding how parts compose wholes is likely to be a growth area in the future of the mechanistic framework. (For some other recent additions, see Kuorikoski and Ylikoski 2013; Kuhlmann 2011; Glennan forthcoming.)

Woodward’s (2001, 2002, 2011, 2014) counterfactual definition of a mechanism (which is indirectly specified via an account of mechanistic models), as well as a descendant elaborated by Menzies (2012), require that models of mechanisms be modular. This means, roughly, that it should be physically possible to intervene on a putative cause variable in a mechanism without disrupting the functional relationships among the other variables in the mechanism. In terms of structural equation models in particular, this means that one should be able to replace the right-hand side of an equation in the model with a particular value (i.e., set the left-hand variable to a value) without needing to change any of the other equations in the model. This is intended to formally capture the sense in which mechanism is composed of separable, interacting parts. For arguments in favor of a modularity condition on mechanistic models see Menzies (2012).

Steel (2008) appeals to a somewhat weaker form of modularity in his probabilistic analysis of mechanisms—one that follows directly from Simon’s (1996 [1962]) idea of nearly decomposable systems. On Simon’s view, the parts of a mechanism have more and stronger causal relations with other components in the mechanism than they do with items outside the mechanism. This gives mechanisms (and parts of mechanisms) a kind of “independence” or “objecthood” defined ultimately in terms of the intensity of interaction among components. Grush (2003), following Haugeland (1998), develops an idea of modularity in terms of the bandwidth of interaction, where modules are high-bandwidth in their internal interactions and low-bandwidth in their external interactions. On this view, modularity is not an all-or-none proposition but a matter of degree; mechanisms are only nearly decomposable. Craver (2007) argues that such a generic notion fails to account for the relevance of different causal interactions for different mechanistic decompositions; what counts as a part of a mechanism can only be defined relative to some prior decision about what one takes the mechanism to be doing. For criticisms of modularity, see Mitchell (2005) and Cartwright (2001, 2002).

Fagan (2012, 2013) emphasizes the interdependent relationship between parts of a mechanism. Components in a mechanism, she points out, often form a more complex unit by virtue of the individual properties that unite them—their “meshing properties”; the complex unit then figures into the mechanism’s behavior. This interdependent relationship—jointness—is exemplified by the lock-and-key model of enzyme action. Fagan applies this notion to research on stem cells (Fagan 2013) but argues that it is a general feature of experimental biology (Fagan 2012).

Many mechanists emphasize the hierarchical organization of mechanisms and the multilevel structure of theories in the special sciences (see especially Craver 2007, Ch. 5). Antecedents of the new mechanism focused almost exclusively on etiological, causal relations. However, the new emphasis on mechanisms in biology and the special sciences demanded an analysis of mechanistic relations across levels of organization.

From a mechanistic perspective, levels are not monolithic divides in the furniture of the universe (as represented by Oppenheim and Putnam 1958), nor are they fundamentally a matter of size or the exclusivity of causal interactions within a level (Wimsatt 1976). Rather, levels of mechanisms are defined locally within a multilevel mechanism: one item is at a lower level of mechanisms than another when the first item is a part of the second and when the first item is organized (spatially, temporally, and actively) with the other components such that together they realize the second item. Thus, the mechanism of spatial memory has multiple levels, some of which include organs such as the hippocampus generating a spatial map, some of which involve the cellular interactions that underlie map generation, and some of which involve the molecular mechanisms that underlie those cellular interactions (Craver 2007). For more on levels, see Section 4.2 below.

Finally, mechanists have found it necessary to distinguish between stable mechanisms, which rely fundamentally upon the more or less fixed arrangement of parts and activities, and ephemeral mechanisms, which involve a process evolving through time without fixed spatial and temporal arrangement (Glennan 2009). The time-keeping mechanism in a clock, for example, is a relatively stable assemblage of components in relatively fixed locations that work the same way, with the same organizational features, each time it works. Ephemeral mechanisms, in contrast, involve a much looser kind of organization: items still interact in space and time, but they do not do so in virtue of robust, stable structures. Many chemical mechanisms in a cell are like that (Richardson and Stephan 2007). Ephemeral mechanisms are surely a primary focus of historical sciences, such as archaeology, history, and evolutionary biology (Glennan 2009).

2.5 What Mechanisms Are Not and What Are Not Mechanisms

The term “mechanism” has been used in many different ways to express many different ideas. The new mechanists’ appropriation of the term is thus likely to cause unhelpful associations, and their liberalization of the term is likely to raise worries that the notion of mechanism has thereby been trivialized (see, e.g., Moss 2012 and Nicholson 2012). Here, we first distinguish the new mechanism from other doctrines with which it shares both name and family resemblance. We then discuss some things to which the new use of the term “mechanism” does not apply.

New mechanists have explicitly eschewed the following associations with the term “mechanism”:

  • Mechanisms are not necessarily deterministic . Mechanisms might be stochastic if, for example, they are composed of stochastic activities (Bogen 2005, 2008a), or, in a more mundane sense (i.e., one consistent with determinism), because it is always possible for one or more factors to interfere with the working of a mechanism; one of the parts might be broken, or an unexpected preventer might interfere with the operation of a mechanism. The truth or falsity of determinism, and its relevance to understanding the special sciences, is an independent issue from the question of whether something is a mechanism.
  • Appeal to mechanisms is not necessarily reductionistic . Mechanisms are often described as multi-level, with activities at different levels being equally essential to how a mechanism works. Mechanistic explanations might look, up, down, or around depending on the choice of an explanandum and the presuppositions of the explanatory context (Bechtel 2009a). Mechanists might be reductionists or anti-reductionists. That said, many mechanists opt for some form of explanatory anti-reductionism, emphasizing the importance of multilevel and upward-looking explanations, without rejecting the central ideas that motivate a broadly physicalist world-picture (e.g., McCauley and Bechtel 2001; Craver 2007). (For further discussion, see Theurer 2013; see also Sections 3.1 and 5 and the entry on reductionism in biology .)
  • Not all mechanisms are machines . Machines are human-made contrivances with each part added and organized by a designer to perform a function; biological and social mechanisms, in contrast, are products of evolution, broadly construed (Darden 2006), and so display ornate forms of organization in comparison with contrivances. One machine might contain multiple mechanisms (a car, for example, has mechanisms for braking, propulsion, playing music and climate control). Machines are also capable of being both active and passive (a stopped clock is still a machine); mechanisms, in contrast, have a productive aspect and are always doing something.
  • Mechanisms are not necessarily sequential or linear . Mechanisms can have feedback loops and cycles wherein the output of the mechanism or components in turn influences the input of the mechanism or components in a subsequent iteration (Bechtel 2011). Also, the interactions among components in a mechanism need not be describable by a linear equation.
  • Mechanisms are not necessarily localizable (Bechtel and Richardson 2010 [1993]). Components of mechanisms might be widely distributed (as are many brain mechanisms) and might violate our intuitive or tutored sense of the boundaries of objects (as an action potential violates the cell boundary). The assumption of localization is often an important heuristic in the search for mechanisms; however, this heuristic often must be abandoned as the mechanism’s organization reveals itself.
  • Mechanisms are not limited to push-pull dynamics . Descartes’ mechanism had this feature, but (as noted above) the new mechanism explicitly liberalizes the notion to account for other kinds of causing.
  • Mechanisms are not just fictions/metaphors . When a scientist says that there is a mechanism that makes proteins in living organisms, she is not just using a machine metaphor; rather, she is saying that there are in fact parts and activities organized in living organisms such that they produce proteins.

One might object that there’s nothing left of mechanism once it sheds these historical associations. One might suspect that it has been trivialized (Dupré 2013).

The idea of mechanism is a central part of the explanatory ideal of understanding the world by learning its causal structure. The history of science contains many other conceptions of scientific explanation and understanding that are at odds with this commitment. Some have held that the world should be understood in terms of divine motives. Some have held that natural phenomena should be understood teleologically. Others have been convinced that understanding the natural world is nothing more than being able to predict its behavior. Commitment to mechanism as a framework concept is commitment to something distinct from and, for many, exclusive of, these alternative conceptions. If this appears trivial, rather than a central achievement in the history of science, it is because the mechanistic perspective now so thoroughly dominates our scientific worldview.

Yet there are many ways of organizing phenomena besides revealing mechanisms. Some scientists are concerned with physical structures and their spatial relations without regard to how they work: an anatomist might be interested in the spatial organization of parts within the body with minimal interest in how those parts articulate together to do something. Many scientists build predictive models of systems without any pretense that these models in fact reveal the causal structures by which the systems work. Some scientists are concerned with taxonomy, sorting like with like without regard to how the sorted items came about or how they work. Finally, in many areas of science, there is a widely recognized and practically significant distinction between knowing that C (e.g., smoking) is a cause of E (lung cancer) and knowing how C causes E . This is not so much an ontological difference as it is a difference in the grain with which one seeks to understand a system’s causal structure. In short, there are many framework concepts in science, and not all of them can be assimilated to mechanisms.

But what, the critic might push further, does not count as a mechanism? Here are some contrast classes:

  • Entities (or objects ) are not mechanisms. Mechanisms do things. If an object is not doing anything (i.e., if there is no phenomenon), then it is not a mechanism.
  • Correlations are not mechanisms. Mechanisms explain at least many correlations, and many correlations can be used to characterize causal or mechanistic relations, but correlations themselves are not mechanistic. The same can be said of mere temporal sequences of events.
  • Inferences , reasons , and arguments are not mechanisms. Though there are mechanisms of inference and reasoning, what makes something an inference or a reason is logical relation and not (merely) a causal relationship between premise and conclusion.
  • Symmetries are not mechanisms. Many kind of symmetry are of fundamental importance in different areas of physics (e.g., translational symmetries, rotational symmetries). These are features of physical systems that are highly general facts or assumptions, not mechanisms.
  • Fundamental laws and fundamental causal relations are not mechanisms. If a law or causal relation is fundamental, then (by definition) there is no mechanism for it.
  • Relations of logical and mathematical necessity are not mechanisms. Such truths hold in all possible worlds and so do not depend for their truth on facts about the causal structure of this world.

This is not an exhaustive list of non-mechanisms or non-mechanistic framework concepts. Yet it demonstrates that even the liberalized concept of mechanism is neither vacuous nor trivial.

Much of the early new mechanical philosophy has focused on the special sciences, such as neuroscience and molecular biology. In the years since, philosophers have extended the mechanistic framework to other scientific disciplines, such as cell biology (Bechtel 2006), cognitive science (Bechtel 2008; Thagard 2006), neuroeconomics (Craver and Alexandrova 2008), organic chemistry (Ramsey 2008), physics (Teller 2010), astrophysics (Illari and Williamson 2012), behavior genetics (Tabery 2014a), and phylogenetics (Matthews forthcoming). Philosophers continue to test the limits of this framework, with the expectation that alternative organizing frameworks might play central roles in other sciences. For example, a debate has emerged in the philosophy of biology over whether or not natural selection is a mechanism (see, for example, Skipper and Millstein 2005; Baker 2005; Barros 2008; Illari and Williamson 2012; Havstad 2011; and Matthewson and Calcott 2011). Similar debates have emerged concerning mechanistic explanation in cognitive science (Bechtel 2008; Piccinini and Craver 2011; Weiskopf 2011; Povich forthcoming).

One area that has received particular attention is the effort to understand computational mechanisms. On some accounts, computational mechanisms form a proper subclass of mechanisms that can be defined explicitly in terms of the kinds of entities, properties, and activities involved in mechanisms in that class (Piccinini 2007; Milkowski 2013). According to this view, computational mechanisms are mechanisms that have the function to manipulate medium independent vehicles in accordance with a general rule that applies to all vehicles and depends on the inputs for its application (Piccinini and Scarantino 2011). Digital computers are distinctive in that their vehicles are digits (Piccinini 2007). Proponents of this account hope to demarcate computing mechanisms from non-computing mechanisms by appeal to the distinctive components proprietary to computing mechanisms. This view contrasts both with a semantic view, according to which computation is essentially a matter of manipulating symbols or representations, and with perspectivalist views, according to which whether a mechanism counts as computing is a matter of whether it is so described (Churchland 1986; Churchland and Sejnowski 1992; Shagrir 2010).

Philosophers of the social sciences have also emphasized and debated the importance of mechanistic knowledge (e.g., Elster 1989; for a useful review of these connections, see Hedström and Ylikoski 2010). In that context, appeals to mechanisms are intended to remedy the relative uninformativeness of social (or macro-level) explanations of social phenomena (such as widespread norms, persistent inequalities, network and institutional structures) by insisting that these explanations ultimately be grounded in mechanistic details about individual agents and actors, their desires and motivations, and, importantly, their relations to one another. The emphasis on relations among actors distances this mechanistic view from methodological individualism (see the entry on methodological individualism ). Mechanists in the social sciences have also tended to shy away from grand, overarching theories and toward more local explanations: scientific knowledge grows by adding items to a toolbox of mechanisms and showing how items from that toolbox can be combined to provide an explanation for a particular phenomenon. Frederica Russo (2009) discusses a number of strategies for modeling social mechanisms (see also Little 1991, 1998; Hedström 2005; Hedström and Swedberg 1998).

3. Explanation: From Formal Analyses to Material Structures

The covering-law model of explanation was a centerpiece of the logical empiricist conception of science. According to that model, explanations are arguments showing that the event to be explained (the explanandum event) was to have been expected on the basis of laws of nature and the antecedent and boundary conditions (the explanans ). For advocates of the covering-law model, the philosophical problem of explanation is thus largely a matter of analyzing the formal structure of explanatory arguments (Hempel and Oppenheim 1948; Hempel 1965). A rainbow, for example, is explained under the covering-law model by reference to laws of reflection and refraction alongside conditions concerning the position of the sun and the nature of light, the position of the raindrops, and the position of the person seeing the rainbow. The description of the rainbow is the conclusion of a deductive argument with law statements and descriptions of conditions as premises, and so the rainbow was to be expected in light of knowledge of the laws and conditions.

Mechanists, in contrast, insist explanation is a matter of elucidating the causal structures that produce, underlie, or maintain the phenomenon of interest. For mechanists, the philosophical problem is largely about characterizing or describing the worldly or ontic structures to which explanatory models (including arguments) must refer if they are to count as genuinely explanatory. A rainbow, for the mechanist, is explained by situating that phenomenon in the causal structure of the world; the explanation is an account of how the phenomenon was produced by entities (like rain drops and eyeballs) with particular properties (like shapes and refractive indices) that causally interact with light propagating from the sun. Mechanists typically distinguish several ways of situating a phenomenon within the causal structure of the world.

Most mechanists recognize two main aspects of mechanistic explanation: etiological and constitutive. Salmon (1984) describes them as two different ways of situating an explanandum phenomenon in the causal nexus (see also Craver 2001b; Glennan 2009). Etiological explanations reveal the causal history of the explanandum phenomenon, as when one says a virus explains a disease. Constitutive explanations, in contrast, explain a phenomenon by describing the mechanism that underlies it, as when one says brain regions, muscles, and joints explain reaching.

Philosophical arguments against the covering law model often focused on its inability to deal with causal, etiological explanations. The model failed to deliver the right verdict on a variety of problem cases precisely because it attempted to provide an account of explanation without any explicit mention of causation (Bromberger 1966; Salmon 1984; Scriven 1959)

New mechanists extend these kinds of criticism to the covering law model of intertheoretic, micro-reduction. According to the covering law model of reductive explanation, a theory about parts reduces, and so explains, a theory about wholes when it is possible to derive the second from the first given bridge laws to connect the two (see Nagel 1961; Schaffner 1993).

Some mechanists argue that the covering law model of constitutive explanation has problems analogous to those that beset the covering-law model of etiological explanations. Action potentials cannot be explained by mere temporal sequences of events utterly irrelevant to the phenomenon, but one can derive a description of the action potential from descriptions of such irrelevant phenomena. Action potentials cannot be explained by mere patterns of correlation that are not indicative of an underlying causal relation. Irrelevant byproducts of a mechanism might be correlated with the behavior of the mechanism, even perfectly correlated such that one could form bridge laws between levels, but would not thereby explain the relationship. Merely finding a neural correlate of consciousness, for example, would not, and is not taken by anyone to, constitute an explanation of consciousness. So mechanists argue that micro-reductive explanations must satisfy causal constraints just as surely as etiological explanations must (Craver 2007).

The covering law model also fails to distinguish models that merely re-describe the phenomenon in general terms from explanations that, in addition to predicting aspects of the phenomenon, reveal the mechanisms that produce it (Craver 2006; Kaplan and Craver 2011; but see Weiskopf 2011). For example, Snell’s law allows one to predict how light bends when passing from one medium to another, but it does not explain why the light bends. New mechanists also argue that the covering law model fails to distinguish predictively adequate but fictional models from explanatory models. Finally, mechanists argue that the intertheoretic model of reduction fails to capture an important dimension of explanatory quality: depth. An implication of the covering law model is that any true law statements that allow one to derive the explanandum law (with suitable corrections and assumptions) will count as a complete explanation. Yet it seems one can deepen an explanation by opening black boxes and revealing how things work down to whatever level one takes as relatively fundamental for the purposes at hand. Such criticisms suggest that the covering-law model of constitutive explanation is too weak to capture the norms of explanation in the special sciences.

Other mechanists have argued that the covering law model is too strong. Philosophers of biology have long argued that there are no laws of the sort the logical empiricist described in biology and other special sciences (Beatty 1995; Mitchell 1997, 2000; Woodward 2001). One might conclude from this that there are no explanations in biology (Rosenberg 1985), but such a radical conclusion is difficult to square with obvious advances in understanding, e.g., protein synthesis, action potentials, cell signaling, and a host of other biological phenomena. In such cases, one finds that scientists appeal to mechanisms to do the explanatory work, even in cases where nothing resembling a law appears to be available.

With increased attention to constitutive explanation, mechanists realized the need for an account of constitutive relevance, a principal for sorting relevant from irrelevant factors in a mechanism (Craver 2007; Ylikoski 2013). A system ( S ) exhibiting phenomenon (ψ) is composed of many different entities ( x ), with various properties, engaging in myriad activities (φ) organized together (see Figure 1 above in Section 2). One central research problem is to say which of these entities, activities, and organizational features contribute to the phenomenon and which do not. In a sense, this is a challenge of defining the boundaries of a mechanism: of saying what is and is not in the mechanism.

Three proposals have been considered. The first, the mutual manipulability account , understands constitutive relevance in terms of the experimental manipulations used to test interlevel relations. According to this account, if it can be shown (i) that the putative components are contained within S , (ii) that some ideal interventions on the putative component ( x ’s φ-ing) change the phenomenon ( S ’s ψ-ing), and (iii) that some ideal interventions on S ’s ψ-ing change x ’s φ-ing, that is sufficient to establish that x is a component in the mechanism. The notion of an ideal intervention in this account is explicitly indebted to and a proposed extension of Woodward’s theory of causal relevance to the constitutive domain (see Craver 2007; see also Kaplan 2012). A concern with the mutual manipulability account, though, is that it is best an epistemic guide to constitutive relevance, not an account of what constitutive relevance is (Couch 2011). The account offers, at best, a sufficient condition of relevance. Also, the notion of an “ideal” intervention, borrowed from Woodward’s account of causal relevance, cannot be applied straightforwardly to constitutive explanations. An ideal intervention on a system cannot intervene on both the independent and the dependent variable at the same time. However, when one intervenes to make S ψ (or prevent S from ψ-ing), one invariably also intervenes on the components of S ’s ψ-ing. And when one intervenes on the components of S ’s ψ-ing, one often intervenes on S ’s ψ-ing. Because x ’s φ-ing and S ’s ψ-ing are related as part to whole, they are not independent, and so require another way to think about ideal interventions (see Baumgartner 2010; 2013; Leuridan 2011; yet see Menzies 2012; Woodward 2014.

A second proposal offers a regularity account of constitutive relevance modeled on Mackie’s notion of understanding a cause as an INUS condition: an Insufficient but Non-redundant part of an Unnecessary but Sufficient condition for the effect in question (Mackie 1974; see also Cummins 1983). On this account, a constitutively relevant component is an insufficient but non-redundant part of an unnecessary but sufficient mechanism for a given phenomenon (Couch 2011; see also Harbecke 2010, 2014). Allow that any number of mechanisms might suffice to bring about S ’s ψ-ing; each possible sufficient mechanism is then unnecessary for ψ-ing. Each of these mechanisms is made of components, none of which is alone sufficient to produce the behavior of the mechanism as a whole, but each of which is necessary in the context of the mechanism for S to ψ. This account presupposes the idea of being necessary in context, and one might reasonably worry about sorting accidentally correlated X s from X s that in fact make a difference to S ’s ψ-ing.

A third approach to constitutive relevance dispenses with the interlevel framing enforced by the mutual manipulability account and attempts to analyze relevance using causal notions only. According to accounts of this sort, constitutive relevance is a kind of causal between-ness. If S ’s ψ-ing is understood as an input-output relationship of some sort, then mechanistic relevance could be understood as being a necessary link in the causal chain between the input and the output (see Harinen forthcoming; Menzies 2012). The putatively interlevel experiments in the mutual manipulability account can then be recast as different kinds of unilevel causal experiments. Romero (forthcoming) provides a helpful framing of these issues and offers the novel suggestion that putatively high-level interventions are in fact fat-handed interventions relative to their lower-level counterparts.

3.3 Mechanisms and Models

The philosophical literature on mechanisms also overlaps with the philosophical literature on scientific models (see the entry on models in science ). Here we distinguish mechanical models from models of mechanisms and we discuss varieties of non-mechanical models.

Glennan (2005) proposed a definition of a mechanical model as follows:

(MM) A mechanical model consists of (i) a description of the mechanism’s behavior (the behavioral description); and (ii) a description of the mechanism that accounts for that behavior (the mechanical description). (446)

Such models can be represented in many different ways (see also Giere 2004). They are evaluated in terms of their ability to predict the features of the phenomenon and in terms of the mapping between items in the model and the entities, activities, and organizational features in the mechanism (Glennan 2005: 17; Kaplan and Craver 2011). Glennan emphasizes that there is no hard line between complete and incomplete models; rather models are continually in the process of articulation and refinement. Whether a model is complete enough is determined by pragmatic considerations.

This last point is related to Darden’s distinction between mechanism schemas and mechanism sketches (Darden and Cain 1989; Darden 2002). In discovering a mechanism, it is often crucial to identify gaps that have to be filled in one’s model. While no model is ever complete in the absolute sense, some models have lacunae that must be filled before the model is complete enough

Mechanism schemas are abstract descriptions of mechanisms that can be filled in with details to yield a specific type or token mechanism. Thus, the schema:

DNA → RNA → Protein

can be filled in with a specific sequence of bases in DNA, its complement in RNA, and a corresponding amino acid sequence in the protein. The arrows can be filled in, showing how transcription and translation work. A mechanism sketch is an incomplete representation of a mechanism that specifies some of the relevant entities, activities, and organizational features but leaves gaps that cannot yet be filled. Black boxes, question marks, and filler-terms (such as “activate”, “cause”, or “inhibitor”) hold the place for some entity, activity or process yet to be discovered. The distinction between sketches and schemas is a matter of completeness: schemas are more complete than sketches in the sense that a sketch omits one or more stages of the mechanism that have to be understood if one really wants to solve one’s discovery problem.

Mechanists also emphasize the distinction between a how-possibly schema and a how-actually-enough schema (Craver and Darden 2013). A how-possibly schema describes how entities and activities might be organized to produce a phenomenon. A how possibly model is n hypothesis about how the mechanism works. Such models might be true (enough) or false. A true (enough) how-possibly model is (though we may not know it) also a how-actually (enough) model. A how-actually-enough schema describes how entities and activities are in fact organized to produce the phenomenon. The term “how-actually-enough” captures the idea that the requisite “accuracy” of a mechanistic model can vary considerably from one pragmatic context to another (Weisberg 2013).A false how possibly model is merely a how possibly model; just-so-stories are merely how possibly models (Dray 1957; Brandon 1985). Used in this way, the term “how possibly model” is similar to the term “hypothesis”: it is entertained as a possibility but not necessarily endorsed.

In contrast to mechanism schemas and sketches, some models of mechanisms work not by describing all of the parts, causal interactions, and organizational features, but rather by abstracting away from such potentially obfuscating details (Craver and Darden 2013; Strevens 2008; Levy and Bechtel 2012). In such cases, idealizing assumptions can be introduced to bring the relevant feature of the mechanism most clearly into view: infinite populations, frictionless planes, perfect geometrical shapes are presumed in order to strip the model of detail that does not matter for, or would only obstruct, the intended purposes of model.

Critics of the new mechanical philosophy have pushed on the importance of abstraction in science, drawing attention to the above discussions of completeness. The goals of completeness and accuracy are taken to conflict with the common practice of being satisfied with models that sacrifice detail and truth for clarity and generality (Strevens 2008; Woodward 2014). The normative distinction between a schema and a sketch, for example, seems to suggest that science progresses by moving from incomplete to complete models. And the distinction between how-possibly and how-actually-enough likewise seems to privilege accuracy over other goals of modeling, which often require distortion and falsity (see Wimsatt 2007; Weisberg 2007; Levy and Bechtel 2012; Batterman and Rice 2014; Chirimuuta 2014; Levy 2014).

Yet mechanists can surely allow that not all models of mechanisms are mechanical models or mechanism schemas. Often other sorts of model are useful for isolating central aspects of a mechanism’s functioning. Dynamical models, for example, can be used to characterize the temporal dynamics of a mechanism (Bechtel 2013a,b; Kaplan and Bechtel 2011). Network models can be used to characterize patterns of connectivity regardless of what units are connected and regardless of what kinds of connections one is particularly interested in characterizing (Hunneman 2010). Minimal models can be used to capture something fundamental about the dynamics of a broad class of mechanisms that share no entities and activities in common (Batterman 2002). A model of a mechanism is a model that describes a mechanism. It need not be a mechanical model or a mechanism schema, in the above sense, to play that role.

Some mechanists reserve the term “mechanical models” for models that describe the entities, activities, and organizational features of a system. According to Glennan’s (2005) account, a mechanical model that leaves out some relevant features is, ipso facto , incomplete and sketchy. One specific instantiation of this debate concerns the explanatory force of functional models in psychology. Piccinini and Craver (2011) argue that such models should be understood as mechanistic sketches, black-box models to be evaluated and filled in as details about the underlying mechanism are discovered. Black box models are incomplete in virtue of leaving out details about underling mechanisms and that those models ultimately depend for their explanatory force on the promise that the functional models do, in fact, correspond to how the mechanism works. (See Weiskopf 2011 for a criticism of this account and Povich forthcoming for a response)

One might talk about mechanistic explanation in a way that abstracts from the kind of model used to describe the mechanism: the commitment to mechanistic explanation is not a commitment about the form of the model but rather a commitment about what such models must represent: namely, causal and mechanistic structures. Models are explanatory in virtue of the fact that they represent the causal/mechanistic structures that produce, underlie, or maintain the phenomenon. They are non-mechanistic if they refer to some non-causal, non-mechanistic kind of relation (Salmon 1984; Craver 2014).

To date, much of the work on mechanistic explanation has been driven by the goal of providing a descriptively and normatively adequate theory of mechanistic explanation. Some claim there are kinds of explanation that rely very little on a precise understanding of the mechanistic details of a system (Woodward 2014; Weiskopf 2011) or that work fundamentally by removing all such details from one’s model (Batterman and Rice 2014). Resolving such debates will require being very clear about precisely what one expects out of a philosophical theory of scientific explanation and what one takes a scientific explanation to be (Strevens 2008; Craver 2014). Research is required to understand the diverse representational forms that scientists use to represent mechanisms (Burnston et al. forthcoming), and to understand the role of idealization in mechanistic explanation (Levy and Bechtel 2012; Huneman 2010). Further work is also required to limn the boundaries between mechanistic explanation and other putative varieties of explanation and to say, as perspicuously as Hempel or the causal-mechanical theory, what a model must do to count as explanatory and precisely how good explanations are to be distinguished from bad.

4. Metaphysics of Mechanisms

In this section, we review some of the ways that the concept of mechanism has been used in diverse areas of metaphysics. Of all the areas we have discussed, this is likely the most in need of future development. Here we discuss the relationship between mechanisms and laws, emergence, realization, natural kinds, and functions.

In much of the early literature on mechanisms, mechanisms are contrasted explicitly with laws of nature (Bechtel 1988; Bechtel and Abrahamsen 2005; MDC 2000). This contrast clearly grew out of an emerging consensus in philosophy that there are few, or perhaps no, laws of biology (see Section 3.1 above). The empirical generalizations one finds in biology tend to be hedged by ceteris paribus clauses; whether they hold or not depends on background conditions that might not hold and on conditions internal to the mechanism that might fail. These generalizations, in short, are mechanistically explicable; what necessity they have derives from a mechanism (Cummins 2000; Glennan 1996). Mechanisms thus seem to play the role of laws in the biological sciences: we seek mechanisms to explain, predict, and control phenomena in nature even if mechanisms lack many of the characteristics definitive of laws in the logical empiricist framework (such as universality, inviolable necessity, or unrestricted scope).

One specific strand of this discussion emerged from consideration of Weber’s (2005) argument that biology is heteronymous , i.e., that it ultimately borrows its explanatory power from the laws of physics and chemistry. Weber uses Hodgkin and Huxley’s model of the action potential as an exemplar of the reducing biological phenomena to physical laws (such as Ohm’s law and the Nernst equation). Craver (2006) responds that the explanatory force of Hodgkin and Huxley’s model, in fact, requires a grasp of the distinctly biological properties of ion channels, which properties were black-boxed in Hodgkin and Huxley’s total current equation (see also Craver 2007; Bogen 2008b; Weber 2008).

Yet the contrast between laws and mechanisms has not always been entirely clear. Some, such as Bogen (2005), Machamer (2004), and Glennan (forthcoming) emphasize that causes and mechanisms are, at base, singular, not general or universal. Leuridan (2010), building on the work of Mitchell (2000), objects that mechanisms cannot replace laws of nature in our conceptual understanding of explanation and the metaphysics of science. Scientists rarely investigate token mechanisms, one might think, but are much more interested in types. And once one starts talking about types of mechanisms, one is back in the business of formulating general regularities about how mechanisms work. So it would appear that the concept of mechanism cannot supplant the work that generalization was supposed to do, but requires the idea of regularity, and so something akin to laws, if it is to do that explanatory work (see Andersen 2011, 2012, 2014a,b; Krickel 2014). For a reply to Leuridan, see Kaiser and Craver (2013).

Work on mechanisms has also helped to clarify the idea of levels of organization and its relation to other forms of organization and non-mechanistic forms of emergence.

Many mechanists, following Simon (1996 [1962]), emphasize that biological systems are hierarchically organized into near-decomposable structures: mechanisms within mechanisms, within mechanism. Using the parable of Tempus and Hora, Simon (1962) argued that a watchmaker who builds hierarchically decomposable watches (Tempus) will make more watches than one who builds holistic watches (Hora). This parable led Simon to the conclusion that evolved structures are more likely to be nearly decomposable into hierarchically organized, more or less stable structures and sub-structures. Some have objected that the story is misleading because evolution does not construct organisms from scratch, piece by piece (Bechtel 2009b). Steel (2008), building on the work of others (Schlosser and Wagner 2004), therefore attempts to reconstruct this argument as a way of showing that evolved systems are more likely to be modular: systems made of independently manipulable parts can quarantine the effects of changes to specific parts, giving them added flexibility to make local changes without causing catastrophic side-effects.

The near decomposability of mechanisms is directly related to the idea that mechanisms span multiple levels of organization. The behavior of the whole is explained in terms of the activities and interactions among the component parts. These activities and interactions are themselves sustained by underlying activities and interactions among component parts, and so on (see Bechtel and Richardson 2010 [1993]). Craver (2007) defines levels of mechanisms in terms of a relationship between the behavior (ψ) exhibited by a system ( S ) and the activity (φ) of some component part ( X ) of that system. On this account, X ’s φ-ing is at a lower level of mechanistic organization than S ’s ψ-ing if and only if (i) X is a part of S , and (ii) X ’s φ-ing is a component in S ’s ψ-ing. In short, to say that something is at a lower mechanistic level than the mechanism as a whole is to say that it is a working part of the mechanism. Though the term “level” is used in many legitimate ways, levels of mechanisms seem to play a central role in structuring the relations among many different models in contemporary biology (e.g., between Mendelian and molecular genetics (Darden 2006), between learning and memory and channel physiology (Craver 2007), and between population-level variation and developmental mechanisms (Tabery 2009, 2014a)).

One implication of this view of levels, combined with certain familiar assumptions about causal relations, is that there can be no causal relationships between items at different levels of mechanisms. There can be causal relationships between things of different sizes, and there can be causal relationships between things described in very different vocabularies; but (again, conjoined with certain assumptions about the temporal asymmetry of cause and effect and the independence of cause and effect) there cannot be causal relationships between the behavior of a mechanism and the activities of the parts that jointly constitute that behavior. Claims about interlevel causation, which are ubiquitous in the scientific literature, are best understood either as targeting a different sense of levels or, concerning levels of mechanisms, as expressing hybrid claims combining constitutive claims about the relationship between the behavior of the mechanism as a whole and the activities of its parts, and causal claims concerning relationships between things not related as part and whole (Craver and Bechtel 2007). For alternative interpretations of levels, see Fehr 2004; Leuridan 2011; Thalos 2013; Eronen 2013; 2015; Baumgartner and Gebhardter forthcoming; Romero forthcoming. For reflections on the metaphysical status of higher-level phenomena and higher-level causes, see Baumgartner 2010; Glennan 2010 a, b; Hoffman-Kolss 2014, as well as the entries on causation and manipulability , physicalism , and scientific reduction .

As noted above, the fact that phenomena at higher levels of mechanisms depend upon the organization of component parts entails that the properties/activities of wholes are not simple sums of the properties/activities of the parts. Levels of mechanisms can thus be contrasted with levels of mere aggregation. Because the whole is greater (in this sense) than the sum of the parts, some (such as Wimsatt) have found it appropriate to describe this as a kind of emergence. Mechanistic (or organizational) emergence thus understood is ubiquitous and banal but extremely important for understanding how scientists explain things.

Also familiar is epistemic emergence , the inability to predict the properties or behaviors of wholes from properties and behaviors of the parts. Epistemic emergence can arise as a result of ignorance, such as failing to recognize a relevant variable, or from failing to know how different variables interact in complex networks. It might also result from limitations on human cognitive abilities or in current-generation representational tools (Bedau 1997; Boogerd 2005; Richardson and Stephan 2007). The practical necessity of studying mechanisms by decomposing them into component parts raises the epistemic challenge of putting the parts back together again in a way that actually works (Bechtel 2013a).

The mechanists’ emphasis on mechanistic/organizational and epistemic emergence contrasts with their desire to distance themselves from spooky emergence (Richardson and Stephan 2007). Spooky emergence would involve the appearance of new properties with no sufficient basis in mechanisms. It is not clear that emergent properties are properly said to be properties of the necessary mechanisms; and it is not clear in what since the emergent property is “emergent” rather than a fundamental feature of the causal structure of the world.

In short, such forms of emergence are altogether distinct from, and so gain no plausibility from, verbal association with organizational/mechanistic and epistemic emergence.

Because the framework concept of a mechanism is so useful for thinking about levels and explanation in the sciences, some scholars have sought in the notion of mechanism a way of fleshing out the ontological relationship of realization.

According to the “flat view” (Gillett 2002) realization is a relationship between different properties of one and the same thing (Kim 1998; Shapiro 2000; Shoemaker 2003, 2007; Polger 2007). The subset view, which holds that a property P1 (e.g., mean kinetic energy of the gas) realizes property P2 (e.g., temperature of the gas) when the causal powers distinctive of P2 (temperature) are a subset of the causal powers distinctive of P2 (mean kinetic energy), is an example of the flat view. P1 and P2 are both attributed to the same thing, the gas (Gillett 2002, 2003). The dimensioned view describes realization as a relationship holding between the properties of wholes and the properties of the parts and their organization. This view of realization comports with the explanatory aims of the special sciences and fits nicely with the evidential base on which interlevel claims are grounded (see Aizawa and Gillett 2011). Gillett has since expanded this notion to handle the realization of objects, properties, and processes (Gillett 2013); for criticism and alternatives, see Polger 2010; Melnyk 2003; Melnyk 2010)

Mechanistic theories of natural kinds develop out of Boyd’s homeostatic property cluster (HPC) view. The HPC view is a theory of natural kinds designed to work in domains with high individual variability. The HPC view is offered as a third way between essentialism and nominalism about kinds in the special sciences (Boyd 1991, 1997, 1999; Kornblith 1993; Wilson 1999, 2005; see also the entry on natural kinds ).

According to this view, a natural kind is characterized by i) a cluster of properties that regularly co-occur, and ii) a similarity generating mechanism that explains why the properties in (i) tend to co-occur. In short, kinds are property clusters explained by mechanisms.

This view of natural kinds has been deployed to argue for taxonomic revision in, for example, the biology of human emotion (Griffiths 1997), the structure of concepts (Machery 2009), and the taxonomy of psychiatry (Kendler, Zachar, and Craver 2010; see Craver 2009): a putatively single kind is split into multiple kinds because it is discovered that distinct properties in a property cluster are explained by distinct mechanisms. This view of kinds can also be used to make sense of kinds that are historically transient and, in some ways, the product of human attitudes and so socially constructed in this straightforward sense; perhaps race is like this (see Kuorikoski and Pöyhönin 2012; Khalidi 2013).

Emphasis on the importance of mechanisms is historically associated with the rejection of teleology and formal causes (e.g., Westfall 1971). Yet contemporary biology and many other special sciences, despite the widespread acceptance of the mechanism framework concept, continue to make use of the concept of function, a teleological notion (see the entry on teleological notions in biology ). How is the notion of function at play in contemporary science related to the concept of mechanism? Craver (2001a), following Cummins (1975), argues that functional description is a perspectival means of situating some part within a higher-level mechanism. According to this view, teleology is not a feature of the world so much as it is imposed upon it by an intentional describer (see also Machamer 1977). Garson (2011, 2012), following Wimsatt (1972b), Wright (1973) and Neander (1991a,b), argues that functions are effects of an item that are part of the etiological explanation (through selection, learning, or reinforcement) for why the item is present; as such, functions are reduced to causal histories. In a third view, Maley and Piccinini (forthcoming) argue that the (teleological) function of an item is its contribution to the goals of organisms, which may be objective or subjective goals. Like Garson, Maley and Piccinini hold that functions are objective (i.e., not observer relative). Unlike Garson, however, they are not grounded in the etiology of the item but in their current contribution to survival or reproduction (the objective goals of organisms) or to what the organism itself desires (the subjective goals of organisms).

What must the world be like for this mechanistic perspective to be accurate? Clearly, there are many ways of answering this question from different metaphysical starting assumptions. And clearly, many metaphysical starting assumptions rule this world picture out as illegitimate. The clearest path forward, it would seem, is to work out precisely what one must be committed to in holding that the world is composed of a hierarchy of mechanisms and precisely what of that can be recovered on the basis of different starting assumptions.

That said, not all applications of the mechanism framework require a fully articulated metaphysics. Work on discovery and explanation might proceed perfectly well without embracing any particular metaphysical world picture. Philosophers with different interests (discovery, explanation, testing, reduction, emergence, and so) are likely to elaborate the concept in different ways. There is every reason to doubt, that the idea of mechanism can be given a one-size fits all metaphysical analysis that will adequately address the diverse philosophical ends to which the concept is being deployed.

5. Relations between Scientific Disciplines: From Theory Reduction to Mechanism Integration

According to Nagel (1961), reduction is a species of covering-law explanation: one theory is reduced to another when it is possible to identify the theoretical terms of the first with those of the second and to literally derive the first from the second. On the assumption that scientific disciplines and theories correspond to one another, reduction serves as a model of interdisciplinary integration as well. Mechanist’s objections to the covering-law model of constitutive explanation (i.e., micro-reduction) are discussed above (see Section 3.1 ); here the focus is rather on how distinct disciplines of science are integrated.

On the Nagel view, reduction is an interlevel relationship. It is also a relationship between theories. Theories about phenomena at a higher level (e.g., gases, lightning, and life) are reduced to (i.e., derived from) theories about phenomena at lower levels (e.g., molecules, electrons, and physiological systems). Finally, the relationship is formally specified and has little to do with either the content of the theories or the material structures those theories describe. From the mechanistic perspective, each of these features of the Nagel model is problematic.

5.2 Mechanism Integration

First, mechanists criticize the idea that reduction should be understood primarily as a relationship between theories . In integrating their results, scientists are not simply building theories simpliciter; they are building theories about mechanisms. Mechanisms can perhaps be described using formal accounts of theories—perhaps they can be axiomatized in predicate logic or reconstructed as set theoretic predicates. But such formal accounts of the structures of scientific theories gloss over the mechanistic structures crucial for understanding how these theories are constructed and evaluated (Craver 2001b).

Mechanists also challenge the idea that disciplines are related by way of the relationship between their theories. Darden and Maull (1977) argued that disciplinary fields often integrate their findings through the construction of interfield theories, appealing to diverse material relationships between items in the different fields’ domains: relations such as cause and effect, part and whole, or structure and function. Darden and Maull did not offer a general account of interfield theories, but Bechtel presciently suggested that such theories often take the form of descriptions of mechanisms (Bechtel 1988: 101–102).

The mechanistic approach also has been claimed to have many advantages over reduction for thinking about interlevel forms of interdisciplinary integration. First, it provides a straightforward way to interpret the talk of levels (see Sections 2.4.5 and 4.2 ). Second, it offers significantly more insight into what interlevel integration is, into the evidential constraints by which interlevel bridges are evaluated, and into the forces driving the co-evolution of work at different levels. Constraints on the parts, their causal interactions, and their spatial, temporal, and hierarchical organization all help to flesh out an interlevel integration. Finally, mechanists repeatedly recognize the need to not only look down to the constitutive mechanisms responsible for a given phenomenon (emphasized by classical reduction models), but also to look up and around to the context within which the phenomenon is embedded: interlevel integration is an effort to see how phenomena at many different levels are related to one another (Bechtel 2009a; Craver 2007).

Mechanists have developed several extended examples of the many forms of mechanism integration pursued in mechanistic research programs. Darden (2005), for example, suggests that philosophers in the grip of classical reduction fundamentally misunderstood the relationship between Mendelian and molecular genetics. While reductionists see it as an instance of interlevel explanation, she argues, it is in fact a case in which different scientists worked on different parts of a mechanism that are etiologically (not constitutively) related to one another. Mendelian genetics did not reduce to molecular biology; rather, classical geneticists and molecular biologists integrated their work by focusing on different working entities in the sequentially operating chromosomal and molecular hereditary mechanisms. Examples have also been drawn from the discovery of the mechanisms of protein synthesis (Darden 2006) and cell biology (Bechtel 2006). Craver (2007) uses examples from the neuroscience of memory to explore how multilevel integration does and ought to proceed. In each case, the search for mechanisms serves as an abstract scaffold onto and around which the findings of diverse scientists converge.

The mechanistic perspective tends to emphasize integrative pluralism in scientific research (Mitchell 2003, 2009). The goal is not to explain the less fundamental in terms of the more fundamental in a step-wise relating of monolithic theories at one level to monolithic theories at another. Rather, such scientific achievements are collaborative and piecemeal, adding incremental constraints to an emerging picture of how a mechanism works both at a level and across levels. The many scientific disciplines that investigate a phenomenon pluralistically co-exist and co-inform one another by integratively contributing to the etiological, constitutive, and contextual mechanistic explanations of that phenomenon (Bechtel 2009a; Tabery 2014a).

The Nagel model of theory reduction offers a clear vision of the “unity of science” (see the entry on unity of science ). According to the model, the unity among scientific disciplines is achieved by reducing theories of higher-level disciplines to the theories of lower-level disciplines. Integration, on that vision, is understood as progress toward a grand, unified body of scientific knowledge. For mechanists, in contrast, integration is piecemeal, local, and pluralistic. What sort of unity could such an “integration” sustain? This question plays out in a back-and-forth between Longino and Tabery concerning disciplinary relationships in the behavioral sciences. Tabery argues that disciplines as disparate as neurobiology and quantitative genetics could pluralistically co-exist and co-inform one another’s causal explanations of complex behaviors by way of mechanism integration. Longino counters that Tabery’s pluralism is only a “moderate” sort because the push for integration ultimately is a push for unification (Longino 2013, 2014; Tabery 2014a,b). Sullivan (2009) also challenges the push for mechanism integration; she argues that there are significant barriers to the kind of integration mechanists envision. Different laboratories use different experimental protocols to study what they assume to be the same phenomenon; however, these different protocols often in fact target different phenomena, so the integration achieved by combining results is only illusory. These discussions are symptomatic of more general philosophical questions faced by mechanists: How are mechanism integrations actually achieved (as opposed to just asserted)? And what is the relationship between mechanism integration and unification? The new mechanical philosophy stands to benefit from future efforts to situate mechanistic integration into more general philosophical views of integration and pluralism.

6. Discovery: From A-ha Moments to Discovery Strategies

What can philosophers say about scientific discovery? Many logical empiricists had a simple answer: Nothing. According to Popper, for example, philosophers can illuminate the epistemology of testing, but they can say nothing of substance about how scientists generate the ideas to be tested (Popper 1959). Such “A-ha!” moments of creativity are in the province of psychology, not philosophy. Reichenbach distinguished the context of discovery from the context of justification (the “context distinction”) (Reichenbach 1938; but see the entry on Hans Reichenbach for an alternate interpretation of this distinction). The process of scientific discovery was thus largely off limits to philosophers.

Not all philosophers of science agreed. Hanson, for example, articulated a logic of discovery involving abductive inferences from anomalous data to new hypotheses designed to account for them (Hanson 1958). Others focused on methodologies of discovery that could either allow one to rationally reconstruct why something was discoverable at a given time (Nickels 1985) or to explain why a new hypothesis is considered promising and worthy of further investigation (Schaffner 1993). Early contributions to the new mechanical philosophy followed this path and characterized investigative strategies scientists use to discover mechanisms (see the entry on scientific discovery ).

Bechtel and Richardson’s Discovering Complexity (2010 [1993]) is organized around a flowchart representing choice-points in the discovery of a mechanism. The process of searching for mechanisms begins with a provisional characterization of the phenomenon. Then follow strategies of localizing the mechanism within the system, and decomposing the phenomenon into distinct sub-functions. Localization of function involves determining which of these sub-functions of the system is performed by which parts. Bechtel and Richardson further characterize the use of excitatory and inhibitory experiments to obtain these kinds of information. Bechtel and Abrahamsen (2013) add a subsequent stage, in which scientists recompose what they have learned about the functional parts by putting them back together to produce the phenomenon in question (perhaps using simulations).

Darden also emphasized mechanisms as an important framework concept in scientific discovery (Darden 1980, 1982, 1986, 1991). In the discovery of protein synthesis (jointly investigated by molecular biologists and biochemists in the 1950s and 1960s), scientists didn’t simply have an “A-ha” moment. Rather, they deployed strategies for revealing how a mechanism works (Darden 2006; Craver and Darden 2013). Darden characterizes the process of mechanism discovery as an “extended, piecemeal process with hypotheses undergoing iterative refinement”; that process occurs via the construction, evaluation, and revision of mechanism schemas in light of observational and experimental constraints (Darden 2006: 272).

Darden’s construction strategies are strategies for generating new hypotheses about a mechanism. In addition to decomposition and localization, Darden shows that scientists often borrow a schema type from another area of science, as when selection-type mechanisms were borrowed to understand how the immune system works, or assemble a mechanism from known modules of functional activity ( modular sub-assembly ), as is common in biochemistry and molecular biology. Sometimes, scientists know one part of the mechanism and attempt to work forward or backward through to the other parts and activities. In the discovery of the mechanism of protein synthesis, for example, molecular biologists worked forward from the structure of DNA to figure out what molecules could interact with it ( forward chaining ), and biochemists worked backward from proteins to figure out what chemical reactions would be necessary to create them ( backward chaining ). They met in the middle at RNA. Protein synthesis is now understood to involve transcribing DNA into RNA and then translating RNA into proteins. Far from being philosophically inscrutable, Darden points out that scientists used what they knew about the working entities and activities in the mechanism to infer what could come next or before in the mechanism of protein synthesis (Darden 2006; see also the entry on molecular biology ).

Evaluation strategies, for Darden, involve constraint-based reasoning to limn the contours of the space of possible mechanisms for a given phenomenon. Often scientists reason about how a mechanism works by building off basic findings concerning the spatial and temporal organization of its parts. Harvey, for example, reasoned his way to the circulation of the blood by considering the locations of the valves of the veins and their orientation with respect to the heart. These organizational constraints, and many others, combined to narrow the space of possible mechanisms to a small region containing a model in which the blood completes a circuit of the body (Craver and Darden 2013).

Darden and Craver also discuss experimental strategies for learning how a mechanism works. These strategies reveal how different entities and activities in a mechanism act, interact, and are organized together. For example, one might intervene to remove a putative component to see if and how the mechanism functions in its absence ( inhibitory experiments). Or one might stimulate that component to see if it can drive the mechanism or modulate its behavior. Or one might activate a mechanism by placing it in the precipitating conditions for the phenomenon and observe how the entity or activity changes as the mechanism works. Craver (2002) discusses these under the heading of “interlevel experiments” (see also Harinen forthcoming). Craver and Darden (2013) also discuss more complex kinds of experiments for learning what sort of entity or activity contributes to a process and for learning more complex features of a mechanism’s organization.

Datteri (2009; Datteri and Tamburrini 2007), explores the use of robotic simulations for the purposes of testing mechanisms. They discuss both how assumptions are built into robotic models and how experiments can be designed to reveal how mechanisms work. This work extends the mechanistic framework into the area of bio-robotics and reveals a set of strategies distinct from those explored in Darden’s work.

Rather than focusing on the process by which mechanism schemas are constructed, evaluated, and revised, Steele focuses on the question of how one extrapolates from a sample population or a model organism to the structure of a mechanism in the target. Will a treatment proven to suppress tumors in mice (a model organism) also suppress tumors in humans (the target population)? After developing a probabilistic account of mechanisms, Steele considers how researchers get around what he calls the extrapolator’s circle: determining

how we could know that the model and the target are similar in causally relevant respects without already knowing the causal relationship in the target. (Steel 2008: 78)

Steel breaks the extrapolator’s circle by developing a mechanisms-based extrapolation strategy—the strategy of comparative process tracing . Once a mechanism for some phenomenon has been elucidated in a model (such as a particular process of carcinogenesis in rats), scientists (toxicologists in this case) then compare key stages (particularly downstream stages) of the model with the stages in the target, paying particular attention to points in the process where differences are most likely to arise. The greater the similarities of the entities, activities, and organization of the mechanisms in both populations, the stronger is the basis for extrapolation; the greater the differences, the weaker the basis (but see Howick et al. 2013; see also the sections on extrapolation in the entries on molecular biology and experiment in biology ).

Discovery in medicine is another domain where the mechanical philosophy has been applied. Thagard draws on the case of H. pylori as a cause of ulcers to provide an account of how investigating mechanisms contributes to scientific discovery.

Thagard draws attention to both statistical evidence that suggests ulcers are somehow associated with H. pylori , as well as mechanistic evidence that can explain how the agent of infection could persist in a hostile environment long enough to cause an ulcer. More recently, philosophers interested in evidence-based medicine have probed the relationship between these two types of evidence in the health sciences. Russo and Williamson argue that both types of evidence are necessary to justify causal inference; the correlational evidence establishes that there is a difference-making relation between some cause and some effect, while the mechanistic evidence establishes how exactly the cause produces its effect—the “Russo-Williamson Thesis” (Russo and Williamson 2007). Philosophers have since refined the Russo-Williamson Thesis, pointing out, for instance, that “type of evidence” could refer to different methodologies for gathering evidence or to different objects of evidence. Difference-making methodologies include observational studies and randomized controlled trials, while mechanistic methodologies include interventionist experiments such as those described above; likewise, the object of evidence could be the evidence of an associated difference or it could be the evidence concerning the mechanism linking the cause and effect (Illari 2011; see also Campaner 2011). Evidence-based medicine hierarchies, which rank different kinds of evidence in terms of its epistemic strength, tend to prioritize evidence from difference-making methodologies (such as randomized controlled trials and meta-analyses) over mechanistic evidence; in reply, these philosophers argue that the different types of evidence are on a par (each with its own strengths and weaknesses) and advocate for integrating difference-making and mechanistic evidence, a sentiment which aligns with the emphasis on mechanism integration discussed in Section 5.2 above (Clarke et al. 2013, 2014).

Many mechanists have explored the strategies that scientists use in discovery. Bechtel and Richardson attended to decomposition and localization; Darden and Craver highlighted forward and backward chaining; Russo and Williamson emphasized drawing on both difference-making and mechanistic evidence. These strategies were found in specific, experimental sciences, such as neuroscience and molecular biology. So one task for philosophers moving forward is to assess whether or not similar strategies exist in other sciences, especially those that operate outside the traditional laboratory, both in the human sciences (such as sociology and economics) and in the physical sciences (such as cosmology).

We also expect tremendous development to come from bridging the gap between the qualitative accounts of mechanisms and mechanistic explanation developed in the new mechanism and quantitative theories of discovery from the discipline of machine learning and causal modeling (Spirtes et al. 2000; Pearl 2009). The latter offer tools to mine correlational data for causal dependencies. Such tools might escape more qualitative, historical approaches and might, in fact, go beyond the common strategies that scientists traditionally use. Such tools also offer a means to assess discovery strategies by exploring the conditions under which they succeed and fail and the efficiency with which they deliver verdicts on causal hypotheses.

The new mechanical philosophy and, more generally, attention to the framework concept of “mechanism” has expanded rapidly over the last two decades bringing with it new orientations toward a wide range of issues in the philosophy of science. Yet it is clear that many of the major topics are only beginning to develop, leaving a lot of work for scholars to elaborate the basic commitments of this framework and to consider what it means to do science outside of that framework. The near future is likely to see continued discussion of the implications and limits of this framework for thinking about science and scientific practice.

  • Aizawa, K. and C. Gillett, 2011, “The Autonomy of Psychology in the Age of Neuroscience”, in Causality in the Sciences , in Illari et al. 2011: 202–23.
  • Allen, G.E., 2005, “Mechanism, Vitalism and Organicism in Late Nineteenth and Twentieth Century Biology: The Importance of Historical Context”, in Craver and Darden 2005: 261–283.
  • Alon, U., 2006, An Introduction to Systems Biology , Boca Raton: Chapman and Hall/CRC Press.
  • Andersen, Holly, 2011, “Mechanisms, Laws, and Regularities”, Philosophy of Science , 78(2): 325–331.
  • –––, 2012, “The Case for Regularity in Mechanistic Causal Explanation”, Synthese , 189: 415–432.
  • –––, 2014a, “A Field Guide to Mechanisms: Part I”, Philosophy Compass , 4: 274–283.
  • –––, 2014b, “A Field Guide to Mechanisms: Part II”, Philosophy Compass , 4: 283–297.
  • Baker [now Byron], J.M., 2005, “Adaptive Speciation: The Role of Natural Selection in Mechanisms of Geographic and Non-geographic Speciation”, in Craver and Darden 2005: 303–326.
  • Barros, D.B., 2008, “Natural Selection as a Mechanism”, Philosophy of Science , 75: 306–322.
  • Batterman, R., 2002, “Asymptotics and the Role of Minimal Models”, The British Journal for the Philosophy of Science , 53: 21–38.
  • Batterman, R. and C. Rice, 2014, “Minimal Model Explanations”, Philosophy of Science , 81: 349–376.
  • Baumgartner, M., 2010, “Interventionism and Epiphenomenalism”, Canadian Journal of Philosophy , 40: 359–384.
  • –––, 2013, “Rendering Interventionism and Non-Reductive Physicalism Compatible”, Dialectica , 67: 1–27.
  • Baumgartner, M. and A. Gebharter, forthcoming, “Constitutive Relevance, Mutual Manipulability, and Fat-Handedness”, The British Journal for the Philosophy of Science . doi:10.1093/bjps/axv003 [ Baumgartner and Gebharter forthcoming available online ]
  • Beatty, J., 1995, “The Evolutionary Contingency Thesis”, in James G. Lennox and Gereon Wolters (eds), Concepts, Theories, and Rationality in the Biological Sciences , Pittsburgh, PA: University of Pittsburgh Press, pp. 45–81.
  • Bechtel, W., 1988, Philosophy of science: An overview for cognitive science , Hillsdale, NJ: Erlbaum. Italian translation Filosofia della scienza e scienza cognitiva, Gius. Laterza & Figli, 1995. Second edition in preparation. [ some of Bechtel 1988 available online ]
  • –––, 2006, Discovering Cell Mechanisms: The Creation of Modern Cell Biology , Cambridge: Cambridge University Press. [ some of Bechtel 2006 available online ]
  • –––, 2008, Mental Mechanisms: Philosophical Perspectives on Cognitive Neuroscience , London: Routledge.
  • –––, 2009a, “Looking Down, Around, and Up: Mechanistic Explanation in Psychology”, Philosophical Psychology , 22: 543–564. [ Bechtel 2009a available online ]
  • –––, 2009b, “Explanation: Mechanism, Modularity, and Situated Cognition”, in P. Robbins and M. Aydede (eds). Cambridge handbook of situated cognition , Cambridge: Cambridge University Press, pp. 155–170. [ Bechtel 2009b available online ]
  • –––, 2011, “Mechanism and Biological Explanation”, Philosophy of Science , 78: 533–557.
  • –––, 2013a, “Addressing the Vitalist’s Challenge to Mechanistic Science: Dynamic Mechanistic Explanation”, in S. Normandin & C.T. Wolfe (eds), Vitalism and the Scientific Image in Post-Enlightenment Life Science, 1800–2010 , Dordrecht: Springer, pp. 345–370.
  • –––, 2013b, “From Molecules to Behavior and the Clinic: Integration in Chronobiology”, Studies in History and Philosophy of Biological and Biomedical Sciences , 44: 493–502.
  • Bechtel, W. and A. Abrahamsen, 2005, “Explanation: A Mechanistic Alternative”, Studies in History and Philosophy of the Biological and Biomedical Sciences , 36: 421–441. [ Bechtel and Abrahamsen 2005 available online ]
  • –––, 2013, “Thinking Dynamically about Biological Mechanisms: Networks of Coupled Oscillators”, Foundations of Science , 18: 707–723. [ Bechtel and Abrahamsen 2013 available online ]
  • Bechtel, W. and R.C. Richardson, 2010 [1993], Discovering Complexity: Decomposition and Localization as Strategies in Scientific Research , Second Edition. Cambridge, MA: MIT Press/Bradford Books.
  • Bedau, M., 1997, “Weak Emergence”, Philosophical Perspectives , 11: 375–399.
  • Beebee, H., C. Hitchcock, and P. Menzies, (eds), 2010, The Oxford Handbook of Causation , Oxford: Oxford University Press.
  • Boas, M., 1952, “The Establishment of the Mechanical Philosophy”, Osiris , 10: 412–541.
  • Bogen, J., 2005, “Regularities and Causality; Generalizations and Causal Explanations”, in Craver and Darden 2005: 397–420.
  • –––, 2008a, “Causally Productive Activities”, Studies in History and Philosophy of Science , 39: 112–123.
  • –––, 2008b, “The Hodgkin-Huxley Equations and the Concrete Model: Comments on Craver, Schaffner, and Weber”, Philosophy of Science , 75: 1034–1046.
  • Boogerd, F.C., F.J. Bruggeman, R.C. Richardson A. Stephan, and H. V. Westerhoff, 2005, “Emergence and Its Place in Nature: A Case Study of Biochemical Networks”, Synthese , 145: 131–164.
  • Boyd, R., 1991, “Realism, Anti-Foundationalism and the Enthusiasm for Natural Kinds”, Philosophical Studies , 61: 127–148.
  • –––, 1997, “Kinds as the ‘Workmanship of Men’: Realism, Constructivism, and Natural Kinds”, in J. Nida-Rumelin (ed.), Rationality, Realism, Revision: Proceedings of the 3rd International Congress of the Society for Analytical Philosophy , New York: Walter de Gruyter, pp. 52–89.
  • –––, 1999, “Homeostasis, Species, and Higher Taxa”, in R.A. Wilson (ed.), Species , MIT Press: Cambridge, 141-185.
  • Brandon, R., 1985, “Grene on Mechanism and Reductionism: More Than Just a Side Issue”, in Peter Asquith and Philip Kitcher (eds), PSA 1984 , v. 2. East Lansing, MI: Philosophy of Science Association, pp. 345–353.
  • Bromberger, S., 1966, “Why Questions”, in R.G. Colodny (ed.), Mind and Cosmos , Pittsburgh: University of Pittsburgh Press, pp. 86–111.
  • Burnston, D.C., B. Sheredos, A. Abrahamsen, and W. Bechtel, in press, “Scientists’ Use of Diagrams in Developing Mechanistic Explanations: A Case Study from Chronobiology”, Pragmatics and Cognition .
  • Campaner, R., 2011, “Understanding Mechanisms in the Health Sciences”, Theoretical Medicine and Bioethics , 32: 5–17.
  • Cartwright, N.D., 1989, Nature’s Capacities and their Measurement , New York: Oxford University Press. [ Cartwright 1989 available online ]
  • –––, 1999, The Dappled World: A Study of the Boundaries of Science , Cambridge: Cambridge University Press.
  • –––, 2001, “Modularity: It Can—and Generally Does—Fail”, in Stochastic Dependence and Causality , D. Costantini, M.C. Galavotti, and P. Suppes (eds), Stanford: CSLI Publications, pp 65–84.
  • –––, 2002, “Against Modularity, the Causal Markov Condition and Any Link Between the Two: Comments on Hausman and Woodward”, British Journal for the Philosophy of Science , 53: 411–453.
  • Chemero A. and M. Silberstein, 2008, “After the Philosophy of Mind: Replacing Scholasticism with Science”, Philosophy of Science , 75: 1–27.
  • Churchland, P.S., 1986, Neurophilosophy: Toward a Unified Science of the Mind/Brain , Cambridge, MA: MIT Press.
  • Churchland, P.S. and T.J. Sejnowski, 1992, The Computational Brain , Cambridge, MA: MIT Press.
  • Chirimuuta, M., 2014, “Minimal Models and Canonical Neural Computations: The Distinctness of Computational Explanation in Neuroscience”. Synthese , 191: 127–153.
  • Clarke, B., D. Gillies, P. Illari, F. Russo, and J. Williamson, 2013, “The Evidence that Evidence-Based Medicine Omits”, Preventive Medicine , 57: 745–747.
  • –––, 2014, “Mechanisms and the Evidence Hierarchy”, Topoi , 33: 339–360.
  • Couch, M.B., 2011, “Mechanisms and Constitutive Relevance”, Synthese , 183: 375–88.
  • Craver, C.F., 2001a, “Role Functions, Mechanisms and Hierarchy”, Philosophy of Science , 68: 31–55.
  • –––, 2001b, “Structures of Scientific Theories”, in P.K. Machamer and M. Silberstein (eds), Blackwell Guide to the Philosophy of Science , Blackwell: Oxford, pp. 55–79.
  • –––, 2006, “When Mechanistic Models Explain”, Synthese , 153: 355–376.
  • –––, 2007, Explaining the Brain: Mechanisms and the Mosaic Unity of Neuroscience , Oxford: Clarendon Press.
  • –––, 2009, “Mechanisms and Natural Kinds”, Philosophical Psychology , 22: 575–594.
  • –––, 2013, “Functions and Mechanisms: A Perspectivalist Account”, in P. Huneman (ed.), Functions: Selection and Mechanisms , Dordrecht: Springer, pp. 133–158.
  • –––, 2014, “The Ontic Conception of Scientific Explanation”, in Andreas Hütteman and Marie Kaiser (eds), Explanation in the Special Sciences: The Cases of Biology and History , Dordrecht: Springer, pp. 27–52.
  • Craver, C.F. and A. Alexandrova, 2008, “No Revolution Necessary: Neural Mechanisms for Economics”, Economics and Philosophy , 24: 381–406.
  • Craver, C.F. and W.M. Bechtel, 2007, “Top-down Causation without Top-down Causes”, Biology and Philosophy , 22: 547–563.
  • Craver, C.F. and L. Darden, 2013, In Search of Mechanisms: Discoveries Across the Life Sciences , Chicago: University of Chicago Press.
  • Cummins, R., 1975, “Functional Analysis”, Journal of Philosophy , 72: 741–764.
  • –––, 1983, The Nature of Psychological Explanation , Cambridge, MA: Bradford/MIT Press.
  • –––, 2000, “‘How Does It Work?’ Vs. ‘What Are The Laws?’ Two Conceptions of Psychological Explanation”, in F. Keil and R. Wilson (eds), Explanation and Cognition , Cambridge, MA: MIT Press, pp. 117–144.
  • Darden, L., 1986, “Reasoning in Theory Construction: Analogies, Interfield Connections, and Levels of Organization”, in P. Weingartner and G. Dorn (eds), Foundations of Biology , Vienna: Holder-Pichler-Tempsky, pp. 99–107.
  • –––, 1991, Theory Change in Science: Strategies from Mendelian Genetics , New York: Oxford University Press.
  • –––, 2002, “Strategies for Discovering Mechanisms: Schema Instantiation, Modular Subassembly, Forward/Backward Chaining”, Philosophy of Science , 69: S354–S365.
  • –––, 2005, “Relations Among Fields: Mendelian, Cytological and Molecular Mechanisms”, Studies in History and Philosophy of Biological and Biomedical Sciences , 36: 349–371.
  • –––, 2006, Reasoning in Biological Discoveries: Mechanism, Interfield Relations, and Anomaly Resolution , New York: Cambridge University Press.
  • Darden, L. and J. Cain, 1989, “Selection Type Theories”, Philosophy of Science , 56:106–129.
  • Darden, L. and N. Maull, 1977, “Interfield Theories”, Philosophy of Science , 44: 43–64.
  • Datteri, E., 2009, “Simulation Experiments in Bionics: A Regulative Methodological Perspective”, Biology and Philosophy , 24: 301–324.
  • Datteri, E. and G. Tamburrini, 2007, “Biorobotic Experiments for the Discovery of Biological Mechanisms”, Philosophy of Science , 74: 409–430.
  • DesAutels, L., 2011, “Against Regular and Irregular Characterizations of Mechanisms”, Philosophy of Science , 78: 914–925.
  • Des Chene, D., 2001, Spirits & Clocks: Machine & Organism in Descartes , Ithaca, NY: Cornell University Press.
  • –––, 2005, “Mechanisms of Life in the Seventeenth Century: Borelli, Perrault, Régis”, Studies in the History and Philosophy of Biological and Biomedical Sciences , 36: 245–260.
  • Dijksterhuis, E.J., 1961, The Mechanization of the World Picture , New York: Oxford University Press.
  • Dowe, P., 1992, “Wesley Salmon’s Process Theory of Causality and the Conserved Quantity Theory”, Philosophy of Science , 59: 195–216.
  • –––, 2011, “The Causal-Process-Model Theory of Mechanisms”, in Illari et al. 2011: 865–879.
  • Dray, W., 1957, Laws and Explanation in History , London: Oxford University Press.
  • Dupré, J., 1993, The Disorder of Things: Metaphysical Foundations of the Disunity of Science , Cambridge, MA: Harvard University Press.
  • –––, 2013, “Living Causes”, Proceedings of the Aristotelian Society , 87: 19–35.
  • Elster, J., 1989, Nuts and Bolts for the Social Sciences , Cambridge, UK: Cambridge University Press.
  • Eronen, M.I., 2013, “No Levels, No Problems: Downward Causation in Neuroscience”, Philosophy of Science , 80: 1042–1052.
  • –––, 2015, “ Levels of Organization: A Deflationary Account”, Biology and Philosophy , 30: 39–58.
  • Fagan, M.B., 2012, “The Joint Account of Mechanistic Explanation”, Philosophy of Science , 79: 448–472.
  • –––, 2013, Philosophy of Stem Cell Biology: Knowledge in Flesh and Blood , London: Palgrave MacMillan.
  • Fehr, C., 2004, “Feminism and Science: Mechanism without Weductionism”, National Women’s Studies Association Journal , 16: 136–156.
  • Fodor, J., 1968, Psychological Explanation , New York: Random House.
  • Garson, J., 2011, “Selected Effects Functions and Causal Role Functions in the Brain: The Case for an Etiological Approach to Neuroscience”, Biology & Philosophy , 26: 547–565.
  • –––, 2012, “Function, Selection, and Construction in the Brain”, Synthese , 189: 451–481.
  • –––, 2013, “The Functional Sense of Mechanism”, Philosophy of Science , 80: 317–333.
  • Giere, R.N., 2004, “How Models Are Used to Represent Reality”, Philosophy of Science , 71: 742–752.
  • Gillett, C., 2002, “The Dimensions of Realization: A Critique of the Standard View”, Analysis , 62: 316–323.
  • –––, 2003, “The Metaphysics of Realization, Multiple Realizability, and the Special Sciences”, The Journal of Philosophy , 100: 591–603.
  • –––, 2013, “Constitution, and Multiple Constitution, in the Sciences: Using the Neuron to Construct a Starting Framework”, Minds and Machines , 23: 309–37.
  • Glennan, S.S., 1996, “Mechanisms and The Nature of Causation”, Erkenntnis , 44: 49–71.
  • –––, 1997, “Capacities, Universality and Singularity”, Philosophy of Science , 64: 605–626.
  • –––, 2002, “Rethinking Mechanistic Explanation”, Philosophy of Science , 69: S342–S353.
  • –––, 2005, “Modeling Mechanisms”, Studies in the History and Philosophy of the Biological and Biomedical Sciences , 36: 375–388.
  • –––, 2009, “Productivity, Relevance and Natural Selection”, Biology and Philosophy , 24: 325–339.
  • –––, 2010a, “Mechanisms”, in Beebee et al. 2010: 315–325.
  • –––, 2010b, “Mechanisms, Causes, and the Layered Model of the World”, Philosophy and Phenomenological Research , 81: 362–381.
  • –––, forthcoming, The New Mechanical Philosophy , Oxford: Oxford University Press.
  • Godfrey-Smith, P., 2010, “Causal Pluralism”, in Beebee et al. 2010: 326–337. [ Godfrey-Smith 2010 available online ]
  • Gould, S.J., 1990, Wonderful Life: Burgess Shale and the Nature of History , New York: W.W. Norton and Company.
  • Griffiths, P.E., 1997, What Emotions Really Are: The Problem of Psychological Categories , Chicago: University of Chicago Press.
  • Grush, R., 2003, “In Defense of Some ‘Cartesian’ Assumptions Concerning the Brain and its Operation”, Biology and Philosophy , 18: 53–93.
  • Hanson, N.R., 1958, Patterns of Discovery , Cambridge: Cambridge University Press.
  • Harbecke, J., 2010, “Mechanistic Constitution in Neurobiological Explanations”, International Studies in the Philosophy of Science , 24: 267–285.
  • –––, 2014 [onlinefirst], “Regularity Constitution and the Location of Mechanistic Levels”, Foundations of Science , 19. doi: 10.1007/s10699–014–9371–1
  • Harinen, T., forthcoming, “Mutual Manipulability and Causal Betweenness”, Synthese , doi:10.1007/s11229-014-0564-5
  • Haugeland, J., 1998, Having Thought , Cambridge, MA: Harvard University Press
  • Havstad, J.C., 2011, “Discussion: Problems for Natural Selection as a Mechanism”, Philosophy of Science , 78: 512–523.
  • Hedström, P., 2005, Dissecting the Social: On the Principles of Analytical Sociology , Cambridge: Cambridge University Press.
  • Hedström, P. and R. Swedberg, 1998, Social Mechanisms: An Analytical Approach to Social Theory , Cambridge: Cambridge University Press.
  • Hedström, P. and P. Ylikoski, 2010, “Causal Mechanisms in the Social Sciences”, Annual Review of Sociology , 36: 49–67.
  • Heil, J. and A. Mele, 1993, Mental Causation , Oxford: Clarendon Press.
  • Hempel, C.G., 1965, Aspects of Scientific Explanation and Other Essays in the Philosophy of Science , New York: Free Press.
  • Hempel, C.G. and P. Oppenheim, 1948, “Studies in the Logic of Explanation”, Philosophy of Science , 15: 135–175.
  • Hitchcock, C.R., 1995, “Discussion: Salmon on Explanatory Relevance”, Philosophy of Science , 62: 304–20.
  • Hoffmann-Kloss, V., 2014, “Interventionism and Higher-Level Causation”, International Studies in the Philosophy of Science , 28: 49–64.
  • Howick, J.P. Glasziou, and J.K. Aronson, 2013, “Problems with Using Mechanisms to Solve the Problem of Extrapolation”, Theoretical Medicine and Bioethics , 34: 275–291.
  • Huneman, P., 2010, “Topological Explanations and Robustness in Biological Sciences”, Synthese , 177: 213–245.
  • Illari, P.M., 2011, “Disambiguating the Russo-Williamson Thesis”, International Studies in the Philosophy of Science , 25: 139–157.
  • Illari, P.M., F. Russo, and J. Williamson (eds), 2011, Causality in the Sciences , Oxford: Oxford University Press.
  • Illari, P.M. and J. Williamson, 2010, “Function and Organization: Comparing the Mechanisms of Protein Synthesis and Natural Selection”, Studies in the History and Philosophy of the Biological and Biomedical Sciences , 41: 279–291.
  • –––, 2012, “What is a Mechanism?: Thinking about Mechanisms Across the Sciences”, European Journal for Philosophy of Science , 2: 119–135.
  • Kaiser, M. and C.F. Craver, 2013, “Mechanisms and Laws: Clarifying the Debate”, in Hsiang-Ke Chao, Szu-Ting Chen and Roberta L. Millstein (eds), Mechanism and Causality in Biology and Economics , Dordrecth: Springer, pp. 125–145. [ Kaiser and Craver 2013 available online ]
  • Kaplan, D.M., 2012, “How to Demarcate the Boundaries of Cognition”, Biology and Philosophy , 27: 545–570.
  • Kaplan, D.M. and W. Bechtel, W, 2011, “Dynamical Models: An Alternative or Complement to Mechanistic Explanations”, Topics , in Cognitive Science , 3: , 438–444. [ Kaplan and Bechtel 2011 available online ]
  • Kaplan, D.M. and C.F. Craver, 2011, “The Explanatory Force of Dynamical Models”, Philosophy of Science , 78: 601–627.
  • Kauffman, S.A., 1971, “Articulation of Parts Explanation in Biology and the Rational Search for Them”, in Roger C. Buck and Robert S. Cohen (eds), PSA 1970, Boston Studies in the Philosophy of Science , volume 8. Dordrecht: Reidel, pp. 257–272. Reprinted in Marjorie Grene and Everett Mendelsohn (eds), 1976, Topics in the Philosophy of Biology , Dordrecht: Reidel, pp. 245–263.
  • Kendler, K., P. Zachar, and C.F. Craver, 2010, “What Kinds of Things are Psychiatric Disorders?”, Psychological Medicine , 41: 1143–1150.
  • Khalidi, M.A., 2013, Natural Categories and Human Kinds , Cambridge: Cambridge University Press.
  • Kim, J., 1998, Mind in a Physical World , Cambridge, MA: The MIT Press.
  • Krickel, B., 2014, The Metaphysics of Mechanism , PhD Dissertation. Humboldt-Universität zu Berlin.
  • Kornblith, H., 1993, Inductive Inference and Its Natural Ground , Cambridge, MA: MIT Press.
  • Kuhlmann, M., 2011, “Mechanisms in Dynamically Complex Systems”, in Illari et al. 2011: 880–906.
  • Kuhlmann, M. and S. Glennan, 2014, “On the Relation between Quantum Mechanical and Neo-Mechanistic Ontologies and Explanatory Strategies”, European Journal for Philosophy of Science , 4: 337–359.
  • Kuhn, T.S., 1962, The Structure of Scientific Revolutions , Chicago: University of Chicago Press.
  • Kuorikoski, J. and S. Pöyhönen, 2012, “Looping Kinds and Social Mechanisms”, Sociological Theory , 30: 187–205.
  • Kuorikoski, J. and P.K. Ylikoski, 2013, EPSA11 Perspectives and Foundational Problems in Philosophy of Science , V. Karakostas & D. Dieks (eds), Heidelberg: Springer, pp. 69–80.
  • Lakatos, I., 1977, The Methodology of Scientific Research Programmes: Philosophical Papers , vol. 1. Cambridge; Cambridge University Press.
  • Laudan, L., 1977, Progress and Its Problems , Berkeley: University of California Press.
  • Leuridan, B., 2010, “Can Mechanisms Really Replace Laws of Nature?”, Philosophy of Science , 77: 317–340.
  • –––, 2011, “Three Problems for the Mutual Manipulability Account of Constitutive Relevance in Mechanisms”, The British Journal for the Philosophy of Science , 63: 399–427.
  • Levy, A., 2013, “Three Kinds of ‘New Mechanism’”, Biology and Philosophy , 28: 99–114.
  • –––, 2014, “What was Hodgkin and Huxley’s Achievement?”, British Journal for the Philosophy of Science , 65(3): 469–492. doi:10.1093/bjps/axs043
  • Levy, A. and W. Bechtel, 2012, “Abstraction and the Organization of Mechanisms”, Philosophy of Science , 80: 241–261.
  • Little, D., 1991, Varieties of Social Explanation: An Introduction to the Philosophy of Social Science , Boulder, CO: Westview.
  • –––, 1998, Microfoundations, Method, and Causation: On the Philosophy of the Social Science , New Brunswick, NJ: Transaction.
  • Longino, H., 2013, Studying Human Behavior: How Scientists Investigate Aggression and Sexuality , Chicago: University of Chicago Press.
  • –––, 2014, “Pluralism, Social Action, and the Causal Space of Human Behavior”, Metascience , 23: 443–459.
  • Machamer, P.K., 1977, “Teleology and Selective Processes”, in R. Colodny (ed.), Logic, Laws, and Life: Some Philosophical Complications , Pittsburgh, PA: University of Pittsburgh Press, pp. 129–142.
  • Machamer, P., 2004, “Activities and Causation: The Metaphysics and Epistemology of Mechanisms”, International Studies in the Philosophy of Science , 18: 27–39.
  • Machamer, P.K., L. Darden, and C.F. Craver, 2000 [MDC], “Thinking about Mechanisms”, Philosophy of Science , 67:1–25.
  • Machery, E., 2009, Doing Without Concepts , New York: Oxford University Press.
  • Mackie, J.L., 1974, The Cement of the Universe , Oxford: Clarendon Press.
  • Maley, C.J. and G. Piccinini, forthcoming, “A Unified Mechanistic Account of Teleological Functions for Psychology and Neuroscience”, in David Kaplan (ed.), Integrating Mind and Brain Science: Mechanistic Perspectives and Beyond , Oxford: Oxford University Press.
  • Matthews, L., forthcoming, “Embedded Mechanisms and Phylogenetics”, Philosophy of Science .
  • Matthewson, J. and B. Calcott, 2011, “Mechanistic Models of Population-Level Phenomena”, Biology and Philosophy , 26: 737–756.
  • McCauley, R.N. and W. Bechtel, 2001, “Explanatory Pluralism and the Heuristic Identity Theory”, Theory and Psychology , 11: 736–760.
  • Melnyk, A., 2003, A Physicalist Manifesto: Thoroughly Modern Materialism , Cambridge: Cambridge University Press.
  • Melnyk, A., 2010, “Comments on Sydney Shoemaker’s Physical Realization ”, Philosophical Studies , 148: 113–123.
  • Menzies, P., 2012, “The Causal Structure of Mechanisms”, Studies in History and Philosophy of Biological and Biomedical Sciences , 43: 796–805.
  • Milkowski, M., 2013, Explaining the Computational Mind , Cambridge, MA: MIT Press.
  • Millstein, R.L., 2006, “Natural Selection as a Population-Level Causal Process”, The British Journal for the Philosophy of Science , 57: 627 -653.
  • Milo, R, S. Shen-Orr, S. Itzkovitz, N. Kashtan, D. Chklovskii, and U. Alon, 2002, “Network Motifs: Simple Building Blocks of Complex Networks”, Science , 298(5594): 824–27.
  • Mitchell, S.D., 1997, “Pragmatic Laws”, Philosophy of Science , 64: S468–S479.
  • –––, 2000, “Dimensions of Scientific Law”, Philosophy of Science 67: 242–265.
  • –––, 2003, Biological Complexity and Integrative Pluralism , Cambridge: Cambridge University Press.
  • –––, 2005, “Modularity: More than a Buzzword? Essay Review, Biological Theory , 1: 98–101.
  • –––, 2009, Unsimple Truths: Science, Complexity, and Policy , Chicago: University of Chicago Press.
  • Moss, L., 2012, “Is the Philosophy of Mechanism Philosophy Enough?” Studies in the History and Philosophy of Science: C , 43: 164–72.
  • Nagel, E., 1961, The Structure of Science: Problems in the Logic of Scientific Explanation , New York: Harcourt, Brace and World.
  • Neander, K., 1991a, “The Teleological Notion of ‘Function’”, Australasian Journal of Philosophy , 69: 454–468.
  • –––, 1991b, “Functions as Selected Effects: the Conceptual Analyst’s Defence”, Philosophy of Science , 58: 168–184.
  • Nicholson, D.J., 2012, “The Concept of Mechanism in Biology”, Studies in History and Philosophy of Biological and Biomedical Sciences , 43: 152–163.
  • Nickels, T., 1985, “Beyond Divorce: Current Status of the Discovery Debate”, Philosophy of Science , 52: 177–206.
  • Oppenheim, P. and H. Putnam, 1958, “The Unity of Science as a Working Hypothesis”, in H. Feigl, M. Scriven and G. Maxwell (eds), Concepts, Theories and the Mind-Body Problem , (Minnesota Studies in the Philosophy of Science, v. 2), Minneapolis: University of Minnesota Press, pp. 3–36.
  • Pearl, J., 2009, Causality: Models, Reasoning, and Inference , second edition, New York: Cambridge University Press.
  • Persson, Johannes, 2010, “Activity-Based Accounts of Mechanism and the Threat of Polygenic Effects”, Erkenntnis , 72: 135–149.
  • Piccinini, G., 2007, “Computing Mechanisms”, Philosophy of Science , 74: 501–526.
  • Piccinini, G. and C.F. Craver, 2011, “Integrating Psychology and Neuroscience: Functional Analyses as Mechanism Sketches”, Synthese , 183: 283–3
  • Piccinini, G. and A. Scarantino, 2011, “Information Processing, Computation, and Cognition”, Journal of Biological Physics , 37: 1–38.
  • Polger, T.W., 2007, “Realization and the Metaphysics of Mind”, Australasian Journal of Philosophy , 85: 233–59.
  • –––, 2010, “Mechanisms and Explanatory Realization Relations”, Synthese , 177: 193–212.
  • Popper, K., 1959, The Logic of Scientific Discovery , London: Hutchinson and Co.
  • Povich, M., forthcoming, “Mechanisms and Model-based fMRI”, Philosophy of Science .
  • Psillos, S., 2004, “A Glimpse of the Secret Connexion: Harmonizing Mechanism with Counterfactuals”, Perspectives on Science , 12: 288–319.
  • Ramsey, J.L., 2008, “Mechanisms and Their Explanatory Challenges in Organic Chemistry”, Philosophy of Science , 75: 970–982.
  • Reichenbach, H., 1938, Experience and Prediction: An Analysis of the Foundations and the Structure of Knowledge , Chicago: The University of Chicago Press.
  • Richardson, R.C. and A. Stephan, 2007, “Emergence”, Biological Theory , 2: 91–96.
  • Roe, S., 2014, The Salmon-Roe Approach to Mechanistic Explanations , Doctoral Dissertation, UC Davis.
  • Romero, F., forthcoming, “Why there isn’t interlevel causation in mechanisms”, Synthese . doi:10.1007/s11229-015-0718-0
  • Rosenberg, A., 1985, The Structure of Biological Science , Cambridge: Cambridge University Press.
  • Russell, B., 1913, “On the Notion of Cause”, Proceedings of the Aristotelian Society , 13: 1–26.
  • Russo, F., 2009, Causality and Causal Modeling in the Social Sciences: Measuring Variation , Dordrecht: Springer.
  • Russo, F. and J. Williamson, 2007, “Interpreting Causality in the Health Sciences”, International Studies in the Philosophy of Science , 21: 157–170.
  • Salmon, W.C., 1984, Scientific Explanation and the Causal Structure of the World , Princeton, NJ: Princeton University Press.
  • –––, 1989, Four Decades of Scientific Explanation , Minneapolis, MN: University of Minnesota Press.
  • –––, 1994, “Causality Without Counterfactuals”, Philosophy of Science , 61: 297–312.
  • –––, 1997, “Causality and Explanation: A Reply to Two Critiques”, Philosophy of Science , 64: 461–77.
  • Schaffer, J., 2000, “Causation by Disconnection”, Philosophy of Science , 67: 285–300.
  • –––, 2004, “Causes Need Not be Physically Connected to Their Effects: The Case for Negative Causation”, in C. Hitchcock (ed.), Contemporary Debates in Philosophy of Science , Malden, MA: Blackwell Publishing. Pp. 197–216.
  • Schaffner, K., 1993, Discovery and Explanation in Biology and Medicine , Chicago, IL: University of Chicago Press.
  • Schlosser, G. and G.P. Wagner, 2004, Modularity in Development and Evolution , Chicago, IL: University of Chicago Press.
  • Scriven, M., 1959, “Explanation and Prediction in Evolutionary Theory”, Science , 130: 477–482.
  • Shagrir, O., 2010, “Computation, San Diego Style”, Philosophy of Science , 77: 862–874.
  • Shapiro, L., 2000, “Multiple Realizations”, Journal of Philosophy , 97: 635–654.
  • Shoemaker, S., 2003, “Realization, Micro-Realization, and Coincidence”, Philosophy and Phenomenological Research , 67: 1–23.
  • –––, 2007, Physical Realization , Oxford: Oxford University Press.
  • Simon, H.A., 1996 [1962], The Sciences of the Artificial , 3 rd ed., Cambridge, MA: MIT Press.
  • Skipper Jr., R.A. and R.L. Millstein, 2005, “Thinking about Evolutionary Mechanisms: Natural Selection”, in Craver and Darden 2005: 327–347.
  • Spirtes, P., C. Glymour, and R. Scheines, 2000, Causation, Prediction, and Search , 2 nd ed., Cambridge, MA: MIT Press.
  • Steel, D.P., 2008, Across the Boundaries: Extrapolation in Biology and Social Science , New York: Oxford University Press.
  • Strevens, M., 2008, Depth: An Account of Scientific Explanation , Cambridge, MA: Harvard University Press.
  • Sullivan, Jacqueline, 2009, “The Multiplicity of Experimental Protocols: A Challenge to Reductionist and Non-Reductionist Models of the Unity of Neuroscience”, Synthese , 167: 511–539.
  • Tabery, J., 2004, “Synthesizing Activities and Interactions in the Concept of a Mechanism”, Philosophy of Science , 71: 1–15.
  • –––, 2009, “Difference Mechanisms: Explaining Variation with Mechanisms”, Biology and Philosophy , 24: 645–664.
  • –––, 2014a, Beyond Versus: The Struggle to Understand the Interaction of Nature and Nurture , Cambridge, MA: The MIT Press.
  • –––, 2014b, “Pluralism, Social Action, and the Causal Space of Human Behavior”, Metascience , 23: 443–459.
  • Teller, P., 2010, “Mechanism, Reduction, and Emergence in Two Stories of the Human Epistemic Enterprise”, Erkenntnis , 73: 413–425.
  • Thagard, P., 1998, “Explaining Disease: Causes, Correlations, and Mechanisms”, Minds and Machines , 8: 61–78.
  • –––, 2000, How Scientists Explain Disease , Princeton, NJ: Princeton University Press.
  • –––, 2006, Hot Thought: Mechanisms and Applications of Emotional Cognition , Cambridge, MA: The MIT Press.
  • Thalos, M., 2013, Without Hierarchy : An Essay on the Scale Freedom of the Universe , Oxford: Oxford University Press.
  • Theurer, K.L., 2013, “Compositional Explanatory Relations and Mechanistic Reduction”, Minds and Machines , 23: 287–307.
  • Weber, M., 2005, Philosophy of Experimental Biology , Cambridge: Cambridge University Press.
  • –––, 2008, “Causes without Mechanisms: Experimental Regularities, Physical Laws, and Neuroscientific Explanation”, Philosophy of Science , 75: 995–1007.
  • Weisberg, M., 2007, “Three Kinds of Idealization”, The Journal of Philosophy , 104: 639–659.
  • –––, 2013, Simulation and Similarity: Using Models to Understand the World , Oxford: Oxford University Press.
  • Weiskopf, D., 2011, “Models and Mechanisms in Psychological Explanation”, Synthese , 183: 313–338.
  • Westfall, R., 1971, The Construction of Modern Science , Cambridge: Cambridge University Press.
  • –––, 2005, Genes and the Agents of Life , Cambridge: Cambridge University Press.
  • Wimsatt, W.C., 1972a, “Complexity and Organization”, in Kenneth F. Schaffner and Robert S. Cohen (eds), PSA 1972, Proceedings of the Philosophy of Science Association , Dordrecht: Reidel, pp. 67–86.
  • –––, 1972b, “Teleology and the Logical Structure of Function Statements”, Studies in the History and Philosophy of Science , 3: 1–80.
  • –––, 1976, “Reductionism, Levels of Organization, and the Mind–Body Problem”, in G. Globus, I. Savodnik, and G. Maxwell (eds), Consciousness and the Brain , New York: Plenum Press, pp. 199–267.
  • –––, 1997, “Aggregativity: Reductive Heuristics for Finding Emergence”, Philosophy of Science , 64: S372–S384.
  • –––, 2007, Re-Engineering Philosophy for Limited Beings: Piecewise Approximations to Reality , Cambridge, MA: Harvard University Press.
  • Woodward, J., 1989, “The Causal/Mechanical Model of Explanation”, in P. Kitcher and W.C. Salmon (eds), Scientific Explanation , (Minnesota Studies in the Philosophy of Science 13), pp. 357–383.
  • –––, 2001, “Law and Explanation in Biology: Invariance Is the Kind of Stability that Matters”, Philosophy of Science , 68: 1–20.
  • –––, 2002, “What Is a Mechanism?: A Counterfactual Account”, Philosophy of Science , 69: S366–S377.
  • –––, 2003, Making Things Happen: A Theory of Causal Explanation , New York: Oxford University Press.
  • –––, 2011, “Mechanisms Revisited”, Synthese , 183: 409–427.
  • –––, 2014, “A Functional Account of Causation; or, A Defense of the Legitimacy of Causal Thinking by Reference to the Only Standard That Matters—Usefulness (as Opposed to Metaphysics or Agreement with Intuitive Judgment)”, Philosophy of Science , 81: 691–713.
  • Wright, L., 1973, “Functions”, Philosophical Review , 82: 139–168.
  • Ylikoski, P., 2013, “Causal and Constitutive Explanation Compared”, Erkenntnis , 78: 277–97.
How to cite this entry . Preview the PDF version of this entry at the Friends of the SEP Society . Look up topics and thinkers related to this entry at the Internet Philosophy Ontology Project (InPhO). Enhanced bibliography for this entry at PhilPapers , with links to its database.

[Please contact the author with suggestions.]

Anscombe, Gertrude Elizabeth Margaret | biology: experiment in | causation: and manipulability | causation: counterfactual theories of | causation: the metaphysics of | Descartes, René | empiricism: logical | Helmholtz, Hermann von | individualism, methodological | models in science | molecular biology | natural kinds | Newton, Isaac | physicalism | physics: intertheory relations in | reduction, scientific | reduction, scientific: in biology | Reichenbach, Hans | science: unity of | scientific discovery | scientific explanation | teleology: teleological notions in biology

Acknowledgments

We are grateful to the following individuals for very helpful feedback on earlier drafts of this entry: Lindley Darden, Melinda Fagan, Stuart Glennan, Phyllis Illari, Beate Krickel, Lucas Matthews, Irena Mikhalevich, Anya Plutynski, Gualtiero Piccinini, Mark Povich, and Katheryn Shrumm. We are also grateful to Pamela Speh for help with Figure 1.

Copyright © 2015 by Carl Craver < ccraver @ wustl . edu > James Tabery < james . tabery @ utah . edu >

  • Accessibility

Support SEP

Mirror sites.

View this site from another server:

  • Info about mirror sites

The Stanford Encyclopedia of Philosophy is copyright © 2023 by The Metaphysics Research Lab , Department of Philosophy, Stanford University

Library of Congress Catalog Data: ISSN 1095-5054

  • Subject List
  • Take a Tour
  • For Authors
  • Subscriber Services
  • Publications
  • African American Studies
  • African Studies
  • American Literature
  • Anthropology
  • Architecture Planning and Preservation
  • Art History
  • Atlantic History
  • Biblical Studies
  • British and Irish Literature
  • Childhood Studies
  • Chinese Studies
  • Cinema and Media Studies
  • Communication
  • Criminology
  • Environmental Science
  • Evolutionary Biology
  • International Law
  • International Relations
  • Islamic Studies
  • Jewish Studies
  • Latin American Studies
  • Latino Studies
  • Linguistics
  • Literary and Critical Theory
  • Medieval Studies
  • Military History
  • Political Science
  • Public Health
  • Renaissance and Reformation
  • Social Work
  • Urban Studies
  • Victorian Literature
  • Browse All Subjects

How to Subscribe

  • Free Trials

In This Article Expand or collapse the "in this article" section Mechanisms in Science

Introduction, central articles and books.

  • Overview Articles and Reference Works
  • Predecessors/Prehistory
  • What Are Mechanisms?
  • Discovery and Evidence
  • Mechanisms, Reduction, and Interfield Integration
  • Mechanistic Explanation (General Background)
  • Mechanisms and Causation
  • Mechanisms and Laws of Nature
  • Mechanisms and Natural Kinds
  • Functionalism, Abstraction, and Mechanisms
  • Levels of Mechanisms and Interlevel Causation
  • Dynamical Models, Minimal Models, and Network Models: Conflict or Cooperation?
  • Objections from the Failure of Localization and Decomposition as a Strategy

Related Articles Expand or collapse the "related articles" section about

About related articles close popup.

Lorem Ipsum Sit Dolor Amet

Vestibulum ante ipsum primis in faucibus orci luctus et ultrices posuere cubilia Curae; Aliquam ligula odio, euismod ut aliquam et, vestibulum nec risus. Nulla viverra, arcu et iaculis consequat, justo diam ornare tellus, semper ultrices tellus nunc eu tellus.

  • Experiments in Physics
  • Idealizations in Science
  • Laws of Nature
  • Mathematical Explanation
  • Models and Theories in Science
  • Natural Kinds
  • Paul Feyerabend
  • Scientific Explanation
  • Scientific Progress
  • Scientific Realism
  • Scientific Representation
  • Scientific Revolutions
  • Theoretical Terms in Science

Other Subject Areas

Forthcoming articles expand or collapse the "forthcoming articles" section.

  • Alfred North Whitehead
  • Feminist Aesthetics
  • Find more forthcoming articles...
  • Export Citations
  • Share This Facebook LinkedIn Twitter

Mechanisms in Science by Carl Craver , Beate Krickel LAST REVIEWED: 30 October 2019 LAST MODIFIED: 30 October 2019 DOI: 10.1093/obo/9780195396577-0395

Many areas of science are animated by the search for mechanisms: experiments are designed to find them, explanations are built to reveal them, models are constructed to describe them, funding is disbursed to prioritize their discovery, and translational research is premised on their value for manipulation and control. Over the last twenty or thirty years, a number of philosophers of science, starting from the biological and neural sciences, have directed attention at the concept of mechanism. They have emphasized that mechanisms play central roles in discovery, explanation, experimentation, modeling, and reduction. At the same time, the idea of mechanism has served for many as a suggestive hint as to the metaphysics of the middle range: covering that domain of phenomena above the size scale of atomic physics and beneath that of planets, encompassing biology, physiology, psychology, and the human and so-called special sciences more generally. What must mechanisms be if they are to play these diverse roles? The term “new mechanism” (to distinguish the modern account from the classical mechanical worldview defended by philosophers such as Descartes or La Mettrie), introduced to describe this research area, runs the risk of homogenizing what has become a heterogeneous body of work, serving many masters and tugging the analysis in many directions at once. Here, we attempt to collect some points of consensus while highlighting areas of productive disagreement and criticism going forward. The authors thank Stuart Glennan and Lindley Darden for feedback on an earlier draft, Sue McKinney for assistance formatting references, and Paola Hernandez-Chavez for support in finding non-English texts.

This section contains central articles and books of notable historical significance within the new mechanism for understanding what the new mechanism sought to achieve and the diverse areas of the philosophy of science in which it was thought to do useful service. Bechtel and Richardson 1993 sketches a road map (or decision-tree) that researchers face in the search for mechanisms. The authors focus specifically on the effort to decompose complex functions into subfunctions, and to localize those subfunctions to different locations in a system, such as a cell or a body. These strategies, they point out, make sense against the background idea that the scientists aim, in their research, to discover mechanisms. Bechtel and Abrahamsen 2005 extends this work to offer a mechanistic theory of explanation opposed to the idea that explanations require knowing the laws of nature. Glennan 1996 also deploys the concept of mechanism, but in the service of a philosophical analysis of causation. For Glennan, mechanisms are the hidden connection Hume sought between a cause and its effect. He has subsequently refined and revised his views in light of criticism and theory development (see Glennan 2002 for contrasts with Salmon’s mechanistic view; see also Glennan 2017 for its latest development, emphasizing nominalism and an activities-based view of causation). Machamer, Darden, and Craver (commonly known as MDC) were less concerned with a particular set of research strategies and with Hume’s search for a hidden connection between cause and effect than they were with the suggestion that attention to mechanisms could revolutionize the philosophy of science, transforming discussions of causation, discovery, functions, laws, levels, and reduction. They characterize mechanisms as entities and activities organized in the production of regular changes from start or set-up conditions to finish and termination conditions, and they argue that the drive to satisfy this explanatory demand shapes the practice of the biological and neural sciences (see Machamer, et al. 2000 ). Following this publication, Darden pursued the issue of mechanism discovery, both alone and in cooperation with Craver and others. Key papers in this development are compiled in Darden 2006 . Darden’s view perhaps has its fullest development in Craver and Darden 2013 . Craver, in contrast, began to work on the relevance of mechanism to the topic of scientific explanation. In Craver 2001 he explores how Cummins’ idea of an analytic account of functional explanation applies to the multilevel mechanistic theories of the physiological sciences. His most systematic treatment of the topic can be found in Craver 2007 , which became a focus of much subsequent discussion.

Bechtel, William, and Adele Abrahamsen. “Explanation: A Mechanist Alternative.” In Special Issue: Mechanisms in Biology . Edited by Carl F. Craver and Lindley Darden. Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences 36.2 (2005): 421–441.

DOI: 10.1016/j.shpsc.2005.03.010

Bechtel and Abrahamsen set the grounds for the epistemic interpretation of mechanisms as explanatory models. They discuss the benefits of this view over law-based accounts of explanation.

Bechtel, William, and Robert C. Richardson. Discovering Complexity: Decomposition and Localization as Strategies in Scientific Research . Princeton, NJ: Princeton University Press, 1993.

Analyzes the heuristics of decomposition and localization that are crucial for the discovery of mechanisms and in the development of mechanistic models in cell biology, cognitive neuroscience, and genetics.

Craver, Carl F. “Role Functions, Mechanisms and Hierarchy.” Philosophy of Science 68 (2001): 31–55.

DOI: 10.1086/392866

Craver unifies Cummins’ view of functions with his multilevel mechanistic picture of explanation ( Cummins 1975 ), arguing that contextual role functions constitute a third aspect of causal-mechanical explanation, beyond etiological and constitutive.

Craver, Carl F. Explaining the Brain: Mechanisms and the Mosaic Unity of Neuroscience . New York: Oxford University Press, 2007.

DOI: 10.1093/acprof:oso/9780199299317.001.0001

Discusses the nature of explanations in neuroscience and defends the view that good neuroscientific explanations are descriptions of multilevel mechanisms.

Craver, Carl F., and Lindley Darden. In Search of Mechanisms: Discoveries across the Life Sciences . Chicago: University of Chicago Press, 2013.

DOI: 10.7208/chicago/9780226039824.001.0001

Explains the relevance of mechanism to discovery, with detailed examples and case-studies from the history of biology and contemporary biology, including the neural sciences. Details stages of mechanism discovery and strategies for solving specific discovery problems.

Darden, Lindley. Reasoning in Biological Discoveries: Essays on Mechanism, Interfield Relations, and Anomaly Resolution . New York: Cambridge University Press, 2006.

DOI: 10.1017/CBO9780511498442

Contains many of Darden’s papers, alone and with coauthors, on how the search for mechanisms constrains scientific discovery, in addition to Darden’s earlier work on interfield theories and an anomaly-driven theory change.

Glennan, Stuart. “Mechanisms and the Nature of Causation.” Erkenntnis 44 (1996): 49–71.

DOI: 10.1007/BF00172853

Proposes a mechanistic theory of causation according to which A and B causally interact only if there is a mechanism between them.

Glennan, Stuart. “Rethinking Mechanistic Explanation.” Philosophy of Science 69.S3 (2002): S342–S353.

DOI: 10.1086/341857

Develops Glennan’s complex system view of mechanisms and contrasts it with Wesley Salmon’s and Peter Railton’s accounts.

Glennan, Stuart. The New Mechanical Philosophy . Oxford: Oxford University Press, 2017.

DOI: 10.1093/oso/9780198779711.001.0001

See especially Glennan’s discussion in chapter 2 of the many ways of solving problems of both defining and demarcating mechanisms.

Machamer, Peter, Lindley Darden, and Carl F. Craver. “Thinking About Mechanisms.” Philosophy of Science 67.1 (2000): 1–25.

DOI: 10.1086/392759

This is probably the most prominent article of the new mechanistic literature. It provides starting points for many discussions, such as how to characterize mechanisms and how to describe the relationship between mechanisms and mechanistic explanation and the role of mechanisms in discovery.

back to top

Users without a subscription are not able to see the full content on this page. Please subscribe or login .

Oxford Bibliographies Online is available by subscription and perpetual access to institutions. For more information or to contact an Oxford Sales Representative click here .

  • About Philosophy »
  • Meet the Editorial Board »
  • A Priori Knowledge
  • Abduction and Explanatory Reasoning
  • Abstract Objects
  • Addams, Jane
  • Adorno, Theodor
  • Aesthetic Hedonism
  • Aesthetics, Analytic Approaches to
  • Aesthetics, Continental
  • Aesthetics, Environmental
  • Aesthetics, History of
  • African Philosophy, Contemporary
  • Alexander, Samuel
  • Analytic/Synthetic Distinction
  • Anarchism, Philosophical
  • Animal Rights
  • Anscombe, G. E. M.
  • Anthropic Principle, The
  • Anti-Natalism
  • Applied Ethics
  • Aquinas, Thomas
  • Argument Mapping
  • Art and Emotion
  • Art and Knowledge
  • Art and Morality
  • Astell, Mary
  • Aurelius, Marcus
  • Austin, J. L.
  • Bacon, Francis
  • Bayesianism
  • Bergson, Henri
  • Berkeley, George
  • Biology, Philosophy of
  • Bolzano, Bernard
  • Boredom, Philosophy of
  • British Idealism
  • Buber, Martin
  • Buddhist Philosophy
  • Burge, Tyler
  • Business Ethics
  • Camus, Albert
  • Canterbury, Anselm of
  • Carnap, Rudolf
  • Cavendish, Margaret
  • Chemistry, Philosophy of
  • Childhood, Philosophy of
  • Chinese Philosophy
  • Cognitive Ability
  • Cognitive Phenomenology
  • Cognitive Science, Philosophy of
  • Coherentism
  • Communitarianism
  • Computational Science
  • Computer Science, Philosophy of
  • Computer Simulations
  • Comte, Auguste
  • Conceptual Role Semantics
  • Conditionals
  • Confirmation
  • Connectionism
  • Consciousness
  • Constructive Empiricism
  • Contemporary Hylomorphism
  • Contextualism
  • Contrastivism
  • Cook Wilson, John
  • Cosmology, Philosophy of
  • Critical Theory
  • Culture and Cognition
  • Daoism and Philosophy
  • Davidson, Donald
  • de Beauvoir, Simone
  • de Montaigne, Michel
  • Decision Theory
  • Deleuze, Gilles
  • Derrida, Jacques
  • Descartes, René
  • Descartes, René: Sensory Representations
  • Descriptions
  • Dewey, John
  • Dialetheism
  • Disagreement, Epistemology of
  • Disjunctivism
  • Dispositions
  • Divine Command Theory
  • Doing and Allowing
  • du Châtelet, Emilie
  • Dummett, Michael
  • Dutch Book Arguments
  • Early Modern Philosophy, 1600-1750
  • Eastern Orthodox Philosophical Thought
  • Education, Philosophy of
  • Engineering, Philosophy and Ethics of
  • Environmental Philosophy
  • Epistemic Basing Relation
  • Epistemic Defeat
  • Epistemic Injustice
  • Epistemic Justification
  • Epistemic Philosophy of Logic
  • Epistemology
  • Epistemology and Active Externalism
  • Epistemology, Bayesian
  • Epistemology, Feminist
  • Epistemology, Internalism and Externalism in
  • Epistemology, Moral
  • Epistemology of Education
  • Ethical Consequentialism
  • Ethical Deontology
  • Ethical Intuitionism
  • Eugenics and Philosophy
  • Events, The Philosophy of
  • Evidence-Based Medicine, Philosophy of
  • Evidential Support Relation In Epistemology, The
  • Evolutionary Debunking Arguments in Ethics
  • Evolutionary Epistemology
  • Experimental Philosophy
  • Explanations of Religion
  • Extended Mind Thesis, The
  • Externalism and Internalism in the Philosophy of Mind
  • Faith, Conceptions of
  • Feminist Philosophy
  • Feyerabend, Paul
  • Fichte, Johann Gottlieb
  • Fictionalism
  • Fictionalism in the Philosophy of Mathematics
  • Film, Philosophy of
  • Foot, Philippa
  • Foreknowledge
  • Forgiveness
  • Formal Epistemology
  • Foucault, Michel
  • Frege, Gottlob
  • Gadamer, Hans-Georg
  • Geometry, Epistemology of
  • God and Possible Worlds
  • God, Arguments for the Existence of
  • God, The Existence and Attributes of
  • Grice, Paul
  • Habermas, Jürgen
  • Hart, H. L. A.
  • Heaven and Hell
  • Hegel, Georg Wilhelm Friedrich: Aesthetics
  • Hegel, Georg Wilhelm Friedrich: Metaphysics
  • Hegel, Georg Wilhelm Friedrich: Philosophy of History
  • Hegel, Georg Wilhelm Friedrich: Philosophy of Politics
  • Heidegger, Martin: Early Works
  • Hermeneutics
  • Higher Education, Philosophy of
  • History, Philosophy of
  • Hobbes, Thomas
  • Horkheimer, Max
  • Human Rights
  • Hume, David: Aesthetics
  • Hume, David: Moral and Political Philosophy
  • Husserl, Edmund
  • Identity in Physics
  • Imagination
  • Imagination and Belief
  • Immanuel Kant: Political and Legal Philosophy
  • Impossible Worlds
  • Incommensurability in Science
  • Indian Philosophy
  • Indispensability of Mathematics
  • Inductive Reasoning
  • Instruments in Science
  • Intellectual Humility
  • Intentionality, Collective
  • James, William
  • Japanese Philosophy
  • Kant and the Laws of Nature
  • Kant, Immanuel: Aesthetics and Teleology
  • Kant, Immanuel: Ethics
  • Kant, Immanuel: Theoretical Philosophy
  • Kierkegaard, Søren
  • Knowledge-first Epistemology
  • Knowledge-How
  • Kristeva, Julia
  • Kuhn, Thomas S.
  • Lacan, Jacques
  • Lakatos, Imre
  • Langer, Susanne
  • Language of Thought
  • Language, Philosophy of
  • Latin American Philosophy
  • Legal Epistemology
  • Legal Philosophy
  • Legal Positivism
  • Leibniz, Gottfried Wilhelm
  • Levinas, Emmanuel
  • Lewis, C. I.
  • Literature, Philosophy of
  • Locke, John
  • Locke, John: Identity, Persons, and Personal Identity
  • Lottery and Preface Paradoxes, The
  • Machiavelli, Niccolò
  • Martin Heidegger: Later Works
  • Martin Heidegger: Middle Works
  • Material Constitution
  • Mathematical Pluralism
  • Mathematical Structuralism
  • Mathematics, Ontology of
  • Mathematics, Philosophy of
  • Mathematics, Visual Thinking in
  • McDowell, John
  • McTaggart, John
  • Meaning of Life, The
  • Mechanisms in Science
  • Medically Assisted Dying
  • Medicine, Contemporary Philosophy of
  • Medieval Logic
  • Medieval Philosophy
  • Mental Causation
  • Merleau-Ponty, Maurice
  • Meta-epistemological Skepticism
  • Metaepistemology
  • Metametaphysics
  • Metaphilosophy
  • Metaphysical Grounding
  • Metaphysics, Contemporary
  • Metaphysics, Feminist
  • Midgley, Mary
  • Mill, John Stuart
  • Mind, Metaphysics of
  • Modal Epistemology
  • Montesquieu
  • Moore, G. E.
  • Moral Contractualism
  • Moral Naturalism and Nonnaturalism
  • Moral Responsibility
  • Multiculturalism
  • Murdoch, Iris
  • Music, Analytic Philosophy of
  • Nationalism
  • Naturalism in the Philosophy of Mathematics
  • Naïve Realism
  • Neo-Confucianism
  • Neuroscience, Philosophy of
  • Nietzsche, Friedrich
  • Nonexistent Objects
  • Normative Ethics
  • Normative Foundations, Philosophy of Law:
  • Normativity and Social Explanation
  • Objectivity
  • Occasionalism
  • Ontological Dependence
  • Ontology of Art
  • Ordinary Objects
  • Other Minds
  • Panpsychism
  • Particularism in Ethics
  • Pascal, Blaise
  • Paternalism
  • Peirce, Charles Sanders
  • Perception, Cognition, Action
  • Perception, The Problem of
  • Perfectionism
  • Persistence
  • Personal Identity
  • Phenomenal Concepts
  • Phenomenal Conservatism
  • Phenomenology
  • Philosophy for Children
  • Photography, Analytic Philosophy of
  • Physicalism
  • Physicalism and Metaphysical Naturalism
  • Physics, Experiments in
  • Political Epistemology
  • Political Obligation
  • Political Philosophy
  • Popper, Karl
  • Pornography and Objectification, Analytic Approaches to
  • Practical Knowledge
  • Practical Moral Skepticism
  • Practical Reason
  • Probabilistic Representations of Belief
  • Probability, Interpretations of
  • Problem of Divine Hiddenness, The
  • Problem of Evil, The
  • Propositions
  • Psychology, Philosophy of
  • Quine, W. V. O.
  • Racist Jokes
  • Rationalism
  • Rationality
  • Rawls, John: Moral and Political Philosophy
  • Realism and Anti-Realism
  • Realization
  • Reasons in Epistemology
  • Reductionism in Biology
  • Reference, Theory of
  • Reid, Thomas
  • Reliabilism
  • Religion, Philosophy of
  • Religious Belief, Epistemology of
  • Religious Experience
  • Religious Pluralism
  • Ricoeur, Paul
  • Risk, Philosophy of
  • Rorty, Richard
  • Rousseau, Jean-Jacques
  • Rule-Following
  • Russell, Bertrand
  • Ryle, Gilbert
  • Sartre, Jean-Paul
  • Schopenhauer, Arthur
  • Science and Religion
  • Science, Theoretical Virtues in
  • Scotus, Duns
  • Self-Knowledge
  • Sellars, Wilfrid
  • Semantic Externalism
  • Semantic Minimalism
  • Senses, The
  • Sensitivity Principle in Epistemology
  • Shepherd, Mary
  • Singular Thought
  • Situated Cognition
  • Situationism and Virtue Theory
  • Skepticism, Contemporary
  • Skepticism, History of
  • Slurs, Pejoratives, and Hate Speech
  • Smith, Adam: Moral and Political Philosophy
  • Social Aspects of Scientific Knowledge
  • Social Epistemology
  • Social Identity
  • Sounds and Auditory Perception
  • Space and Time
  • Speech Acts
  • Spinoza, Baruch
  • Stebbing, Susan
  • Strawson, P. F.
  • Structural Realism
  • Supererogation
  • Supervenience
  • Tarski, Alfred
  • Technology, Philosophy of
  • Testimony, Epistemology of
  • Thomas Aquinas' Philosophy of Religion
  • Thought Experiments
  • Time and Tense
  • Time Travel
  • Transcendental Arguments
  • Truth and the Aim of Belief
  • Truthmaking
  • Turing Test
  • Two-Dimensional Semantics
  • Understanding
  • Uniqueness and Permissiveness in Epistemology
  • Utilitarianism
  • Value of Knowledge
  • Vienna Circle
  • Virtue Epistemology
  • Virtue Ethics
  • Virtues, Epistemic
  • Virtues, Intellectual
  • Voluntarism, Doxastic
  • Weakness of Will
  • Weil, Simone
  • William of Ockham
  • Williams, Bernard
  • Wittgenstein, Ludwig: Early Works
  • Wittgenstein, Ludwig: Later Works
  • Wittgenstein, Ludwig: Middle Works
  • Wollstonecraft, Mary
  • Privacy Policy
  • Cookie Policy
  • Legal Notice
  • Accessibility

Powered by:

  • [66.249.64.20|162.248.224.4]
  • 162.248.224.4
  • Search Menu
  • Browse content in Arts and Humanities
  • Browse content in Archaeology
  • Anglo-Saxon and Medieval Archaeology
  • Archaeological Methodology and Techniques
  • Archaeology by Region
  • Archaeology of Religion
  • Archaeology of Trade and Exchange
  • Biblical Archaeology
  • Contemporary and Public Archaeology
  • Environmental Archaeology
  • Historical Archaeology
  • History and Theory of Archaeology
  • Industrial Archaeology
  • Landscape Archaeology
  • Mortuary Archaeology
  • Prehistoric Archaeology
  • Underwater Archaeology
  • Urban Archaeology
  • Zooarchaeology
  • Browse content in Architecture
  • Architectural Structure and Design
  • History of Architecture
  • Residential and Domestic Buildings
  • Theory of Architecture
  • Browse content in Art
  • Art Subjects and Themes
  • History of Art
  • Industrial and Commercial Art
  • Theory of Art
  • Biographical Studies
  • Byzantine Studies
  • Browse content in Classical Studies
  • Classical History
  • Classical Philosophy
  • Classical Mythology
  • Classical Literature
  • Classical Reception
  • Classical Art and Architecture
  • Classical Oratory and Rhetoric
  • Greek and Roman Papyrology
  • Greek and Roman Epigraphy
  • Greek and Roman Law
  • Greek and Roman Archaeology
  • Late Antiquity
  • Religion in the Ancient World
  • Digital Humanities
  • Browse content in History
  • Colonialism and Imperialism
  • Diplomatic History
  • Environmental History
  • Genealogy, Heraldry, Names, and Honours
  • Genocide and Ethnic Cleansing
  • Historical Geography
  • History by Period
  • History of Emotions
  • History of Agriculture
  • History of Education
  • History of Gender and Sexuality
  • Industrial History
  • Intellectual History
  • International History
  • Labour History
  • Legal and Constitutional History
  • Local and Family History
  • Maritime History
  • Military History
  • National Liberation and Post-Colonialism
  • Oral History
  • Political History
  • Public History
  • Regional and National History
  • Revolutions and Rebellions
  • Slavery and Abolition of Slavery
  • Social and Cultural History
  • Theory, Methods, and Historiography
  • Urban History
  • World History
  • Browse content in Language Teaching and Learning
  • Language Learning (Specific Skills)
  • Language Teaching Theory and Methods
  • Browse content in Linguistics
  • Applied Linguistics
  • Cognitive Linguistics
  • Computational Linguistics
  • Forensic Linguistics
  • Grammar, Syntax and Morphology
  • Historical and Diachronic Linguistics
  • History of English
  • Language Evolution
  • Language Reference
  • Language Acquisition
  • Language Variation
  • Language Families
  • Lexicography
  • Linguistic Anthropology
  • Linguistic Theories
  • Linguistic Typology
  • Phonetics and Phonology
  • Psycholinguistics
  • Sociolinguistics
  • Translation and Interpretation
  • Writing Systems
  • Browse content in Literature
  • Bibliography
  • Children's Literature Studies
  • Literary Studies (Romanticism)
  • Literary Studies (American)
  • Literary Studies (Asian)
  • Literary Studies (European)
  • Literary Studies (Eco-criticism)
  • Literary Studies (Modernism)
  • Literary Studies - World
  • Literary Studies (1500 to 1800)
  • Literary Studies (19th Century)
  • Literary Studies (20th Century onwards)
  • Literary Studies (African American Literature)
  • Literary Studies (British and Irish)
  • Literary Studies (Early and Medieval)
  • Literary Studies (Fiction, Novelists, and Prose Writers)
  • Literary Studies (Gender Studies)
  • Literary Studies (Graphic Novels)
  • Literary Studies (History of the Book)
  • Literary Studies (Plays and Playwrights)
  • Literary Studies (Poetry and Poets)
  • Literary Studies (Postcolonial Literature)
  • Literary Studies (Queer Studies)
  • Literary Studies (Science Fiction)
  • Literary Studies (Travel Literature)
  • Literary Studies (War Literature)
  • Literary Studies (Women's Writing)
  • Literary Theory and Cultural Studies
  • Mythology and Folklore
  • Shakespeare Studies and Criticism
  • Browse content in Media Studies
  • Browse content in Music
  • Applied Music
  • Dance and Music
  • Ethics in Music
  • Ethnomusicology
  • Gender and Sexuality in Music
  • Medicine and Music
  • Music Cultures
  • Music and Media
  • Music and Religion
  • Music and Culture
  • Music Education and Pedagogy
  • Music Theory and Analysis
  • Musical Scores, Lyrics, and Libretti
  • Musical Structures, Styles, and Techniques
  • Musicology and Music History
  • Performance Practice and Studies
  • Race and Ethnicity in Music
  • Sound Studies
  • Browse content in Performing Arts
  • Browse content in Philosophy
  • Aesthetics and Philosophy of Art
  • Epistemology
  • Feminist Philosophy
  • History of Western Philosophy
  • Metaphysics
  • Moral Philosophy
  • Non-Western Philosophy
  • Philosophy of Language
  • Philosophy of Mind
  • Philosophy of Perception
  • Philosophy of Science
  • Philosophy of Action
  • Philosophy of Law
  • Philosophy of Religion
  • Philosophy of Mathematics and Logic
  • Practical Ethics
  • Social and Political Philosophy
  • Browse content in Religion
  • Biblical Studies
  • Christianity
  • East Asian Religions
  • History of Religion
  • Judaism and Jewish Studies
  • Qumran Studies
  • Religion and Education
  • Religion and Health
  • Religion and Politics
  • Religion and Science
  • Religion and Law
  • Religion and Art, Literature, and Music
  • Religious Studies
  • Browse content in Society and Culture
  • Cookery, Food, and Drink
  • Cultural Studies
  • Customs and Traditions
  • Ethical Issues and Debates
  • Hobbies, Games, Arts and Crafts
  • Lifestyle, Home, and Garden
  • Natural world, Country Life, and Pets
  • Popular Beliefs and Controversial Knowledge
  • Sports and Outdoor Recreation
  • Technology and Society
  • Travel and Holiday
  • Visual Culture
  • Browse content in Law
  • Arbitration
  • Browse content in Company and Commercial Law
  • Commercial Law
  • Company Law
  • Browse content in Comparative Law
  • Systems of Law
  • Competition Law
  • Browse content in Constitutional and Administrative Law
  • Government Powers
  • Judicial Review
  • Local Government Law
  • Military and Defence Law
  • Parliamentary and Legislative Practice
  • Construction Law
  • Contract Law
  • Browse content in Criminal Law
  • Criminal Procedure
  • Criminal Evidence Law
  • Sentencing and Punishment
  • Employment and Labour Law
  • Environment and Energy Law
  • Browse content in Financial Law
  • Banking Law
  • Insolvency Law
  • History of Law
  • Human Rights and Immigration
  • Intellectual Property Law
  • Browse content in International Law
  • Private International Law and Conflict of Laws
  • Public International Law
  • IT and Communications Law
  • Jurisprudence and Philosophy of Law
  • Law and Politics
  • Law and Society
  • Browse content in Legal System and Practice
  • Courts and Procedure
  • Legal Skills and Practice
  • Primary Sources of Law
  • Regulation of Legal Profession
  • Medical and Healthcare Law
  • Browse content in Policing
  • Criminal Investigation and Detection
  • Police and Security Services
  • Police Procedure and Law
  • Police Regional Planning
  • Browse content in Property Law
  • Personal Property Law
  • Study and Revision
  • Terrorism and National Security Law
  • Browse content in Trusts Law
  • Wills and Probate or Succession
  • Browse content in Medicine and Health
  • Browse content in Allied Health Professions
  • Arts Therapies
  • Clinical Science
  • Dietetics and Nutrition
  • Occupational Therapy
  • Operating Department Practice
  • Physiotherapy
  • Radiography
  • Speech and Language Therapy
  • Browse content in Anaesthetics
  • General Anaesthesia
  • Neuroanaesthesia
  • Clinical Neuroscience
  • Browse content in Clinical Medicine
  • Acute Medicine
  • Cardiovascular Medicine
  • Clinical Genetics
  • Clinical Pharmacology and Therapeutics
  • Dermatology
  • Endocrinology and Diabetes
  • Gastroenterology
  • Genito-urinary Medicine
  • Geriatric Medicine
  • Infectious Diseases
  • Medical Toxicology
  • Medical Oncology
  • Pain Medicine
  • Palliative Medicine
  • Rehabilitation Medicine
  • Respiratory Medicine and Pulmonology
  • Rheumatology
  • Sleep Medicine
  • Sports and Exercise Medicine
  • Community Medical Services
  • Critical Care
  • Emergency Medicine
  • Forensic Medicine
  • Haematology
  • History of Medicine
  • Browse content in Medical Skills
  • Clinical Skills
  • Communication Skills
  • Nursing Skills
  • Surgical Skills
  • Browse content in Medical Dentistry
  • Oral and Maxillofacial Surgery
  • Paediatric Dentistry
  • Restorative Dentistry and Orthodontics
  • Surgical Dentistry
  • Medical Ethics
  • Medical Statistics and Methodology
  • Browse content in Neurology
  • Clinical Neurophysiology
  • Neuropathology
  • Nursing Studies
  • Browse content in Obstetrics and Gynaecology
  • Gynaecology
  • Occupational Medicine
  • Ophthalmology
  • Otolaryngology (ENT)
  • Browse content in Paediatrics
  • Neonatology
  • Browse content in Pathology
  • Chemical Pathology
  • Clinical Cytogenetics and Molecular Genetics
  • Histopathology
  • Medical Microbiology and Virology
  • Patient Education and Information
  • Browse content in Pharmacology
  • Psychopharmacology
  • Browse content in Popular Health
  • Caring for Others
  • Complementary and Alternative Medicine
  • Self-help and Personal Development
  • Browse content in Preclinical Medicine
  • Cell Biology
  • Molecular Biology and Genetics
  • Reproduction, Growth and Development
  • Primary Care
  • Professional Development in Medicine
  • Browse content in Psychiatry
  • Addiction Medicine
  • Child and Adolescent Psychiatry
  • Forensic Psychiatry
  • Learning Disabilities
  • Old Age Psychiatry
  • Psychotherapy
  • Browse content in Public Health and Epidemiology
  • Epidemiology
  • Public Health
  • Browse content in Radiology
  • Clinical Radiology
  • Interventional Radiology
  • Nuclear Medicine
  • Radiation Oncology
  • Reproductive Medicine
  • Browse content in Surgery
  • Cardiothoracic Surgery
  • Gastro-intestinal and Colorectal Surgery
  • General Surgery
  • Neurosurgery
  • Paediatric Surgery
  • Peri-operative Care
  • Plastic and Reconstructive Surgery
  • Surgical Oncology
  • Transplant Surgery
  • Trauma and Orthopaedic Surgery
  • Vascular Surgery
  • Browse content in Science and Mathematics
  • Browse content in Biological Sciences
  • Aquatic Biology
  • Biochemistry
  • Bioinformatics and Computational Biology
  • Developmental Biology
  • Ecology and Conservation
  • Evolutionary Biology
  • Genetics and Genomics
  • Microbiology
  • Molecular and Cell Biology
  • Natural History
  • Plant Sciences and Forestry
  • Research Methods in Life Sciences
  • Structural Biology
  • Systems Biology
  • Zoology and Animal Sciences
  • Browse content in Chemistry
  • Analytical Chemistry
  • Computational Chemistry
  • Crystallography
  • Environmental Chemistry
  • Industrial Chemistry
  • Inorganic Chemistry
  • Materials Chemistry
  • Medicinal Chemistry
  • Mineralogy and Gems
  • Organic Chemistry
  • Physical Chemistry
  • Polymer Chemistry
  • Study and Communication Skills in Chemistry
  • Theoretical Chemistry
  • Browse content in Computer Science
  • Artificial Intelligence
  • Computer Architecture and Logic Design
  • Game Studies
  • Human-Computer Interaction
  • Mathematical Theory of Computation
  • Programming Languages
  • Software Engineering
  • Systems Analysis and Design
  • Virtual Reality
  • Browse content in Computing
  • Business Applications
  • Computer Security
  • Computer Games
  • Computer Networking and Communications
  • Digital Lifestyle
  • Graphical and Digital Media Applications
  • Operating Systems
  • Browse content in Earth Sciences and Geography
  • Atmospheric Sciences
  • Environmental Geography
  • Geology and the Lithosphere
  • Maps and Map-making
  • Meteorology and Climatology
  • Oceanography and Hydrology
  • Palaeontology
  • Physical Geography and Topography
  • Regional Geography
  • Soil Science
  • Urban Geography
  • Browse content in Engineering and Technology
  • Agriculture and Farming
  • Biological Engineering
  • Civil Engineering, Surveying, and Building
  • Electronics and Communications Engineering
  • Energy Technology
  • Engineering (General)
  • Environmental Science, Engineering, and Technology
  • History of Engineering and Technology
  • Mechanical Engineering and Materials
  • Technology of Industrial Chemistry
  • Transport Technology and Trades
  • Browse content in Environmental Science
  • Applied Ecology (Environmental Science)
  • Conservation of the Environment (Environmental Science)
  • Environmental Sustainability
  • Environmentalist Thought and Ideology (Environmental Science)
  • Management of Land and Natural Resources (Environmental Science)
  • Natural Disasters (Environmental Science)
  • Nuclear Issues (Environmental Science)
  • Pollution and Threats to the Environment (Environmental Science)
  • Social Impact of Environmental Issues (Environmental Science)
  • History of Science and Technology
  • Browse content in Materials Science
  • Ceramics and Glasses
  • Composite Materials
  • Metals, Alloying, and Corrosion
  • Nanotechnology
  • Browse content in Mathematics
  • Applied Mathematics
  • Biomathematics and Statistics
  • History of Mathematics
  • Mathematical Education
  • Mathematical Finance
  • Mathematical Analysis
  • Numerical and Computational Mathematics
  • Probability and Statistics
  • Pure Mathematics
  • Browse content in Neuroscience
  • Cognition and Behavioural Neuroscience
  • Development of the Nervous System
  • Disorders of the Nervous System
  • History of Neuroscience
  • Invertebrate Neurobiology
  • Molecular and Cellular Systems
  • Neuroendocrinology and Autonomic Nervous System
  • Neuroscientific Techniques
  • Sensory and Motor Systems
  • Browse content in Physics
  • Astronomy and Astrophysics
  • Atomic, Molecular, and Optical Physics
  • Biological and Medical Physics
  • Classical Mechanics
  • Computational Physics
  • Condensed Matter Physics
  • Electromagnetism, Optics, and Acoustics
  • History of Physics
  • Mathematical and Statistical Physics
  • Measurement Science
  • Nuclear Physics
  • Particles and Fields
  • Plasma Physics
  • Quantum Physics
  • Relativity and Gravitation
  • Semiconductor and Mesoscopic Physics
  • Browse content in Psychology
  • Affective Sciences
  • Clinical Psychology
  • Cognitive Psychology
  • Cognitive Neuroscience
  • Criminal and Forensic Psychology
  • Developmental Psychology
  • Educational Psychology
  • Evolutionary Psychology
  • Health Psychology
  • History and Systems in Psychology
  • Music Psychology
  • Neuropsychology
  • Organizational Psychology
  • Psychological Assessment and Testing
  • Psychology of Human-Technology Interaction
  • Psychology Professional Development and Training
  • Research Methods in Psychology
  • Social Psychology
  • Browse content in Social Sciences
  • Browse content in Anthropology
  • Anthropology of Religion
  • Human Evolution
  • Medical Anthropology
  • Physical Anthropology
  • Regional Anthropology
  • Social and Cultural Anthropology
  • Theory and Practice of Anthropology
  • Browse content in Business and Management
  • Business Ethics
  • Business Strategy
  • Business History
  • Business and Technology
  • Business and Government
  • Business and the Environment
  • Comparative Management
  • Corporate Governance
  • Corporate Social Responsibility
  • Entrepreneurship
  • Health Management
  • Human Resource Management
  • Industrial and Employment Relations
  • Industry Studies
  • Information and Communication Technologies
  • International Business
  • Knowledge Management
  • Management and Management Techniques
  • Operations Management
  • Organizational Theory and Behaviour
  • Pensions and Pension Management
  • Public and Nonprofit Management
  • Strategic Management
  • Supply Chain Management
  • Browse content in Criminology and Criminal Justice
  • Criminal Justice
  • Criminology
  • Forms of Crime
  • International and Comparative Criminology
  • Youth Violence and Juvenile Justice
  • Development Studies
  • Browse content in Economics
  • Agricultural, Environmental, and Natural Resource Economics
  • Asian Economics
  • Behavioural Finance
  • Behavioural Economics and Neuroeconomics
  • Econometrics and Mathematical Economics
  • Economic History
  • Economic Systems
  • Economic Methodology
  • Economic Development and Growth
  • Financial Markets
  • Financial Institutions and Services
  • General Economics and Teaching
  • Health, Education, and Welfare
  • History of Economic Thought
  • International Economics
  • Labour and Demographic Economics
  • Law and Economics
  • Macroeconomics and Monetary Economics
  • Microeconomics
  • Public Economics
  • Urban, Rural, and Regional Economics
  • Welfare Economics
  • Browse content in Education
  • Adult Education and Continuous Learning
  • Care and Counselling of Students
  • Early Childhood and Elementary Education
  • Educational Equipment and Technology
  • Educational Strategies and Policy
  • Higher and Further Education
  • Organization and Management of Education
  • Philosophy and Theory of Education
  • Schools Studies
  • Secondary Education
  • Teaching of a Specific Subject
  • Teaching of Specific Groups and Special Educational Needs
  • Teaching Skills and Techniques
  • Browse content in Environment
  • Applied Ecology (Social Science)
  • Climate Change
  • Conservation of the Environment (Social Science)
  • Environmentalist Thought and Ideology (Social Science)
  • Natural Disasters (Environment)
  • Social Impact of Environmental Issues (Social Science)
  • Browse content in Human Geography
  • Cultural Geography
  • Economic Geography
  • Political Geography
  • Browse content in Interdisciplinary Studies
  • Communication Studies
  • Museums, Libraries, and Information Sciences
  • Browse content in Politics
  • African Politics
  • Asian Politics
  • Chinese Politics
  • Comparative Politics
  • Conflict Politics
  • Elections and Electoral Studies
  • Environmental Politics
  • European Union
  • Foreign Policy
  • Gender and Politics
  • Human Rights and Politics
  • Indian Politics
  • International Relations
  • International Organization (Politics)
  • International Political Economy
  • Irish Politics
  • Latin American Politics
  • Middle Eastern Politics
  • Political Behaviour
  • Political Economy
  • Political Institutions
  • Political Methodology
  • Political Communication
  • Political Philosophy
  • Political Sociology
  • Political Theory
  • Politics and Law
  • Public Policy
  • Public Administration
  • Quantitative Political Methodology
  • Regional Political Studies
  • Russian Politics
  • Security Studies
  • State and Local Government
  • UK Politics
  • US Politics
  • Browse content in Regional and Area Studies
  • African Studies
  • Asian Studies
  • East Asian Studies
  • Japanese Studies
  • Latin American Studies
  • Middle Eastern Studies
  • Native American Studies
  • Scottish Studies
  • Browse content in Research and Information
  • Research Methods
  • Browse content in Social Work
  • Addictions and Substance Misuse
  • Adoption and Fostering
  • Care of the Elderly
  • Child and Adolescent Social Work
  • Couple and Family Social Work
  • Developmental and Physical Disabilities Social Work
  • Direct Practice and Clinical Social Work
  • Emergency Services
  • Human Behaviour and the Social Environment
  • International and Global Issues in Social Work
  • Mental and Behavioural Health
  • Social Justice and Human Rights
  • Social Policy and Advocacy
  • Social Work and Crime and Justice
  • Social Work Macro Practice
  • Social Work Practice Settings
  • Social Work Research and Evidence-based Practice
  • Welfare and Benefit Systems
  • Browse content in Sociology
  • Childhood Studies
  • Community Development
  • Comparative and Historical Sociology
  • Economic Sociology
  • Gender and Sexuality
  • Gerontology and Ageing
  • Health, Illness, and Medicine
  • Marriage and the Family
  • Migration Studies
  • Occupations, Professions, and Work
  • Organizations
  • Population and Demography
  • Race and Ethnicity
  • Social Theory
  • Social Movements and Social Change
  • Social Research and Statistics
  • Social Stratification, Inequality, and Mobility
  • Sociology of Religion
  • Sociology of Education
  • Sport and Leisure
  • Urban and Rural Studies
  • Browse content in Warfare and Defence
  • Defence Strategy, Planning, and Research
  • Land Forces and Warfare
  • Military Administration
  • Military Life and Institutions
  • Naval Forces and Warfare
  • Other Warfare and Defence Issues
  • Peace Studies and Conflict Resolution
  • Weapons and Equipment

Varieties of Understanding: New Perspectives from Philosophy, Psychology, and Theology

  • < Previous chapter
  • Next chapter >

11 Mechanistic versus Functional Understanding

  • Published: September 2019
  • Cite Icon Cite
  • Permissions Icon Permissions

Many natural and artificial entities can be predicted and explained both mechanistically, in term of parts and proximate causal processes, as well as functionally, in terms of functions and goals. Do these distinct “stances” or “modes of construal” support fundamentally different kinds of understanding? Based on recent work in epistemology and philosophy of science, as well as empirical evidence from cognitive and developmental psychology, this chapter argues for the “weak differentiation thesis”: the claim that mechanistic and functional understanding are distinct in that they involve importantly different objects. The chapter also considers more tentative arguments for the “strong differentiation thesis”: the claim that mechanistic and functional understanding involve different epistemic relationships between mind and world.

Signed in as

Institutional accounts.

  • Google Scholar Indexing
  • GoogleCrawler [DO NOT DELETE]

Personal account

  • Sign in with email/username & password
  • Get email alerts
  • Save searches
  • Purchase content
  • Activate your purchase/trial code

Institutional access

  • Sign in with a library card Sign in with username/password Recommend to your librarian
  • Institutional account management
  • Get help with access

Access to content on Oxford Academic is often provided through institutional subscriptions and purchases. If you are a member of an institution with an active account, you may be able to access content in one of the following ways:

IP based access

Typically, access is provided across an institutional network to a range of IP addresses. This authentication occurs automatically, and it is not possible to sign out of an IP authenticated account.

Sign in through your institution

Choose this option to get remote access when outside your institution. Shibboleth/Open Athens technology is used to provide single sign-on between your institution’s website and Oxford Academic.

  • Click Sign in through your institution.
  • Select your institution from the list provided, which will take you to your institution's website to sign in.
  • When on the institution site, please use the credentials provided by your institution. Do not use an Oxford Academic personal account.
  • Following successful sign in, you will be returned to Oxford Academic.

If your institution is not listed or you cannot sign in to your institution’s website, please contact your librarian or administrator.

Sign in with a library card

Enter your library card number to sign in. If you cannot sign in, please contact your librarian.

Society Members

Society member access to a journal is achieved in one of the following ways:

Sign in through society site

Many societies offer single sign-on between the society website and Oxford Academic. If you see ‘Sign in through society site’ in the sign in pane within a journal:

  • Click Sign in through society site.
  • When on the society site, please use the credentials provided by that society. Do not use an Oxford Academic personal account.

If you do not have a society account or have forgotten your username or password, please contact your society.

Sign in using a personal account

Some societies use Oxford Academic personal accounts to provide access to their members. See below.

A personal account can be used to get email alerts, save searches, purchase content, and activate subscriptions.

Some societies use Oxford Academic personal accounts to provide access to their members.

Viewing your signed in accounts

Click the account icon in the top right to:

  • View your signed in personal account and access account management features.
  • View the institutional accounts that are providing access.

Signed in but can't access content

Oxford Academic is home to a wide variety of products. The institutional subscription may not cover the content that you are trying to access. If you believe you should have access to that content, please contact your librarian.

For librarians and administrators, your personal account also provides access to institutional account management. Here you will find options to view and activate subscriptions, manage institutional settings and access options, access usage statistics, and more.

Our books are available by subscription or purchase to libraries and institutions.

  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Rights and permissions
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

Loading metrics

Open Access

Perspective

The Perspective section provides experts with a forum to comment on topical or controversial issues of broad interest.

See all article types »

The principle of uncertainty in biology: Will machine learning/artificial intelligence lead to the end of mechanistic studies?

* E-mail: [email protected]

Affiliation Systems Biology Department, Centro Nacional de Biotecnología, CSIC, C/ Darwin, Madrid, Spain

ORCID logo

  • Victor de Lorenzo

PLOS

Published: February 8, 2024

  • https://doi.org/10.1371/journal.pbio.3002495
  • Reader Comments

Molecular Biology has long tried to discover mechanisms, considering that unless we understand the principles, we cannot develop applications. Now machine learning and artificial intelligence enable direct leaps to application without understanding the principles. Will this herald a decline in mechanistic studies?

Citation: de Lorenzo V (2024) The principle of uncertainty in biology: Will machine learning/artificial intelligence lead to the end of mechanistic studies? PLoS Biol 22(2): e3002495. https://doi.org/10.1371/journal.pbio.3002495

Copyright: © 2024 Victor de Lorenzo. This is an open access article distributed under the terms of the Creative Commons Attribution License , which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Funding: This work was supported by European Union Horizon Programs (MIX-UP H2020-BIO-CN-2019-870294 and HORIZON-CL6-2021-UE 101060625 to VdL) and by European Structural and Investment Funds (BIOSINT-CM (Y2020/TCS- 6555-CAM to VdL). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Competing interests: The author has declared that no competing interests exist.

Biology as a scientific and research domain seems to have undergone major breakthroughs and paradigm shifts every time external disciplines have intersected with it. Typically, after a given time of somewhat uneasy coexistence, the community embraces the new conceptual frame and its associated technologies as a lens through which biological phenomena can be (re)interpreted. The happy encounter between biology and chemistry gave birth to biochemistry, enzymology, and metabolism. Much later, the interest of post-war physicists for live systems brought about the onset of molecular biology, which reached its biggest milestones in the elucidation of the DNA double helix and the deciphering of the genetic code. During the many decades dominated by molecular biology and molecular genetics, the emphasis has been on mechanistic understanding of biological phenomena, enabled by rigorous hypothesis-driven approaches imported from physics and formal mathematical logic. These departed from mere trial-and-error approaches that prevailed in previous stages and have enabled all-rational understanding of many key biological processes on the same principles that govern the rest of the material world. This is, after all, the ultimate mission of science as a human endeavour: rational understanding of reality with universal laws and principles.

Yet, the notion that by knowing the functioning of specific biological components we can understand the functioning of whole live systems became insufficient in view of the avalanche of data later generated by the plethora of “omics” technologies. Molecular biology reduces the complexity of a given phenomenon to a point where rigorous logic can be applied, experimental results unambiguously interpreted, and conclusions fixed as permanent pieces of knowledge. Yet, for this to happen, there should be a limited number of actors in an experiment. It thus follows that molecular biology is not capable of handling systems with too many components. Another conceptual and technical framework was clearly required. And thus systems biology [ 1 ], which largely relies on network theory (ultimately a branch of physics [ 2 ]), came about with the motto that “for understanding the whole, one has to study the whole”.

Using a systems biology approach, motifs, patterns, and correlations can be identified that in turn raise questions and testable hypothesis. However, the scientific agenda up to this point remained mechanistic comprehension of biological phenomena from first principles. The merging of molecular biology with systems biology has produced, for example, a comprehensive theory of gene expression regulation that captures the response mechanisms of live systems to changing scenarios in space and time in a simplified manner. This understanding has enabled the development of many heterologous expression systems and transcriptional circuits that can be parametrised and connected among them, and their behaviour predicted with high accuracy. Adding complexity, regulatory devices can be abstracted as connectable Boolean gates, thereby paving the way to biological computation based on logic circuits, not unlike their silicon-based counterparts [ 3 ]. Alas, the underlying assumption that biological systems can be understood as finite state machines (as typical computers are) breaks down when we face biological questions involving a very high number of variables that themselves vary on the fly and thus scape rigorous computation or simple relational logic [ 4 ].

There are at least 3 intertwined qualities of life that can make full comprehension of living systems ultimately unreachable. First, they grow, mutate, and evolve. This means that one live object we inspect at a given moment in space and time will not be identical to the same a moment later, let alone if we perturb it to sample or measure specific properties. Second, parameters associated to specific biological devices are context-dependent. One promoter in a location of the genome will show different input–output transfer functions when placed on another site or when cells are grown under different nutritional or environmental conditions. Since the number of such possible conditions is virtually infinite, so is the variability of the parameters at stake. Finally, biological matter is often soft, i.e., not hard in a physical sense. Instead, biology is about flexible materials, plastic shapes, glues, etc. Such components also change their form, which adds another degree of difficulty to making predictions with an accuracy remotely comparable to other branches of science and technology. This does not mean that Biology escapes the laws of physics, but that its ever-growing complexity makes its complete comprehension with standard approaches unreachable—even theoretically. This is the point where machine learning and artificial intelligence (ML/AI) come to the rescue.

ML is about adopting learning algorithms whose objective is to obtain a result (e.g., patterns, rules, correlations) dependent on the input variables (data) but without assuming any prefixed interplay among them. Through the use of computational and statistical methods, such algorithms are trained to make classifications and scores with an explicit degree of reliability. On this basis, it cannot come as a surprise that the large volume of data generated by omics technologies become an excellent input for training ML platforms. In turn, they can deliver dependable predictions on specific, yet complex, biological questions that are not yet amenable to mechanistic understanding.

One revealing case is the prediction of protein folding, a long-standing issue in biology that is now solved to a large extent by the ML/AI-run AlphaFold (AF) [ 5 ] and other AI-driven platforms. AF predictions, accurate as they may be, are not built on first principles but largely out of experience and statistical correlations among diverse data on the matter. AF is, in turn, the basis of a whole collection of remarkable platforms for protein engineering, including altogether new-to-nature structures and activities (see, e.g., https://loschmidt.chemi.muni.cz/portal/ ). To an extent, ML/AI is a return to the experience-based, trial-and-error, black-box approach, which was the source of knowledge in the prescientific era: We know how things behave and how to make them work, but we do not know ultimately why. But … does it matter? ( Box 1 ).

Box 1. A biological principle of uncertainty?

Mechanistic understanding from first principles anticipates solar eclipses and the successful test of the first prototype of an atomic bomb with 100% accuracy. This is hardly attainable in Biology if the phenomenon at stake has many components. On the other hand, ML-based Alpha Fold enables prediction of complex protein structures with an average accuracy of >80%. True, it is not 100% but it is good enough for moving ahead in most cases. If we accept that we cannot follow all variables of a system—in particular if they change through time and space—resorting to patterns and probabilities might enable us to nevertheless solve practical problems without knowing the underlying basis.

This raises a deeper question about how we study the composition and evolution of biological systems. As physicists gave up on describing the exact position of particles beyond a certain subatomic scale, biologists may also have to give up mechanistic understanding of live systems beyond a given level of complexity in favour of a probabilistic, experience-based approach. In specific scales, mechanistic comprehension—and thus complete predictability—could be inherently impossible. Thus, efforts in that direction are likely to be ultimately futile. Yet, we can still handle description and even understanding of biological systems in terms of chances, scores, and so on.

This should not lead us to disappointment, but rather should be celebrated, as it may open new conceptual frameworks for the understanding of biological phenomena as open-ended occurrences [ 6 ] that evolve through trial-and-error progressions, in a way not dissimilar to the processes that ML/AI tries to somehow capture.

Things in biology can become way more complicated than protein folding. Let us take a quite extreme example: the gut microbiome. We know it is an essential partner of the human body that interplays with the immune system, determining health and even psychological well-being through the gut–brain axis. The microbiome encompasses a hypercomplex association of microorganisms of all types undergoing continuous genetic exchange as well as massive molecular trade with the human host. Numerous microbiomes of individuals of diverse age, lifestyles, diets, locations, etc. have been characterised in detail in the last decades, associating distinct patterns of species to either healthy conditions or disease [ 7 ] (what has been called dysbiosis). But the enormity of the factors at play makes mechanistic studies virtually impossible. Is that a problem? It could well happen that AI-enabled developments will allow, e.g., for translational development of microbiome science without a deep understanding of how or why. Same in other areas of biology or biotechnology where the need of solutions is more pressing than the importance to understand their basis.

The emphasis on pursuing the laws for understanding the material world could be seen as one of the noblest human undertakings, but it has always been accompanied by the expectation of enabling accurate predictions for practical purposes. The more predictive a branch of science is, the more credible and appealing it is. Biology can make good predictions in systems or subsystems with a limited level of complexity, beyond which it relies on experience-based probabilities. Whether the results of ML/AI are just information or authentic scientific knowledge, as Sydney Brenner would put it [ 8 ], remains a legitimate question. A second question is whether there will be a decreasing interest in mechanistic understanding of biological phenomena in favour of using AI to obtain useful rules to make things happen in living systems with a high probability, even if we do not know why they do. Remarkably, even synthetic biology, which has the explicit aim of rational engineering of biology, increasingly relies on ML for adjusting parameters [ 9 ], dealing with context sensitivity [ 10 ], and guiding the artificial assembly of complex pathways [ 11 ]. We should celebrate this latest intersection of biology with a somewhat alien discipline—surely to be followed by other creative encounters in the future.

Acknowledgments

VdL is indebted to various participants of a Twitter/X discussion on the subject of this Perspective during 7–14 October 2023 ( https://shorturl.at/imnFM ).

  • 1. Alon U. An introduction to systems biology: design principles of biological circuits. CRC Press; 2019.
  • View Article
  • PubMed/NCBI
  • Google Scholar

The concept of mechanism in biology

Affiliation.

  • 1 Konrad Lorenz Institute for Evolution and Cognition Research, Adolf Lorenz Gasse, 2, Altenberg A-3422, Austria. [email protected]
  • PMID: 22326084
  • DOI: 10.1016/j.shpsc.2011.05.014

The concept of mechanism in biology has three distinct meanings. It may refer to a philosophical thesis about the nature of life and biology ('mechanicism'), to the internal workings of a machine-like structure ('machine mechanism'), or to the causal explanation of a particular phenomenon ('causal mechanism'). In this paper I trace the conceptual evolution of 'mechanism' in the history of biology, and I examine how the three meanings of this term have come to be featured in the philosophy of biology, situating the new 'mechanismic program' in this context. I argue that the leading advocates of the mechanismic program (i.e., Craver, Darden, Bechtel, etc.) inadvertently conflate the different senses of 'mechanism'. Specifically, they all inappropriately endow causal mechanisms with the ontic status of machine mechanisms, and this invariably results in problematic accounts of the role played by mechanism-talk in scientific practice. I suggest that for effective analyses of the concept of mechanism, causal mechanisms need to be distinguished from machine mechanisms, and the new mechanismic program in the philosophy of biology needs to be demarcated from the traditional concerns of mechanistic biology.

Copyright © 2011 Elsevier Ltd. All rights reserved.

Publication types

  • Historical Article
  • Research Support, Non-U.S. Gov't
  • Biology / history*
  • Concept Formation*
  • History, 17th Century
  • History, 18th Century
  • History, 19th Century
  • History, 20th Century
  • Models, Biological*
  • Philosophy / history*
  • Science / history*

Internet Explorer is no longer supported by Microsoft. To browse the NIHR site please use a modern, secure browser like Google Chrome, Mozilla Firefox, or Microsoft Edge.

National Institute for Health and Care Research logo | Homepage

Mechanistic studies, explanation and examples

mechanistic hypothesis biology

Published: 06 June 2019

Version: 1.3 March 2024

MRC-NIHR funding collaboration

EME encourages hypothesis-testing mechanistic studies. These studies can explore the mechanisms of action of the intervention, the causes of differing responses, or promote an understanding of any potential adverse effects and how these could be reduced. They could also contribute to understanding of the disease process.

We will fund hypothesis driven studies that can be tested using sample analysis, images, or other data. Commonly these mechanistic studies use the difference between treatment groups in a randomised clinical trial to explore the mechanism of action of the treatment, but we welcome other designs involving interventions used with therapeutic intent in patients.

It is important that applicants have a clear hypothesis about the mechanisms under study, and can demonstrate how the proposed tests or data analysis will confirm or refute the hypothesis. The rationale for the hypothesis will need to be justified to persuade the Funding Committee that there is sufficient grounds to fund the research.

In addition, the application must evidence how:

  • The proposed study would add to the existing body of knowledge
  • The proposed costs of the research are reasonable and commensurate with the proposed work involved. It is expected that applications to the Mechanisms of Action of Health Intervention call would typically be valued up to £500,000   
  • The study design will answer the research question proposed
  • The proposed study is feasible and deliverable
  • The team has the necessary skill mix and experience and project management of the study is sound
  • The proposed mechanistic evaluation will not undermine the delivery of the underlying efficacy or effectiveness trial (if it is still ongoing)

In general, where the investigation primarily aims to identify mechanisms of pathophysiology or disease, or to demonstrate proof-of-concept evidence of the validity and importance of new discoveries or treatments this would fall under the remit of the Medical Research Council (MRC) schemes for funding. Where the investigation aims to identify mechanisms of an intervention for which there are existing proof of concept or efficacy data, this would fall under the remit of the EME scheme for funding.

We will not fund exploratory analyses of samples or data to seek out previously unknown analytes or unexpected associations with outcomes. Examples of exploratory analyses that EME would be unlikely to fund are “biomarker discovery”, “proteomics”, “peptidomics”, “genome-wide association studies”, and similar strategies. EME may fund collection, processing and storage of samples from patients for the duration of an RCT or other NIHR-funded study testing a therapeutic intervention in patients, for future hypothesis driven analyses where these samples represent a unique collection that would otherwise not be available. In keeping with principles of open access, we would expect samples to be made available for use for other investigators, and would expect to see an initial plan for sample sharing to be submitted alongside any proposal.

We do not fund any research involving animals or animal tissues.

How we support mechanistic studies

We support mechanistic studies in three ways:

  • We fund mechanistic studies integrated within a main efficacy study (see studies EME 08/43/03 ,  EME 13/50/17 and others as examples).
  • We also support mechanistic studies that use samples or data from eligible on-going or completed RCTs or other studies that are funded by the NIHR testing a therapeutic intervention in patients (see study  EME 14/205/01 below for an example)
  • In addition, the remit has expanded to also include mechanistic studies using patients and samples collected from non-NIHR funded projects, for example, from MRC, Wellcome Trust, equivalent funding from the Devolved Administrations, and major charities that include external review processes if recruitment has been completed.

Applications for stand-alone mechanistic studies using samples or data are submitted in a two-stage application process. Please see the relevant commissioning brief for further information.

We will potentially fund mechanistic studies regardless of whether the result from the completed clinical study has shown benefit, no difference or harm.

Examples of EME funded studies with mechanistic components

Studies where determinants of the treatment response are investigated.

An investigator studied the clinical effectiveness of intermittent parent-determined montelukast as a treatment in pre-school wheezing children. The amount of enzyme that determines the production of the target molecule for montelukast varies with different genotypes of the promoter gene. The investigator designed a study that investigated the effect of intermittent parent-determined montelukast treatment on pre-school wheeze, whilst simultaneously determining if the treatment effect was modified by the child’s genotype. ( EME 08/43/03 )

A team of researchers are studying the efficacy of an IL-1 receptor antagonist in pustular psoriasis. It is possible that patients with pustular psoriasis have a deficiency in interleukin-36 receptor antagonist (IL36RA) leading to unrestrained IL-36 signalling and increased IL-1 production. The investigators will investigate the abnormal IL-1 signalling caused by IL-36 and other gene mutations to determine if they predict response to IL-1 receptor blockade. ( EME 13/50/17 )

Studies where the mechanism by which a treatment works is investigated

A study is using standard clinical pain assessment tools to test whether morphine provides analgesia and increased physiological stability during retinopathy of prematurity screening in neonates. In addition to standard clinical pain assessment the investigators are using recently developed electrophysiological techniques to examine the effects of morphine on the underlying brain and spinal cord activity evoked by the painful procedures, to determine how nociceptive information is transmitted to and processed by the cortex and where morphine interacts with this pathway. ( EME 14/187/01 )

In a study of ulipristal acetate against a levonorgestrel-releasing intra-uterine system (Mirena) for the long-term treatment of heavy menstrual bleeding a sub-set of patients will undergo imaging of the uterus using high resolution MRI of the uterine matrix and fibroids and Dynamic Contrast Enhanced MRI to measure uterine perfusion. This will determine if ulipristal influences menstrual bleeding via a reduction in uterine blood flow. ( EME 12/206/52 )

Studies using data or samples from other NIHR-funded trials

Intravenous tranexamic acid may reduce the severity of traumatic brain injury by reducing the severity of intracranial bleeding. It is also possible that tranexamic acid has a direct anti-inflammatory action on the brain which may also be beneficial in patients with a brain injury. An investigator is using brain microdialysis techniques in a sub-set of patients recruited to the HTA-funded CRASH-3 study to determine whether markers of brain inflammation are changed by tranexamic acid and whether this is related to outcome. ( EME 14/205/01 )

An EME-funded study aims to understand the mechanisms of change in group-based weight loss and related behaviour change interventions. The investigators are using recordings of group sessions from two randomised controlled trials and a feasibility study, all NIHR funded. Qualitative research methods are being used to extract target features at group level and then determine if they are associated with successful weight loss. ( EME 14/202/03 )

Examples of mechanistic components that would be outside EME’s remit

An investigator proposes testing a new intervention to treat exacerbations of refractory asthma with a control group treated according to standard BTS guidelines. She proposes collecting urine specimens at randomisation on all patients. Using physicochemical protein capture and high-performance liquid chromatography with tandem mass spectrometry she proposes to characterise the urinary proteome and then look for as yet unknown differences between “responders” and “non-responders” to the new treatment. Any identified biomarkers might provide the basis for future stratified care.

Comment: there is no proposed mechanism or identified analyte. This is a “biomarker discovery” project. EME would consider funding the sample collection and storage but would not fund the analysis.

A researcher proposes a study of an on-line, self-administered cognitive behavioural therapy (CBT) programme against an attention control for patients with obsessional-compulsive disorder. He requests funding for functional MRI studies in all patients to confirm previous findings that, in patients with obsessional-compulsive disorder, there is an abnormality of fMRI-determined amygdala activation to human face images.

Comment: there is no hypothesis that relates to the intervention under study, the researcher is requesting funding to study the whole cohort to confirm previous studies in this disease.

A very large effectiveness study funded by the NIHR is looking at an agent that may reduce the number of patients who develop metastatic deposits of colonic adenocarcinoma after primary resection. The agent may work via a mechanism involving endothelial cell surface antigens. A researcher proposes an add-on mechanistic study. This involves measuring the expression of these markers in venous biopsies in both groups to see how the treatment changes the surface markers of interest, and parallel work involving in vitro studies on endothelial cell cultures from an immortalised cell line.

Comment: the venous biopsy component would be within EME’s remit as it examines one possible mechanism of action of the treatment. Parallel laboratory studies not involving patient material or data are out of remit.

An investigator proposed a study of a structured exercise regimen for rehabilitation of patients with lung cancer treated with pulmonary resection. The control group was “care as usual”. She proposed a mechanistic study that included measures of muscle strength, maximal oxygen consumption during exercise (VO2max) and ultrasound-determined quadriceps mass.

Comment: the mechanistic component in this study would simply confirm the intervention had the expected effect on measures of physical fitness and does not explore novel mechanisms.

Psychology Dictionary

MECHANISTIC THEORY

is one of many theories which attempt to explain human behaviours. Mechanistic theory implies that human behaviours can be explained in the exact same way that mechanical and physiological processes are explained and understood.

Mechanistic Theory

Introduction

The mechanistic theory is an important topic in the biological and social sciences since it allows us to understand complex systems. It adopts an original approach by viewing all aspects of the natural world and social processes as interrelated pieces of a vast machine. It may unearth the rich historical context in which mechanistic theory was established, investigate the main principles that define it, and expose its practical applicability in diverse disciplines of study by delving deeply into it.

Mechanistic theory is a point of view that looks both natural and social phenomena as systems made up of interconnected components that work together in a machine-like fashion. According to this theory, it may successfully predict and explain the behavior and results of the entire system by developing a comprehensive understanding of how different pieces interact and work together.

Background and Context

Historical perspective.

The mechanistic viewpoint can trace its origins back to the Scientific Revolution :

  • The mechanistic theory originated during the Scientific Revolution of the 16th and 17th centuries.
  • René Descartes and Isaac Newton played significant roles in developing the mechanistic perspective.
  • The theory was further refined during the industrial revolution when parallels were drawn between machines and natural systems.

Mechanistic Theory

Current State of Research

Modern breakthroughs in mechanistic theory have progressed beyond simple mechanical analogies, incorporating insights from domains as diverse as systems biology, cognitive science, and artificial intelligence research.:

  • Modern developments in mechanistic theory incorporate insights from various fields.
  • The current focus is on understanding complex, dynamic systems rather than purely deterministic models.

The mechanistic theory is particularly useful in scientific research, where it guides the formation of hypotheses and the design of experiments.:

  • Mechanistic theory provides a valuable framework for understanding complex systems.
  • It plays a critical role in scientific research, guiding hypothesis formulation and experimental design.

Key Concepts and Terminologies

Two essential terms in the mechanistic theory are :

  • System: In the mechanistic theory refers to an assembly of components interacting in a coordinated manner.
  • Mechanism : refers to the processes through which these components interact to produce system behavior.

Mechanistic Theory

Practical Applications and Implications

The mechanistic hypothesis underpins a wide range of disciplines, including biology and sociology. It aids in the understanding of processes at the cellular level in biology, and it aids in the explanation of social structures and events in sociology. The theory can also be used to develop artificial systems like computers and robotics by borrowing inspiration from natural mechanisms.

Mechanistic Theory

FAQ: Frequently Asked Questions

What is the mechanistic theory, and why is it important.

The mechanistic theory is a point of view that sees natural and social processes as systems made up of interconnected pieces, similar to a machine. The significance of this theory arises from its ability to provide a thorough understanding, prediction, and control over the behavior of complex systems.

How do scientists conduct research based on the mechanistic theory?

In their study, scientists frequently use a combination of observation, experimentation , and mathematical modeling. This hybrid approach assists them in identifying and comprehending the fundamental mechanisms that govern the behavior of distinct systems.

What are the different branches of the mechanistic theory?

The mechanistic theory is adaptable and finds applications in a variety of domains, each with its own adaptations . Mechanistic biology investigates biological systems, mechanistic psychology investigates mental and behavioral processes, and mechanistic sociology investigates social structures and events.

How do concepts from the mechanistic theory apply to daily life?

The principles of the mechanistic theory can be employed to understand a wide range of everyday phenomena. For instance, the functioning of a car engine, which involves various parts working together, or the dynamics of a sports team, where individual roles and interactions lead to collective performance, can be understood using the mechanistic theory.

How does the mechanistic theory influence the design of artificial systems?

 The mechanistic theory plays a significant role in the design of artificial systems li ke computers and robots. By understanding natural mechanisms, designers can mimic these processes to create efficient, sophisticated systems that accomplish specific tasks.

How has the mechanistic theory evolved over time?

The mechanistic theory has its roots in the 16th and 17th centuries' Scientific Revolution. It has evolved over time to incorporate findings from many domains like as systems biology, cognitive science, and artificial intelligence. The emphasis has also changed from solely deterministic models to comprehending complex, dynamic systems.

What limitations does the mechanistic theory have?

While the mechanistic paradigm is effective, it has limits. It has the potential to oversimplify complicated systems by disregarding characteristics such as randomness and unpredictability. Furthermore, not all systems, particularly those with non-linear or emergent features , may fit cleanly into a mechanistic framework.

How does the mechanistic theory relate to other scientific theories?

Other scientific theories are frequently supplemented by mechanistic theory. In biology, for example, it works alongside evolutionary theory to explain organismal complexity. It can be used in conjunction with theories of social behavior to explain social structures and occurrences in the social sciences.

 References

Bechtel, W. (2005).  Discovering Cell Mechanisms: The Creation of Modern Cell Biology  (Cambridge Studies in Philosophy and Biology). Cambridge: Cambridge University Press. doi:10.1017/CBO9781139164962

Bechtel, William, and Robert C. Richardson. “Front Matter.” Discovering Complexity: Decomposition and Localization as Strategies in Scientific Research , The MIT Press, 2010, pp. i–vi. JSTOR , http://www.jstor.org/stable/j.ctt5hhhpx.1. Accessed 16 May 2023.

Avatar photo

Leave a Reply

Your email address will not be published. Required fields are marked *

Latest Posts

mechanistic hypothesis biology

Counseling Approaches to Client Care: Theories to Apply in Practice

mechanistic hypothesis biology

The Future Of Education: Can You Earn A Psychology Degree Online?

mechanistic hypothesis biology

Insomnia & Mental Illness: What is the Correlation?

Psychology of Decision Making

Stop Guessing: Here Are 3 Steps to Data-Driven Psychological Decisions

mechanistic hypothesis biology

Getting Help with Grief: Understanding Therapy & How It Can Help

mechanistic hypothesis biology

Exploring the Psychology of Risk and Reward

mechanistic hypothesis biology

Understanding ADHD in Women: Symptoms, Treatment & Support

mechanistic hypothesis biology

Meeting the Milestones: A Guide to Piaget's Child Developmental Stages

mechanistic hypothesis biology

Counseling, Therapy, and Psychology: What Is The Difference?

mechanistic hypothesis biology

The Psychology of Metaphysical Belief Systems

mechanistic hypothesis biology

4 Key Considerations When Supporting a Loved One Through a Legal Battle for Justice 

How Exercise Can Boost Your Mental Health as You Age

Finding Balance: The Psychological Benefits of Staying Active

Popular psychology terms, medical model, hypermnesia, affirmation, backup reinforcer, brainwashing, message-learning approach, affiliative behavior, kinesthetic feedback, behavioral sequence, contrast effect.

  • BiologyDiscussion.com
  • Follow Us On:
  • Google Plus
  • Publish Now

Biology Discussion

The Origin of Life: Concept, Approaches and Theories

mechanistic hypothesis biology

ADVERTISEMENTS:

In this article we will discuss about:- 1. Concept of the Origin of Life 2. Approaches of the Origin of Life 3. Theories.

Concept of the Origin of Life:

The origin of life on earth is one of the most elusive and oldest problems of bio­logy. For about one half of the long span of earth’s history, the earth was barren and lifeless. But at some indeterminate time the entity called life came into exis­tence.

What form it took and what circum­stances brought it into being remain ever a problematic issue. The attention of the keenest minds of science and philosophy tried to answer this question. From time immemorial several explanations have been put forward to explain this phenomenon. But recent researches and experimenta­tions have abandoned most of the expla­nations.

Approaches of the Origin of Life :

By the turn of this century several main approaches have been made on the origin of life. These approaches are the vitalism, special creation, panspermia, mechanistic theory and materialism.

1. Vitalism:

This concept attributes the distinctive properties of the living things to a supernatural life force.

2. Special Creation:

This concept is nothing but the literal interpretation of the Biblical aspects of genesis.

3. Panspermia:

By assuming the eternal and universal existence of life, this concept avoids the question of the origin of life on this earth.

These three ideas regarding the origin of life, viz., vitalism, special creation and panspermia, need no detailed considera­tion, because they are based on assump­tion without having any scientific back­ground. But the other two theories, the mechanistic theory and the materialistic, doctrine need tremendous consideration.

Theories of the Origin of Life:

1. Mechanistic Theories:

These theories are based on the assumption that (i) the origin and activities of living things are derived from natural laws and (ii) that the living things are derived from inorga­nic matters in accordance with those natural laws. But these ideas were con­fronted with serious difficulties.

The real potent question is how can a living thing be formed from inorganic matters? These theories believed that the earth was total­ly devoid of organic compounds in pre-biological days arid that only living organisms can produce organic com­pounds.

The organic compounds cons­titute the most vital part of the living or­ganisms and as such their origin in the absence of organic compounds remains a formidable problem for ever.

The mechanistic theories resolved that the first living thing was a macromolecule or “living molecule” that was originated by the accidental coming together of the components that composed the living molecule in appropriate and proper arrangements and proportions.

It is also believed that the living molecule was probably a molecule of protein—the most important part of the living things. The mechanistic concepts advocate the origin of first living things by the chance combi­nation of these elements. This has come about by trail and error through an enor­mously long span of time.

2. Materialistic Theories:

These theories hold a different view by applying the natural laws to explain the origin of life and reject the contention of the mechanistic theories that only living things can synthesise organic compounds. On the contrary these theories propose that the organic compounds were formed abiogenecally, (i.e. without the intervention of living things) prior to the living organisms came into existence.

Instead of an acci­dental get-together of the elements to produce living things all at once as advo­cated in mechanistic doctrine, materialis­tic theories advocate the origin of life as result of series of steps towards increas­ing complexities and eventually leading up to the living state.

One of the most important steps is the abiogenic synthesis of the organic ‘compounds. Every step leads into a higher grade of organisation of matter that did not exist at the lower level. With the gradual formation of molecules, new physical as well as chemi­cal properties appeared in the scene which were non-existing in the realm of uncombined elements.

By this way and with the advent of living organisms, biological laws came into operation which did not exist before. The materialistic theories hold that the origin of life was not an accident, but it was the end result of matters evolving into higher and higher forms step by step.

Evolution of living things from non-living matters was the speculation of the thinkers of the past. To name a few, they were F. Engels, H. Spencer, T. H. Huxley, C. Mitchell and E. A. Schafer. F. Engels in his “Dialectics of Nature” condemned the concept of the spontaneous generation and vitalism and maintained that the living things have originated from a continuous evolution of matters.

A Russian Biochemist, A. I. Oparin in 1922, holds that the organic compounds had existed on the earth before the origin of living things and the living things have evolved from these pre-existing compounds.

He views the evolutionary sequences (such as the formation of inorganic compounds, organic compounds, primitive and higher living things) not as separate events, but as merely the steps in the process of un­folding of the same biological process.

J. B. S. Haldane in 1928 was also of the opinion of the abiogenic formation of organic substances in the prebiological history of the earth. He postulated that the organic compounds must have accu­mulated before the origin of life.

He assumed that the primitive atmos­phere contained only carbon dioxide, ammonia and water vapour. There was no oxygen. Haldane was of the opinion that such a combination, exposed to ultra­violet light, can evolve into a large variety of organic compounds.

Oparin’s Hypothesis:

In his attempt to solve the problem of the origin of organic compounds on this earth, Oparin based his concept on the then accepted theory on the origin of earth proposed by Sir James Jeans. Although Sir Jean’s theory of the fiery origin of the planets is no longer acceptable, it provided a basis for the formulation of Oparin’s theory of the origin of organic compounds on the earth.

Oparin assumes that the first carbon compounds were the hydrocarbons and not the carbon dioxide. To put the question in other way, he assumes the first appearance of carbon in the reduced state and not in the oxidised state because the pri­mordial atmosphere was devoid of oxygen.

With the cooling down of the atmosphere to 1000°C. and lower, the highly reactive free radicals (viz. CH and CH 2 ) combined to produce a variety of simple saturated and unsaturated hydrocarbons.

The crust of the earth had begun to form but the temperature was still too hot to allow condensation of steam to water. But the atmosphere above the crust contained carbon as well as nitrogen in reduced state. The reduced state of carbon was the hydrocarbons while that of nitrogen was ammonia. Such reducing atmosphere forms the basis of Oparin’s theory of the origin of life.

The simplest substances, such as hydro­gen, steam, methane, hydrogen sulfide, ammonia and nitrogen are continuously producing simple hydrocarbons and their derivatives—the alcohols, aldehydes, sugars, ketones and acids.

Of the deriva­tives there are the amino-acids and the polymerisation of the amino-acids results in the formation of protein. The mole­cules of protein have the capability of coming together and form large complexes. By this way colloidal hydrophilic com­plexes are produced and these complexes play the active role in the transformation of inanimate things into living systems.

Related Articles:

  • Origin of Life on Planet Earth!
  • Modern Concept on the Origin of Life (With Diagram)

Biology , Organic Evolution , Life , Origin , Origin of Life

  • Anybody can ask a question
  • Anybody can answer
  • The best answers are voted up and rise to the top

Forum Categories

  • Animal Kingdom
  • Biodiversity
  • Biological Classification
  • Biology An Introduction 11
  • Biology An Introduction
  • Biology in Human Welfare 175
  • Biomolecules
  • Biotechnology 43
  • Body Fluids and Circulation
  • Breathing and Exchange of Gases
  • Cell- Structure and Function
  • Chemical Coordination
  • Digestion and Absorption
  • Diversity in the Living World 125
  • Environmental Issues
  • Excretory System
  • Flowering Plants
  • Food Production
  • Genetics and Evolution 110
  • Human Health and Diseases
  • Human Physiology 242
  • Human Reproduction
  • Immune System
  • Living World
  • Locomotion and Movement
  • Microbes in Human Welfare
  • Mineral Nutrition
  • Molecualr Basis of Inheritance
  • Neural Coordination
  • Organisms and Population
  • Photosynthesis
  • Plant Growth and Development
  • Plant Kingdom
  • Plant Physiology 261
  • Principles and Processes
  • Principles of Inheritance and Variation
  • Reproduction 245
  • Reproduction in Animals
  • Reproduction in Flowering Plants
  • Reproduction in Organisms
  • Reproductive Health
  • Respiration
  • Structural Organisation in Animals
  • Transport in Plants
  • Trending 14

Privacy Overview

web counter

Ecological Rants

Ecological opinions of charley krebs and judy myers, tag archives: mechanistic hypotheses, have we moved on from hypotheses into the new age of ecology.

For the last 60 years a group of Stone Age scientists like myself have preached to ecology students that one needs hypotheses to do proper science. Now it has always been clear that not all ecologists followed this precept, and a recent review hammers this point home (Betts et al. 2021). I have always asked my students to read the papers from the Stone Age about scientific progress – Popper (1959), Platt (1964), Peters (1991) and even back to the Pre-Stone Age, Chamberlin (1897). There has been much said about this issue, and the recent Betts et al. (2021) paper pulls much of it together by reviewing papers from 1991 to 2015. Their conclusion is dismal if you think ecological science should make progress in gathering evidence. No change from 1990 to 2015. Multiple alternative hypotheses = 6% of papers, Mechanistic hypotheses = 25% of papers, Descriptive hypotheses = 12%, No hypotheses = 75% of papers. Why should this be after years of recommending the gold standard of multiple alternative hypotheses? Can we call ecology a science with these kinds of scores? 

The simplest reason is that in the era of Big Data we do not need any hypotheses to understand populations, communities, and ecosystems. We have computers, that is enough. I think this is a rather silly view, but one would have to interview believers to find out what they view as progress from big data in the absence of hypotheses. The second excuse might be that we cannot be bothered with hypotheses until we have a complete description of life on earth, food webs, interaction webs, diets, competitors, etc. Once we achieve that we will be able to put together mechanistic hypotheses rapidly. An alternative statement of this view is that we need very much natural history to make any progress in ecology, and this is the era of descriptive natural history and that is why 75% of papers do not list the word hypothesis .

But this is all nonsense of course, and try this view on a medical scientist, a physicist, an aeronautical engineer, or a farmer. The fundamental principle of science is cause-and-effect or the simple view that we would like to see how things work and why often they do not work. Have your students read Romesburg (1981) for an easy introduction and then the much more analytical book by Pearl and Mackenzie (2018) to gain an understanding of the complexity of the simple view that there is a cause and it produces an effect . Hone et al. (2023) discuss these specific problems with respect to improving our approach to wildlife management

What can be done about the dismal situation described by Betts et al. (2021)? One useful recommendation for editors and reviewers would be to request for every submitted paper for a clear statement of the hypothesis they are testing, and hopefully for alternative hypotheses. There should be ecology journals specifically for natural history where the opposite gateway is set: no use of ‘ hypothesis ’ in this journal. This would not solve all the Betts et al. problems because some ecology papers are based on the experimental design of ‘do something’ and then later ‘try to invent some way to support a hypotheses’, after the fact science. One problem with this type of literature survey is, as Betts et al. recognized, is that papers could be testing hypotheses without using this exact word. So words like ‘proposition’, ‘thesis’, ‘conjectures’ could camouflage thinking about alternative explanations without the actual word ‘hypothesis’.

One other suggestion to deal with this situation might be for journal editors to disallow all papers with hypotheses that are completely untestable. This type of rejection could be instructive to authors to assist rewriting your paper to be more specific about alternative hypotheses. If you can make a clear causal set of predictions that a particular species will go extinct in 100 years, this could be described as a ‘possible future scenario’ that could be guided by some mechanisms that are specified. Or if you have a hypothesis that ‘climate change will affect species geographical ranges, you are providing  a very vague inference that is difficult to test without being more specific about mechanisms, particularly if the species involved is rare.

There is a general problem with null hypotheses which state there is “no effect”. In some few cases these null hypotheses are useful but for the most part they are very weak and should indicate that you have not thought enough about alternative hypotheses.

So read Platt (1964) or at least the first page of it, the first chapter of Popper (1959), and Betts et al. (2021) paper and in your research try to avoid the dilemmas they discuss, and thus help to move our science forward lest it become a repository of ‘stamp collecting’.

Betts, M.G., Hadley, A.S., Frey, D.W., Frey, S.J.K., Gannon, D., Harris, S.H., et al. (2021) When are hypotheses useful in ecology and evolution? Ecology and Evolution, 11, 5762-5776. doi: 10.1002/ece3.7365.

Chamberlin, T.C. (1897) The method of multiple working hypotheses. Journal of Geology, 5, 837-848 (reprinted in Science 148: 754-759 in 1965). doi. 10.1126/science.148.3671.754.

Hone, J., Drake, A. & Krebs, C.J. (2023) Evaluation options for wildlife management and strengthening of causal inference BioScience, 73, 48-58.doi: 10.1093/biosci/biac105.

Pearl, J., and Mackenzie, D. 2018. The Book of Why. The New Science of Cause and Effect . Penguin, London, U.K. 432 pp. ISBN: 978-1541698963.

Peters, R.H. (1991) A Critique for Ecology . Cambridge University Press, Cambridge, England. ISBN: 0521400171.

Platt, J.R. (1964) Strong inference. Science , 146, 347-353.doi: 10.1126/science.146.3642.347.

Popper, K.R. (1959) The Logic of Scientific Discovery . Hutchinson & Co., London. ISBN: 978-041-5278-447.

Romesburg, H.C. (1981) Wildlife science: gaining reliable knowledge. Journal of Wildlife Management, 45, 293-313. doi:10.2307/3807913.

On Assumptions in Ecology Papers

What can we do as ecologists to improve the publishing standards of ecology papers? I suggest one simple but bold request. We should require at the end of every published paper a annotated list of the assumptions made in providing the analysis reported in the paper. A tabular format could be devised with columns for the assumption, the perceived support of and tests for the assumption, and references for this support or lack thereof. I can hear the screaming already, so this table could be put in the Supplementary Material which most people do not read. We could add to each paper in the final material where there are statements of who did the writing, who provided the money, and add a reference to this assumptions table in the Supplementary Material or a statement that no assumptions about anything were made to reach these conclusions.

The first response I can detect to this recommendation is that many ecologists will differ in what they state are assumptions to their analysis and conclusions. As an example, in wildlife studies, we commonly make the assumption that an individual animal having a radio collar will behave and survive just like another animal with no collar. In analyses of avian population dynamics, we might commonly assume that our visiting nests does not affect their survival probability. We make many such assumptions about random or non-random sampling. My question then is whether or not there is any value in listing these kinds of assumptions. My response is that this approach of listing what the authors think they are assuming should alert the reviewers to the elephants in the room that have not been listed.

My attention was called to this general issue by the recent paper of Ginzburg and Damuth (2022) in which they contrasted the assumptions of two general theories of functional responses of predators to prey – “prey dependence” versus “ratio dependence”. We have in ecology many such either-or discussions that never seem to end. Consider the long-standing discussion of whether populations can be regulated by factors that are “density dependent” or “density independent”, a much-debated issue that is still with us even though it was incisively analyzed many years ago.  

Experimental ecology is not exempt from assumptions, as outlined in Kimmel et al. (2021) who provide an incisive review of cause and effect in ecological experiments. Pringle and Hutchinson (2020) discuss the failure of assumptions in food web analysis and how these might be resolved with new techniques of analysis. Drake et al. (2021) consider the role of connectivity in arriving at conservation evaluations of patch dynamics, and the importance of demographic contributions to connectivity via dispersal. The key point is that, as ecology progresses, the role of assumptions must be continually questioned in relation to our conclusions about population and community dynamics in relation to conservation and landscape management.

Long ago Peters (1991) wrote an extended critique of how ecology should operate to avoid some of these issues, but his 1991 book is not easily available to students (currently available on Amazon for about $90). To encourage more discussion of these questions from the older to the more current literature, I have copied Peters Chapter 4 to the bottom of my web page at https://www.zoology.ubc.ca/~krebs/books.html for students to download if they wish to discuss these issues in more detail.

Perhaps a possible message in all this has been that ecology has always wished to be “physics-in-miniature” with grand generalizations like the laws we teach in the physical sciences. Over the last 60 years the battle in the ecology literature has been between this model of physics and the view that every population and community differ, and everything is continuing to change under the climate emergency so that we can have little general theory in ecology. There are certainly many current generalizations, but they are relatively useless for a transition from the general to the particular for the development of a predictive science. The consequence is that we now bounce from individual study to individual study, typically starting from different assumptions, with very limited predictability that is empirically testable. And the central issue for ecological science is how can we move from the present fragmentation in our knowledge to a more unified science. Perhaps starting to examine the assumptions of our current publications would be a start in this direction.  

Drake, J., Lambin, X., and Sutherland, C. (2021). The value of considering demographic contributions to connectivity: a review. Ecography 44 , 1-18. doi: 10.1111/ecog.05552.

Ginzburg, L.R. and Damuth, J. (2022). The Issue Isn’t Which Model of Consumer Interference Is Right, but Which One Is Least Wrong. Frontiers in Ecology and Evolution 10 , 860542. doi: 10.3389/fevo.2022.860542.

Kimmel, K., Dee, L.E., Avolio, M.L., and Ferraro, P.J. (2021). Causal assumptions and causal inference in ecological experiments. Trends in Ecology & Evolution 36 , 1141-1152. doi: 10.1016/j.tree.2021.08.008.

Peters, R.H. (1991) ‘ A Critique for Ecology .’ (Cambridge University Press: Cambridge, England.) ISBN:0521400171 (Chapter 4 pdf available at https://www.zoology.ubc.ca/~krebs/books.html )

Pringle, R.M. and Hutchinson, M.C. (2020). Resolving Food-Web Structure. Annual Review of Ecology, Evolution, and Systematics 51 , 55-80. doi: 10.1146/annurev-ecolsys-110218-024908.

On Evolution and Ecology and Climate Change

If ecology can team up with evolution to become a predictive science, we can all profit greatly since it will make us more like physics and the hard sciences. It is highly desirable to have a grand vision of accomplishing this, but there could be a few roadblocks on the way. A recent paper by Bay et al. (2018) illustrates some of the difficulties we face.

The yellow warbler ( Setophaga petechia ) has a broad breeding range across the United States and Canada, and could therefore be a good species to survey because it inhabits widely different climatic zones. Bay et al. (2018) identified genomic variation associated with climate across the breeding range of this migratory songbird, and concluded that populations requiring the greatest shifts in allele frequencies to keep pace with future climate change have experienced the largest population declines, suggesting that failure to adapt may have already negatively affected population abundance. This study by Bay et al. (2018) sampled 229 yellow warblers from 21 locations across North America, with an average of 10 birds per sample area (range n = 6 to 21). They examined 104,711 single-nucleotide polymorphisms. They correlated genetic structure to 19 climate variables and 3 vegetation indices, a measure of surface moisture, and average elevation. This is an important study claiming to support an important conclusion, and consequently it is also important to break it down into the three major assumptions on which it rests.

First, this study is a space for time analysis, a subject of much discussion already in plant ecology (e.g. Pickett 1989, Blois et al. 2013). It is an untested assumption that you can substitute space for time in analyzing for future evolutionary changes.

Second, the conclusions of the Bay et al. paper rest on an assumption that you have adequate data on the genetics involved in change and on the demography of the species. A clear understanding of the ecology of the species and what limits its distribution and abundance would seem to be prerequisites for understanding the mechanisms of how evolutionary changes might occur.

The third assumption is that, if there is a correlation between the genetic measures and the climate or vegetation indices, one can identify the precise ‘genomic vulnerability’ of the local population. Genomic variation was most closely related to precipitation variables at each site. The geographic area with one of the highest scores in genomic vulnerability was in the desert area of the intermountain west (USA). As far as I can determine from their Figure 1, there was only one sampling site in this whole area of the intermountain west. Finally Bay et al. (2018) compared the genomic vulnerability data to the population changes reported for each site. Population changes for each sampled site were obtained from the North American Breeding Bird Survey data from 1996 to 2012.

The genetic data and its analysis are more impressive, and since I am not a genetics expert I will simply give it a A grade for genetics. It is the ecology that worries me. I doubt that the North American Breeding Bird Survey is a very precise measure of population changes in any particular area. But following the Bay et al. paper, assume that it is a good measure of changing abundance for the yellow warbler. From the Bay et al. paper abstract we see this prediction:

“Populations requiring the greatest shifts in allele frequencies to keep pace with future climate change have experienced the largest population declines, suggesting that failure to adapt may have already negatively affected populations.”

The prediction is illustrated in Figure 1 below from the Bay et al. paper.

mechanistic hypothesis biology

Consider a single case, the Great Basin, area S09 of the Sauer et al. (2017) breeding bird surveys. From the map in Bay et al. (2018) Figure 2 we get the prediction of a very high genomic vulnerability (above 0.06, approximate red dot in Figure 1 above) for the Great Basin, and thus a strongly declining population trend. But if we go to the Sauer et al. (2017) database, we get this result for the Great Basin (Figure 2 here), a completely stable yellow warbler population for the last 45 years.

mechanistic hypothesis biology

Figure 2. Data for the Great Basin populations of the Yellow Warbler from the North American Breeding Bird Survey, 1967 to 2015 (area S09). (From Sauer et al. 2017)

One clue to this discrepancy is shown in Figure 1 above where R 2 = 0.10, which is to say the predictability of this genomic model is near zero.

So where does this leave us? We have what appears to be an A grade genetic analysis coupled with a D- grade ecological model in which explanations are not based on any mechanism of population dynamics, so that the model presented is useless for any predictions that can be tested in the next 10-20 years. I am far from convinced that this is a useful exercise. It would be a good paper for a graduate seminar discussion. Marvelous genetics, very poor ecology.

And as a footnote I note that mammalian ecologists have already taken a different but more insightful approach to this whole problem of climate-driven adaptation (Boutin and Lane 2014).

Bay, R.A., Harrigan, R.J., Underwood, V.L., Gibbs, H.L., Smith, T.B., and Ruegg, K. 2018. Genomic signals of selection predict climate-driven population declines in a migratory bird. Science 359 (6371): 83-86. doi: 10.1126/science.aan4380.

Blois, J.L., Williams, J.W., Fitzpatrick, M.C., Jackson, S.T., and Ferrier, S. 2013. Space can substitute for time in predicting climate-change effects on biodiversity. Proceedings of the National Academy of Sciences 110 (23): 9374-9379. doi: 10.1073/pnas.1220228110.

Boutin, S., and Lane, J.E. 2014. Climate change and mammals: evolutionary versus plastic responses. Evolutionary Applications 7 (1): 29-41. doi: 10.1111/eva.12121.

Pickett, S.T.A. 1989. Space-for-Time substitution as an alternative to long-term studies. In Long-Term Studies in Ecology: Approaches and Alternatives. Edited by G.E. Likens. Springer New York, New York, NY. pp. 110-135.

Sauer, J.R., Niven, D.K., Hines, J.E., D. J. Ziolkowski, J., Pardieck, K.L., and Fallon, J.E. 2017. The North American Breeding Bird Survey, Results and Analysis 1966 – 2015. USGS Patuxent Wildlife Research Center, Laurel, MD. https://www.mbr-pwrc.usgs.gov/bbs/

On Post-hoc Ecology

Back in the Stone Age when science students took philosophy courses, a logic course was a common choice for students majoring in science. Among the many logical fallacies one of the most common was the Post Hoc Fallacy, or in full “Post hoc, ergo propter hoc”, “After this, therefore because of this.” The Post Hoc Fallacy has the following general form:

  • A occurs before B .
  • Therefore A is the cause of B .

Many examples of this fallacy are given in the newspapers every day. “I lost my pencil this morning and an earthquake occurred in California this afternoon.” Therefore….. Of course, we are certain that this sort of error could never occur in the 21 st century, but I would like to suggest to the contrary that its frequency is probably on the rise in ecology and evolutionary biology, and the culprit ( A ) is most often climate change.

Hilborn and Stearns (1982) pointed out many years ago that most ecological and evolutionary changes have multiple causes, and thus we must learn to deal with multiple causation in which a variety of factors combine and interact to produce an observed outcome. This point of view places an immediate dichotomy between the two extremes of ecological thinking – single factor experiments to determine causation cleanly versus the “many factors are involved” world view. There are a variety of intermediate views of ecological causality between these two extremes, leading in part to the flow chart syndrome of boxes and arrows aptly described by my CSIRO colleague Kent Williams as “horrendograms”. If you are a natural resource manager you will prefer the simple end of the spectrum to answer the management question of ‘what can I possibly manipulate to change an undesirable outcome for this population or community?’

Many ecological changes are going on today in the world, populations are declining or increasing, species are disappearing, geographical distributions are moving toward the poles or to higher altitudes, and novel diseases are appearing in populations of plants and animals. The simplest explanation of all these changes is that climate change is the major cause because in every part of the Earth some aspect of winter or summer climate is changing. This might be correct, or it might be an example of the Post Hoc Fallacy. How can we determine which explanation is correct?

First, for any ecological change it is important to identify a mechanism of change. Climate, or more properly weather, is itself a complex factor of temperature, humidity, and rainfall, and for climate to be considered a proper cause you must advance some information on physiology or behaviour or genetics that would link some specific climate parameter to the changes observed. Information on possible mechanisms makes the potential explanation more feasible. A second step is to make some specific predictions that can be tested either by experiments or by further observational data. Berteaux et al. (2006) provided a careful list of suggestions on how to proceed in this manner, and Tavecchia et al. (2016) have illustrated how one traditional approach to studying the impact of climate change on population dynamics could lead to forecasting errors.

A second critical focus must be on long-term studies of the population or community of interest. In particular, 3-4 year studies common in Ph.D. theses must make the assumption that the results are a random sample of annual ecological changes. Often this is not the case and this can be recognized when longer term studies are completed or more easily if an experimental manipulation can be carried out on the mechanisms involved.

The retort to these complaints about ecological and evolutionary inference is that all investigated problems are complex and multifactorial, so that after much investigation one can conclude only that “many factors are involved”. The application of AIC analysis attempts to blunt this criticism by taking the approach that, given the data (the evidence), what hypothesis is best supported? Hobbs and Hilborn (2006) provide a guide to the different methods of inference that can improve on the standard statistical approach. The AIC approach has always carried with it the awareness of the possibility that the correct hypothesis is not present in the list being evaluated, or that some combination of relevant factors cannot be tested because the available data does not cover a wide enough range of variation. Burnham et al. (2011) provide an excellent checklist for the use of AIC measures to discriminate among hypotheses. Guthery et al. (2005) and Stephens et al. (2005) carry the discussion in interesting ways. Cade (2015) discusses an interesting case in which inappropriate AIC methods lead to questionable conclusions about habitat distribution preferences and use by sage-grouse in Colorado.

If there is a simple message in all this it is to think very carefully about what the problem is in any investigation, what the possible solutions or hypotheses are that could explain the problem, and then utilize the best statistical methods to answer that question. Older statistical methods are not necessarily bad, and newer statistical methods not automatically better for solving problems. The key lies in good data, relevant to the problem being investigated. And if you are a beginning investigator, read some of these papers.

Berteaux, D., et al. 2006. Constraints to projecting the effects of climate change on mammals. Climate Research 32 (2): 151-158. doi: 10.3354/cr032151.

Burnham, K.P., Anderson, D.R., and Huyvaert, K.P. 2011. AIC model selection and multimodel inference in behavioral ecology: some background, observations, and comparisons. Behavioral Ecology and Sociobiology 65 (1): 23-35. doi: 10.1007/s00265-010-1029-6.

Guthery, F.S., Brennan, L.A., Peterson, M.J., and Lusk, J.J. 2005. Information theory in wildlife science: Critique and viewpoint. Journal of Wildlife Management 69 (2): 457-465. doi: 10.1890/04-0645.

Hilborn, R., and Stearns, S.C. 1982. On inference in ecology and evolutionary biology: the problem of multiple causes. Acta Biotheoretica 31 : 145-164. doi: 10.1007/BF01857238

Hobbs, N.T., and Hilborn, R. 2006. Alternatives to statistical hypothesis testing in ecology: a guide to self teaching. Ecological Applications 16 (1): 5-19. doi: 10.1890/04-0645

Stephens, P.A., Buskirk, S.W., Hayward, G.D., and Del Rio, C.M. 2005. Information theory and hypothesis testing: a call for pluralism. Journal of Applied Ecology 42 (1): 4-12. doi: 10.1111/j.1365-2664.2005.01002.x

Tavecchia, G., et al. 2016. Climate-driven vital rates do not always mean climate-driven population. Global Change Biology 22 (12): 3960-3966. doi: 10.1111/gcb.13330.

Climate Change and Ecological Science

One dominant paradigm of the ecological literature at the present time is what I would like to call the Climate Change Paradigm . Stated in its clearest form, it states that all temporal ecological changes now observed are explicable by climate change . The test of this hypothesis is typically a correlation between some event like a population decline, an invasion of a new species into a community, or the outbreak of a pest species and some measure of climate. Given clever statistics and sufficient searching of many climatic measurements with and without time lags, these correlations are often sanctified by p< 0.05. Should we consider this progress in ecological understanding?

An early confusion in relating climate fluctuations to population changes was begun by labelling climate as a density independent factor within the density-dependent model of population dynamics. Fortunately, this massive confusion was sorted out by Enright (1976) but alas I still see this error repeated in recent papers about population changes. I think that much of the early confusion of climatic impacts on populations was due to this classifying all climatic impacts as density-independent factors.

One’s first response perhaps might be that indeed many of the changes we see in populations and communities are indeed related to climate change. But the key here is to validate this conclusion, and to do this we need to talk about the mechanisms by which climate change is acting on our particular species or species group. The search for these mechanisms is much more difficult than the demonstration of a correlation. To become more convincing one might predict that the observed correlation will continue for the next 5 (10, 20?) years and then gather the data to validate the correlation. Many of these published correlations are so weak as to preclude any possibility of validation in the lifetime of a research scientist. So the gold standard must be the deciphering of the mechanisms involved.

And a major concern is that many of the validations of the climate change paradigm on short time scales are likely to be spurious correlations. Those who need a good laugh over the issue of spurious correlation should look at Vigen (2015), a book which illustrates all too well the fun of looking for silly correlations. Climate is a very complex variable and a nearly infinite number of measurements can be concocted with temperature (mean, minimum, maximum), rainfall, snowfall, or wind, analyzed over any number of time periods throughout the year. We are always warned about data dredging, but it is often difficult to know exactly what authors of any particular paper have done. The most extreme examples are possible to spot, and my favorite is this quotation from a paper a few years ago:

“A total of 864 correlations in 72 calendar weather periods were examined; 71 (eight percent) were significant at the p< 0.05 level. …There were 12 negative correlations, p< 0.05, between the number of days with (precipitation) and (a demographic measure). A total of 45- positive correlations, p<0.05, between temperatures and (the same demographic measure) were disclosed…..”

The climate change paradigm is well established in biogeography and the major shifts in vegetation that have occurred in geological time are well correlated with climatic changes. But it is a large leap of faith to scale this well established framework down to the local scale of space and a short-term time scale. There is no question that local short term climate changes can explain many changes in populations and communities, but any analysis of these kinds of effects must consider alternative hypotheses and mechanisms of change. Berteaux et al. (2006) pointed out the differences between forecasting and prediction in climate models. We desire predictive models if we are to improve ecological understanding, and Berteaux et al. (2006) suggested that predictive models are successful if they follow three rules:

(1) Initial conditions of the system are well described (inherent noise is small);

(2) No important variable is excluded from the model (boundary conditions are defined adequately);

(3) Variables used to build the model are related to each other in the proper way (aggregation/representation is adequate).

Like most rules for models, whether these conditions are met is rarely known when the model is published, and we need subsequent data from the real world to see if the predictions are correct.

I am much less convinced that forecasting models are useful in climate research. Forecasting models describe an ecological situation based on correlations among the measurements available with no clear mechanistic model of the ecological interactions involved. My concern was highlighted in a paper by Myers (1998) who investigated for fish populations the success of published juvenile recruitment-environmental factor (typically temperature) correlations and found that very few forecasting models were reliable when tested against additional data obtained after publication. It would be useful for someone to carry out a similar analysis for bird and mammal population models.

Small mammals show some promise for predictive models in some ecosystems. The analysis by Kausrud et al. (2008) illustrates a good approach to incorporating climate into predictive explanations of population change in Norwegian lemmings that involve interactions between climate and predation. The best approach in developing these kinds of explanations and formulating them into models is to determine how the model performs when additional data are obtained in the years to follow publication.

The bottom line is to avoid spurious climatic correlations by describing and evaluating mechanistic models that are based on observable biological factors. And then make predictions that can be tested in a realistic time frame. If we cannot do this, we risk publishing fairy tales rather than science.

Berteaux, D., et al. (2006) Constraints to projecting the effects of climate change on mammals. Climate Research, 32, 151-158. doi: 10.3354/cr032151

Enright, J. T. (1976) Climate and population regulation: the biogeographer’s dilemma. Oecologia, 24, 295-310.

Kausrud, K. L., et al. (2008) Linking climate change to lemming cycles. Nature, 456, 93-97. doi: 10.1038/nature07442

Myers, R. A. (1998) When do environment-recruitment correlations work? Reviews in Fish Biology and Fisheries, 8, 285-305. doi: 10.1023/A:1008828730759

Vigen, T. (2015) Spurious Correlations, Hyperion, New York City. ISBN: 978-031-633-9438

Was the Chitty Hypothesis of Population Regulation a ‘Big Idea’ in Ecology and was it successful?

Jeremy Fox in his ‘Dynamic Ecology’ Blog has raised the eternal question of what have been the big ideas in ecology and were they successful, and this has stimulated me to write about the Chitty Hypothesis and its history since 1952. I will write this from my personal observations which can be faulty, and I will not bother to put in many references since this is a blog and not a formal paper.

In 1952 when Dennis Chitty at Oxford finished his thesis on vole cycles in Wales, he was considered a relatively young heretic because he did not see any evidence in favour of the two dominant paradigms of population dynamics – that populations rose and fell because of food shortage or predation. David Lack vetoed the publication of his Ph.D. paper because he did not agree with Chitty’s findings (Lack believed that food supplies explained all population changes). His 1952 thesis paper was published only because of the intervention of Peter Medawar. Chitty could see no evidence of these two factors in his vole populations and he began to suspect that social factors were involved in population cycles. He tested Jack Christian’s ideas that social stress was a possible cause, since it was well known that some rodents were territorial and highly aggressive, but stress as measured by adrenal gland size did not fit the population trends very well. He then began to suspect that there might be genetic changes in fluctuating vole populations, and that population processes that occurred in voles and lemmings may occur in a wide variety of species, not just in the relatively small group of rodent species, which everyone could ignore as a special case of no generality. This culminated in his 1960 paper in the Canadian Journal of Zoology. This paper stimulated many field ecologists to begin experiments on population regulation in small mammals.

Chitty’s early work contained a ‘big idea’ that population dynamics and population genetics might have something to contribute to each other, and that one could not assume that every individual had equal properties. These ideas of course were not just his, and Bill Wellington had many of the same ideas in studying tent caterpillar population fluctuations. When Chitty suggested these ideas during the late 1950s he was told by several eminent geneticists who must remain nameless that his ideas were impossible, and that ecologists should stay out of genetics because the speed of natural selection was so slow that nothing could be achieved in ecological time. Clearly thinking has now changed on this general idea.

So if one could recognize these early beginnings as a ‘big idea’ it might be stated simply as ‘study individual behaviour, physiology, and genetics to understand population changes’, and it was instrumental in adding another page to the many discussions of population changes that had previously mostly included only predators, food supplies, and potentially disease. All this happened before the rise of behavioural ecology in the 1970s.

I leave others to judge the longer term effects of Chitty’s early suggestions. At present the evidence is largely against any rapid genetic changes in fluctuating populations of mammals and birds, and maternal effects now seem a strong candidate for non-genetic inheritance of traits that affect fitness in a variety of vertebrate species. And in a turn of fate, stress seems to be a strong candidate for at least some maternal effects, and we are back to the early ideas of Jack Christian and Hans Selye of the 1940s, but with greatly improved techniques of measurement of stress in field populations.

Dennis Chitty was a stickler for field experiments in ecology, a trend now long established, and he made many predictions from his ideas, often rejected later but always leading to more insights of what might be happening in field populations. He was a champion of discussing mechanisms of population change, and found little use for the dominant paradigm of the density dependent regulation of populations. Was he successful? I think so, from my biased viewpoint. I note he had less recognition in his lifetime than he deserved because he offended the powers that be. For example, he was never elected to the Royal Society, a victim of the insularity and politics of British science. But that is another story.

Chitty, D. (1952) Mortality among voles ( Microtus agrestis ) at Lake Vyrnwy, Montgomeryshire in 1936-9. Philosophical Transactions of the Royal Society of London, 236, 505-552.

Chitty, D. (1960) Population processes in the vole and their relevance to general theory. Canadian Journal of Zoology, 38, 99-113.

In Defence of Hypothesis Testing in Ecology

In two recent scientific meetings I have attended (which must remain nameless to protect the innocent), I have found myself wondering about the state of hypothesis testing in ecological science. I have always assumed that science consists of testing hypotheses, yet I would estimate roughly that 75% of the talks I have been able to attend showed no sign of any hypothesis. I need to qualify that. Some of these studies are completely descriptive – what species of ferns occur in national park X? Much effort now is devoted to sequencing genomes, the ultimate in descriptive biology. This kind of research work can be classified as alpha-biology, basic description which is necessary before any problems can be formulated. In my particular specialty of population cycles in mammals, much descriptive work had to be carried out to recognize the phenomenon of “cycles”. But then the question arises – at what point should we stop simple descriptions of mammal populations rising and falling? Do we need to study the dynamics of every rodent species that exists? Or in genetics, is our objective to sequence the genome of every species on earth? My point is that after we have enough basic description, we should move into hypothesis testing, or asking why some phenomenon occurs, the mechanisms behind the simple observations. The important point here is that we should not have a single hypothesis or explanation for any set of observations but rather several alternative hypotheses. As a simple example, if we find our favourite plant species is declining in abundance, we should not simply try to connect this decline with climatic warming without having a series of alternative explanations with the emphasis that our observations or experiments should be capable of distinguishing among the alternative hypotheses.

The alternative argument is that we do not know enough about ecological systems to set up a series of credible alternative hypotheses. It is quite possible to go on describing events endlessly in science in the hope that some wisdom will emerge. I do not think this is a profitable use of time or money in science. In ecology in particular I would argue that there is not a single question one can ask that cannot be answered by at least 2 or 3 different mechanistic hypotheses. Our job is to articulate these alternatives and to do whatever studies or experiments are needed to distinguish among them. Of course it is always possible that the correct answer is not among the 2 or 3 hypotheses we suggest at the start of an investigation, and this is often why one study leads to a further one. Consequently we cannot accept statements like “I have no idea why this observation has occurred”. Such a statement means you have not thought deeply enough about what you are studying. Ecological surprises certainly occur while we study any particular community or ecosystem, but we know enough now to suggest several possible mechanisms by which any ecological surprise might be generated.

So I think it incumbent on every ecologist to ask (1) what is the problem or question my research is addressing? And (2) what probable mechanisms can be invoked as the cause of this problem or the answer to this question. Vagueness may be a virtue in politics but it is not a virtue in science. And I look forward to future conferences in which every paper specifies a precise hypothesis and alternative hypotheses. Chamberlin (1897) stated the case for multiple hypotheses, Karl Popper (1963) asked very specifically what your hypothesis forbids from happening, and John Platt (1964) pulled it together in a critical paper. There was important work done before the Iphone was invented. Good reading.

Chamberlin, T. C. 1897. The method of multiple working hypotheses. Journal of Geology 5 :837-848 (reprinted in Science 148: 754-759 in 1965).

Platt, J. R. 1964. Strong inference. Science 146 :347-353.

Popper, K. R. 1963. Conjectures and Refutations: The Growth of Scientific Knowledge . Routledge and Kegan Paul, London.

Follow Ecological Rants

Get every new post delivered to your Inbox

Join other followers:

Mechanistic microclimate models and plant pest risk modelling

  • Open access
  • Published: 10 May 2024

Cite this article

You have full access to this open access article

mechanistic hypothesis biology

  • Jonathan R. Mosedale   ORCID: orcid.org/0000-0001-9008-5439 1 ,
  • Dominic Eyre 2 ,
  • Anastasia Korycinska 2 ,
  • Matthew Everatt 2 ,
  • Sam Grant 2 ,
  • Brittany Trew   ORCID: orcid.org/0000-0002-0649-828X 1 ,
  • Neil Kaye 3 ,
  • Deborah Hemming   ORCID: orcid.org/0000-0001-5245-0546 3 , 4 &
  • Ilya M. D. Maclean   ORCID: orcid.org/0000-0001-8030-9136 1  

82 Accesses

Explore all metrics

Climatic conditions are key determining factors of whether plant pests flourish. Models of pest response to temperature are integral to pest risk assessment and management, helping to inform surveillance and control measures. The widespread use of meteorological data as predictors in these models compromises their reliability as these measurements are not thermally coupled to the conditions experienced by pest organisms or their body temperatures. Here, we present how mechanistic microclimate models can be used to estimate the conditions experienced by pest organisms to provide significant benefits to pest risk modelling. These well-established physical models capture how landscape, vegetation and climate interact to determine the conditions to which pests are exposed. Assessments of pest risk derived from microclimate conditions are likely to significantly diverge from those derived from weather station measurements. The magnitude of this divergence will vary across a landscape, over time and according to pest habitats and behaviour due to the complex mechanisms that determine microclimate conditions and their effect on pest biology. Whereas the application of microclimate models was once restricted to relatively homogeneous habitats, these models can now be applied readily to generate hourly time series across extensive and varied landscapes. We outline the benefits and challenges of more routine application of microclimate models to pest risk modelling. Mechanistic microclimate models provide a heuristic tool that helps discriminate between physical, mathematical and biological causes of model failure. Their use can also help understand how pest ecology, behaviour and physiology mediate the relationship between climate and pest response.

Avoid common mistakes on your manuscript.

Key messages

The use of meteorological data is a common source of errors in pest risk modelling.

We propose the more frequent use of mechanistic microclimate models.

These models estimate the actual conditions to which pest organisms are exposed.

They help us understand pest responses to climate and identify causes of pest model failure.

Routine integration of microclimate modelling into pest risk modelling systems is now feasible

Introduction

Plant pests reduce agricultural productivity and food security with annual losses reaching c.18% of global crop production (Oerke 2006 ; Savary et al. 2019 ; Fones et al. 2020 ). Significant economic costs are also imposed on the forestry sector where pest outbreaks reduce ecosystem function, resilience and biodiversity (Godfray et al. 2010 ; Bradshaw et al. 2016 ). The routine application of pesticides and other measures to control pests incurs additional economic, environmental and human health costs (Pretty et al. 2000 ; Pimentel and Burgess 2014 ; Bourguet and Guillemaud 2016 ). These impacts of plant pests are forecast to change as the distributions of many pest species shift poleward, and invasive species more readily establish, under warmer climatic conditions (Bebber et al. 2013 , 2014 ).

Quantitative pest risk models form a key component of pest management; the ability to forecast and map the risk of pest outbreaks is essential for the deployment of effective monitoring and control methods (Rock and Shaffer 1983 ; Tonnang et al. 2017 ; Coop et al. 2020 ; Hemming and Macneill 2020 ). Various indicators of pest risk can be modelled, informed by pest biology and the mitigation options available. Quantitative measures of pest risk may be derived from models of pest survival, growth or development. For an insect pest, model outputs may be a predicted date of completing a key developmental stage, or indicators of population growth such as voltinism or the likely exposure to causes of pest mortality. For a fungal pest or pathogen growth, model outputs may describe risk in terms of the favourability of conditions for pest growth, reproduction or dispersal. Models may express the response of a ‘typical’ or ‘average’ individual, a set of individuals or that of a pest population.

Climate is fundamental for determining the conditions in which ectothermic pest species flourish, imposing limits on their distribution and governing the biological processes that determine individual fitness and population dynamics (Uvarov 1931 ; Taylor 1981 ; Gillooly et al. 2001 ; Damos and Savopoulou-Soultani 2011 ; Clarke 2017 ). Climate-related variables are therefore key predictors in nearly all pest risk modelling. Long-term pest risk projections are commonly based on current or future climate suitability to endemic or invasive pests (Meyerson and Reaser 2002 ; Venette et al. 2010 ; Magarey et al. 2011 ; Hlásny et al. 2021 ), whereas short-term projections, that improve the efficacy and lower the economic and environmental costs of pest control measures, are typically driven by seasonal weather and short-range forecasts (Brunner et al. 1982 ; Damos and Savopoulou-Soultani 2010 ; Crimmins et al. 2020 ; Barker et al. 2020 ). Models used to inform the management of insect pests are often driven solely by temperature (Taylor 1981 ; Gillooly et al. 2001 ; Damos and Savopoulou-Soultani 2011 ), while models of slug pests (Choi et al. 2004 ; Wilson et al. 2015 ) or of fungal pathogens and oomycetes, are commonly expressed as functions of temperature and moisture (Jacome et al. 1991 ; Huber and Gillespie 1992 ; Magarey and Sutton 2007 ).

Although statistical models have been employed in pest risk analyses, for example to extrapolate from existing distributions of pest species to predict their potential establishment and range in new regions or under different climate scenarios (Lantschner et al. 2019 ), mechanistic or semi-mechanistic approaches are generally preferred (Sutherst and Maywald 1985 ; Savary et al. 2019 ) due to the acknowledged limitations of purely statistical approaches (Sutherst 2014 ; Lantschner et al. 2019 ; Srivastava et al. 2021 ). The use of mechanistic models facilitates a ‘biophysical’ approach to pest risk modelling which has particular benefit when forecasting how pests respond to novel conditions, whether as a result of climate change, pest or host migration. Under such conditions, the relationships between variables described by purely statistical models are less likely to endure than those derived from explicit biophysical processes and interactions.

A key aspect of a biophysical approach is an ability to reliably measure or estimate the abiotic conditions under which organisms live. Unfortunately, the importance of climate-derived variables in pest risk modelling does not denote the suitability of standard meteorological data as model inputs. Indeed, the use of meteorological data has been identified as one of the most common causes of failure in pest forecasting (Magarey and Isard 2017 ) whether in the form of data from individual weather stations (Trnka et al. 2007 ; Elliott et al. 2011 ; Jacquemin et al. 2014 ; Babu et al. 2014 ), in-field weather loggers (Chuang et al. 2014 ) or as gridded climate datasets derived from interpolated meteorological station data (Jarvis and Baker 2001 ; Hijmans et al. 2005 ; Magarey et al. 2011 ; Kriticos et al. 2012 ; Fick and Hijmans 2017 ; Lantschner et al. 2019 ; Crimmins et al. 2020 ; Barker et al. 2020 ).

The problem is that neither the microclimates experienced by pest organisms nor their body temperatures are thermally coupled to standard meteorological data derived from weather stations (Baker 1980 ; Ferro et al. 1979 ; Ruesink 1976 ; Samietz et al. 2007 ; Wellington 1950 ). This divergence between meteorological data and the conditions experienced by organisms is not simply due to a mismatch between the spatial resolution of meteorological data and the size of organisms (sensu Potter et al. 2013 ). Weather stations, measuring ambient air temperature at 2 m above ground, are designed and located to minimise the effects of local topography and vegetation on energy fluxes between the biosphere and atmosphere (WMO 2018 ). Yet it is these same energy fluxes that determine the microclimate conditions experienced by organisms (Daly 2006 ; WMO 2018 ; Bramer et al. 2018 ). Furthermore, the complex interactions mediating the relationship between meteorological and microclimate conditions cannot be reduced to a simple correlative relationship (Monteith and Unsworth 2013 ; Briscoe et al. 2023 ).

In this paper, we argue that the continued use of standard meteorological data in pest risk analyses ignores advances in mechanistic microclimate modelling over the past fifty years (Waggoner and Reifsnyder 1968 ; Goudriaan 1977 ; Deardorff 1978 ; Ogée et al. 2003 ; Kearney and Porter 2017 ; Maclean and Klinges 2021 ). These models are based on well-established physical processes (Richardson 1922 ; Penman 1948 ; Monteith and Szeicz 1962 ) and are today able to model microclimate conditions, at fine temporal and spatial resolutions, across entire landscapes (e.g. Trew et al. 2022 ). We propose their application can help resolve key shortcomings in existing biophysical approaches to pest risk modelling.

High-resolution microclimate and other environmental data forms a key part of new approaches to agricultural and forest management, such as ‘precision agriculture’ (Robert 2002 ). These approaches use data generated by ‘in-field’ sensor technologies and remote sensing to tailor management inputs and activities, including pest control, to conditions within individual fields or plots (Schultze et al. 2021 ). The deployment and maintenance of an extensive network of sensors, however, remains expensive and labour-intensive (Skocir et al. 2021 ), especially outside of highly controlled or homogeneous environments. The integration of mechanistic microclimate models has the potential of lowering the cost and extending the scope of precision agriculture approaches.

In the first part of this paper, we consider how pest thermal responses are measured and represented in pest risk models. In doing so, we revisit long-standing debates on how the type of mathematical model, conditions of fitting and use of meteorological data in model application can all introduce significant bias to model results. Secondly, we consider how microclimate conditions diverge from local weather conditions and the implications for pest risk modelling. Thirdly, we outline the theory of mechanistic microclimate modelling and how recent technological advances are transforming access to and use of these models. In the final part, we outline potential applications, and remaining challenges, of using the outputs of mechanistic microclimate models as inputs to pest risk models. We argue that the integration of microclimate models provides a powerful heuristic tool to improve our understanding of pest ecology and the reliability of pest risk modelling systems.

Although we focus primarily on modelling the effects of temperature on insect pests, our arguments apply to the modelling of other climate-related conditions relevant to pest and pathogen risk, such as incident radiation, humidity, soil moisture and leaf wetness.

Modelling thermal response

Thermal response (or ‘performance’) curves, ‘TPCs’, (Angilletta 2009 ) describe the relationship between temperature and a biological function or process. TPCs are typically derived from empirical observations made at different constant temperatures under controlled conditions (Quinn 2017 ; Shi et al. 2017 ). The concept commonly underlies models of pest life cycles and phenology, voltinism and population dynamics.

When measured over a sufficiently wide range of temperatures, ectothermic organisms typically display a skewed, unimodal response to temperature (Vasseur et al. 2014 ). The most likely cause of this response are the thermodynamics of enzyme reaction rates and protein stability. The response can be described by the fitting of various types of nonlinear statistical models to empirical observations (Kontodimas et al. 2004 ; Damos and Savopoulou-Soultani 2011 ; Rebaudo and Rabhi 2018 ). Pest modelling typically applies TPCs to actual field conditions using rate summation, where the response (such as developmental completion) is measured by accumulating response fractions per unit time (Liu et al. 1995 ). A key assumption of the approach is that the thermal response at any given instant is independent of the wider thermal regime.

Despite the empirical evidence and theory supporting a nonlinear response to temperature as the norm, in practice the majority of invertebrate pest risk models apply a simplified linear response (Quinn 2017 ; Shi et al. 2017 ; Rebaudo and Rabhi 2018 ). In part, this reflects the greater effort and resources required to fit a nonlinear response across a temperature range. The use of linear responses is also justified by assuming organisms are rarely exposed to extreme temperatures (Campbell et al. 1974 ). Among the most common insect pest risk models are those that use units of heat accumulation, such as degree days, to predict developmental rates. In such cases, the x-axis intercept of the thermal response curve defines a lower limit for the calculation of heat accumulation and the reciprocal of the slope defines a thermal constant that designates the amount of heat accumulation required, for example, to complete a developmental stage or life cycle (see Fig.  1 ).

figure 1

Demonstrating the effects of different temperature data on linear and nonlinear thermal response models. In a, response curves of linear (green) and Brière 2 nonlinear (violet) models of codling moth ( Cydia pomonella ) larval development (Aghdam et al. 2009 ) that are applied to the frequency distributions of three contrasting hypothetical temperature datasets (derived from normal distributions with different means and standard deviations). The range of temperature values where both model responses are approximately linear is also shown. In b–d , we show model responses to each temperature dataset and indicate both the mean response to the temperature distribution (dashed lines) and the response to overall mean temperature of the distribution (dash-dot lines) with colours corresponding to the model type as in ( a ). In ( b ), temperatures exceed the shared zone of linearity in model response and hence for the nonlinear model the mean response differs from the response to mean temperature, whereas the corresponding values for the linear model are the same. In ( c ), temperatures fall below the shared zone of linearity and therefore the mean response differs from the response to mean temperature for both linear and nonlinear models. In ( d ), temperatures vary predominantly within the shared zone of linearity and hence mean response and response to mean temperature are very similar for both response models. The differences between mean response and response to mean temperature of the different models demonstrate how both the choice of model type and the use of aggregated temperature data can introduce bias in the predictions of thermal pest response. Note response and temperature scales are identical for figures ( b – d ) although minimum and maximum values vary

In principle, the type of model used to describe the thermal response, and the conditions under which it is fitted, should determine the temperature data used for model application. In practice, however, this logic is often inverted. Readily available weather station measurements, in the form of daily averages or degree days, have been widely used as inputs to pest response models (Babu et al. 2014 ; Magarey and Isard 2017 ; Crimmins et al. 2020 ; Barker et al. 2020 ). By using such data, the likelihood increases of values falling within the range of linear response (Fig.  1 ) and can suggest the suitability of a simplified, linear model of thermal response. However, the approach also introduces several well-documented sources of error that can affect the reliability of not only linear, but also nonlinear, response models.

Firstly, the statistical method used to calculate degree days (or daily averages) and the extrapolation of a linear response to higher temperatures, or use of subjective cut-offs, can all affect model predictions, as has been acknowledged and discussed for several decades (Allen 1976 ; Higley et al. 1986 ; McMaster and Wilhelm 1997 ; Bonhomme 2000 ).

Of greater significance is the implicit presumption when using spatially or temporally aggregated temperature data that pest response to mean temperature is the same as the mean pest response to fluctuating temperatures. It is a presumption that does not prevail when the underlying response is nonlinear and temperatures fluctuate significantly—irrespective of whether the TPC is simplified to a linear response or not. The implications of this mathematical property of nonlinear functions, referred to as Jensen’s inequality (Jensen 1906 ; Ruel and Ayres 1999 ), have been widely discussed in pest modelling literature by reference to rate summation (Tanigoshi et al. 1976 ) and the Kaufmann effect (Bryant et al. 1999 ; Ikemoto and Egami 2013 ). The key implication is that changes to temperature variance can be as important as changes to mean temperature in determining a biological response (Ruel and Ayres 1999 ; Vasseur et al. 2014 ; Bütikofer et al. 2020 ). Referring to Fig.  1 , temperature fluctuation around T low will result in a higher accumulated response than that predicted from the mean temperature, while variation around the optimal temperature will result in a lower response. The magnitude of the divergence will depend upon the degree of nonlinearity and skewness of the response curve, as well as temperature mean and variation.

Finally, and most importantly, when models are fitted to observations made under controlled and constant conditions, ambient air temperature corresponds to pest microclimate and body temperature. No such correspondence exists when applied to field conditions where the use of meteorological temperatures, measured within standardised environments (e.g. Stevenson screen at 2 m above ground level), exclude many of the factors, such as direct sunlight and wind, that affect pest microclimates and their body temperatures.

In summary, the application of models using rate summation from TPCs requires not only addressing the nonlinearity of thermal responses and how they are studied, but also the spatiotemporal variation of temperature to which organisms are exposed in the field (Maiorano et al. 2012 ; von Schmalensee et al. 2021 ). Recent studies (von Schmalensee et al. 2021 ) have demonstrated how the use of nonlinear TPCs driven by microclimate measurements can provide much more accurate predictions of pest biology than using linear degree day models or nonlinear TPCs driven by weather station data. These same studies suggest reliability is more affected by microclimate divergence than the parameterisation of TPCs under constant conditions. The divergence of pest microclimates from meteorological conditions is therefore a key element and challenge to the reliable application of pest modelling.

The importance of pest microclimates

The concept of microclimate does not simply refer to the spatial or temporal resolution of climate observations but incorporates the effects of topography and vegetation on energy fluxes and hence on local air and surface conditions (see Table  1 ). These effects vary ‘horizontally’ across a landscape and ‘vertically’ relative to a canopy, trunk, leaf or ground surface. A range of empirical and theoretical studies have shown how differences in microclimate conditions can affect the growth, development, mortality and diversity of pest organisms (Baker 1980 ; Pincebourde and Woods 2012 ; Saudreau et al. 2013 ; Faye et al. 2017 ; Pincebourde and Casas 2019 ).

The divergence between microclimate and meteorological conditions varies according to the type of microhabitat, prevailing weather conditions, the spatiotemporal scale and the type of metric. Locations near the top of vegetative canopies typically experience greater extremes than ambient air temperature during the day. Greater heat accumulation at the top of the canopy occurs because radiative heating of leaf surfaces generally exceeds cooling except under conditions of low sunlight. Locations lower within the canopy are buffered from air temperature extremes experienced above the canopy (De Frenne et al. 2019 , 2021 ; Haesen et al. 2021 ). This relationship between microclimate and ambient air temperatures varies under different conditions. Summer temperatures beneath a dense canopy are generally lower than above-canopy air temperatures due to canopy shading, but reduced air movement within the canopy can cause warmer relative temperatures under certain wind conditions.

The divergence between meteorological data and microclimate conditions varies not only relative to a canopy or ground surface, but also spatially across a landscape. Many of these landscape effects are often termed ‘mesoclimatic’ (Maclean et al. 2019 ) and include the effects of elevation, topographical wind sheltering, cold-air drainage and coastal exposure. At high spatial resolutions, the effects of slope and aspect on the exposure to direct solar radiation can significantly affect surface and beneath-canopy temperatures. Regions characterised by more variable topography will often result in more variable microclimate conditions. In alpine environments, for example, diurnal and seasonal averages may diverge from nearby weather station data by up to 20 °C (Scherrer and Körner 2011 ).

Microclimate conditions therefore not only play a role in determining current geographical ranges, but also mediate biotic responses to climate change and thereby the propensity of current ranges to enlarge or diminish under future conditions (Maclean and Early 2023 ). The importance of microclimate conditions may be most significant at the margins of a pest range where it is more subject to climate-related limiting factors. Microclimates can permit species to persist under otherwise unfavourable macroclimatic conditions, or conversely, cause local extinctions despite a favourable climate (Suggitt et al. 2018 ; Ma and Ma 2022 ). Depending on the type of habitat and weather, a microclimate may reduce or increase exposure to the extreme temperatures that can be a direct cause of pest mortality (Alford et al. 2018 ; Pincebourde and Casas 2019 ). Winter minima temperatures, for example, are often associated with cold-air pooling and temperature inversions (Cooke and Roland 2003 ) that are influenced by fine-scale topographical features and rarely captured by standard weather station measurements (Schultze et al. 2021 ). Conversely, under dense canopy, minimum temperatures are generally higher than in open areas due to the effect of vegetation (De Frenne et al. 2021 ).

The importance of microclimate conditions to the reliable application of pest risk and crop growth models is well-established (Cellier et al. 1993 ; Bonhomme 2000 ) but not often captured. Where pest risk modelling has explicitly addressed the importance of microclimate, it has often done so by estimating local mesoclimate or microclimates from meteorological measurements by statistical downscaling methods using covariates of those properties, such as elevation or aspect, thought to affect local conditions (Royer et al. 1989 ; Marques da Silva et al. 2015 ; Rebaudo et al. 2016 ; Zellweger et al. 2019 ; Ogris et al. 2019 ). These estimated local temperatures are then used to drive thermal response-type models parameterised under constant, controlled conditions. An alternative approach uses models parameterised from observed pest distributions or behaviour in the field and relates these directly to local meteorological measurements.

In certain cases, different suites of models develop according to whether they are derived primarily from controlled experiments or from field observations. For the bark beetle, Ips typographus , one set of models uses estimates of bark temperature to drive nonlinear response models parameterised under controlled conditions (Baier et al. 2007 ; Ogris et al. 2019 ). Another set of models uses meteorological data to drive a degree day response parameterised from field observations, to map pest ranges under current and future conditions (Jönsson et al. 2007 , 2011 ; Bentz et al. 2019 ).

Although the second approach does avoid discrepancy between the temperature data used in model parameterisation and model application, it does not address the issue of microclimates.

Whereas the first approach explicitly downscales meteorological data to local conditions, the second approach subsumes the processes relating weather to microclimate into the model of pest response to meteorological conditions. Both approaches suffer the same limitation; namely that the relationships they describe are valid within highly prescribed spatiotemporal limits. These limits are imposed by the many and complex physical processes that relate macroclimate to microclimate conditions. These processes vary in relative importance according to topographical, vegetative and weather conditions. Models depend on local parameterisation (where this is possible) and are unreliable when applied to novel combinations of geography and climate.

The adoption of a biophysical approach, that includes direct measurement and/or mechanistic modelling of the abiotic conditions to which pest organisms are exposed, offers a way to improve the reliability of pest risk modelling. Such an approach can explicitly address (i) the physical mechanisms determining the relationship between meteorological conditions and those experienced by pest organisms, (ii) how these processes vary spatially and temporally across the landscape, and (iii) how microclimate conditions relate to the body temperature of pest organisms. The development of more performative and accessible mechanistic microclimate models over recent years, offers the potential of their routine incorporation in integrated pest risk modelling and mapping.

Mechanistic microclimate modelling

Physical theory and model methods.

The mechanisms that determine microclimate conditions reflect well-understood physical processes (Bramer et al. 2018 ) expressed by equations for the exchange of energy and mass between different components of the environment and with organisms within that environment (Campbell and Norman 1998 ; Briscoe et al. 2023 ).

The mechanistic modelling of above-canopy microclimates has its origin in the work on weather forecasting by Richardson ( 1922 ), who demonstrated the basic laws of turbulent mixing in the surface layer of the atmosphere. Theory and methods were subsequently developed in the agricultural and forest sciences and have been well-established since the 1950s (Penman 1948 ; Monin and Obukhov 1954 ; MacHattie and McCormack 1961 ; Monteith and Szeicz 1962 ; Allen et al. 1976 ; Goudriaan 1977 ). Local conditions relative to a ground or vegetative surface are modelled using the physical processes that govern energy exchange (sensible, latent and radiative heat fluxes). By including a term for the rate of heat storage (by the ground or by vegetation), the energy fluxes can be assumed to reach steady state allowing equations to be rearranged to solve for temperature iteratively or using for example the Penman–Monteith equation (e.g. Jarvis & Stewart 1979 ).

The effects of terrain, vegetation and soil properties are incorporated by calculating their effects on components of the energy budget. The incorporation of mesoclimate processes, such as lapse rates, coastal effects and wind sheltering effects of terrain, must be informed by whether the weather data used as model inputs have partly or fully accounted for such effects. Local field measurements may capture many effects while the gridded products interpolated from weather stations often include correction for elevation, coastal and urban heat effects (Hollis et al. 2019 ) to the spatial resolution of the final product.

Beneath a canopy, vegetative properties and canopy structure affect radiation transmission, vapour exchange and wind profiles, which in turn determine heat exchange, air and surface temperatures. Modelling microclimates beneath canopies, and within the soil, is typically undertaken by estimating heat and vapour fluxes between a series of homogeneous horizontal layers (Maclean and Klinges 2021 ). Establishing temperature profiles requires a theory that quantifies how heat is exchanged between these levels in the canopy. Doing so using the same principles of heat transfer that are used in above-canopy models would imply that vertical heat exchange occurs from warm to cool parts of the canopy in a predictable way determined by average wind speed (i.e. a gradient diffusion process). However, accumulated evidence (reviewed by Finnigan & Raupach 1987 , and Raupach 1988 ) shows that this is not a valid assumption as frequent and pronounced counter gradient fluxes can be observed (Denmead and Bradley 1985 , 1987 ). An alternative theory for modelling microclimate within canopies has been developed by adopting a Lagrangian (fluid-following) viewpoint (e.g. Raupach 1989 ). Here, the dispersion pattern of heat is assumed to be determined by the motion of the air that exchanges heat from multiple leaves within the canopy. Of dominant importance is the effect of the persistence of the turbulence relative to the travel time of the heat. At any given point in the canopy, heat arrives from multiple sources with varying travel times, but by assuming the canopy to be horizontally homogeneous, it is possible to find a computationally tractable solution to derive the temperature profile. These principles form the basis of several more recent below-canopy microclimate models (e.g. Bonan et al. 2021 ; Ogée et al. 2003 ) meaning that it is possible to estimate the microclimate temperature of almost any canopy environment.

Mechanistic microclimate models are typically used to return time series at specified heights or depths relative to the ground and/or canopy surface. Time increments can vary from seconds to several days, and outputs are usually for single locations or gridded points across a landscape. The outputs variables will typically include not only air or surface temperatures but also humidity, wind speeds and solar radiation which in turn allow integration with the biophysical modelling of pest organism body temperatures if required.

Model inputs and interfaces

Awareness of mechanistic microclimate models has permeated slowly across disciplines within the biological sciences. Early use was largely limited to the study of relatively homogeneous environments, such as crop fields, for which vegetation, topographical and other parameters could be measured in situ and/or constant values applied. Only in recent years have we begun to realise many potential applications of these models (Briscoe et al. 2023 ) .

Microclimate models are typically run at high temporal and spatial resolutions, making model execution computationally intensive. The array of data inputs (see Fig.  2 ) and significant processing resources required for microclimate modelling presented major barriers to the application of these models only a decade ago (Sutherst 2014 ). Today, many of the same technologies that are transforming agricultural practices—including numerical weather forecasting, geographical positioning systems, in-field sensors, remote sensing and unmanned aerial vehicles—also facilitate the routine and extensive use of microclimate models.

figure 2

Key to microclimate modelling is the climate forcing data used to drive models. Model re-analyses data, such as ECMWF Re-Analysis 5 (Hersbach et al. 2020 ) and NCEP re-analysis dataset (Saha et al. 2014 ), provide a physically consistent, hourly description of global atmospheric characteristics, including cloud cover and incoming radiation. These datasets assimilate empirical weather station observations with short-range, weather forecasting models to generate low spatial, but high temporal, resolution climate data. Remote sensing products provide high spatial resolution topographical and land surface data, including properties such as albedo (Bonafoni and Sekertekin 2020 ).

An additional requirement of many microclimate models are descriptors of vegetation structure, such as plant area index and canopy height, to permit the quantification of microclimate processes such as radiation transfer through a canopy. For some applications, constant values may suffice, or dynamic canopy or crop development models can be applied to simulate seasonal changes in vegetative properties. Such models may be driven by the same climate inputs already described.

In other cases, the direct or remote measurement of actual canopy characteristics may be required, either as direct inputs into microclimate models or as drivers of crop growth or canopy development models (Toda and Richardson 2018 ; Kasampalis et al. 2018 ). Novel data products derived from sensors such as light detection and ranging (LIDAR), digital aerial photogrammetry, thermal radiometry and synthetic aperture radar (Jensen et al. 2021 ; Steele-Dunne et al. 2017 ), including those mounted on the International Space Station (Fisher et al. 2020 ; Dubayah et al. 2020 ), demonstrate the ability to quantify the vertical structure and properties of plant canopies across a landscape. The availability of global datasets of climate, land surface and increasingly canopy structure, combined with faster processing and programmatic access, today allows the near real-time modelling of microclimates across landscapes at high spatial and temporal resolution (Zellweger et al. 2019 ; Kearney et al. 2020 ; Duffy et al. 2021 ; Klinges et al. 2022 ).

Programmatic access to the inputs of microclimate models (Kemp et al. 2012 ; Kearney et al. 2020 ; Klinges et al. 2022 ) also facilitates the development of on-line portals that generate microclimate conditions (e.g. http://bioforecasts.science.unimelb.edu.au/app_direct/soil/ ), improving access to microclimate modelling for non-specialist users. Existing portals generate time series for single point locations, but could be extended to provide more extensive microclimate modelling for specific crop or forest systems. Faster, high-level programming languages (Bezanson et al. 2017 ), grid-modelling frameworks, access to cloud computing services (Yang et al. 2017 ) and statistical emulations of mechanistic microclimate models using Gaussian processes (Baker et al. 2022 ; Gómez-Dans et al. 2016 ) all offer the opportunity for more rapid modelling over larger geographical and temporal extents.

Recent years have also seen the publication of freely available regional, continental and global datasets of microclimate measurements. These include hourly gridded temperature estimates of above and below ground microclimates under different shade conditions or substrate types (Kearney et al. 2014 ; Kearney 2019 ) and a compilation of soil and near-surface temperature measurements from across the globe (Lembrechts et al. 2020 ).

The range of potential data inputs, algorithmic methods and precomputed microclimate datasets is such that guidance concerning their suitability for different end-uses is now required (Meyer et al. 2023 ). Technological advances have also greatly facilitated the validation of microclimate model outputs, whether by direct measurement of remotely sensed surface temperature (Marques da Silva et al. 2015 ; Zellweger et al. 2019 ) or the availability of reliable and cheap temperature and humidity micro-loggers (Lembrechts et al. 2020 ; Maclean et al. 2021 ).

Applying microclimate models to pest risk assessment and management

The greatest benefit of mechanistic microclimate models will be realised when they form part of a broader biophysical approach to pest risk modelling. Potential applications of the approach include the short-term forecasting of pest risk, typically to inform the timing and targeting of pest monitoring or treatments, as well as longer-term risk mapping of potential shifts in geographical range of invasive or migratory species and/or as a response to climate change (Kearney and Porter 2009 ). Microclimate conditions under future climate scenarios (Maclean 2020 ; Briscoe et al. 2023 ) can be modelled using probabilistic estimates of future projected climate from climate/earth system models (Smith et al. 2017 ) or by using the outputs of regional climate models. The outputs can be used to estimate pest risk under future climate scenarios in an analogous approach to that used with crop growth models to forecast the suitability of novel crops (Gardner et al. 2021 ).

There remain, however, many challenges to realising the potential of mechanistic microclimate models as a component of pest risk modelling and analysis. The integration of microclimate models in pest risk modelling requires a sufficient understanding of pest ecology to provide information on the microhabitats occupied by the organism at key stages in its life cycle. Existing microclimate models may need to be extended to capture the conditions of specific habitats. For example, modelling the conditions experienced by wood-boring beetles, a group of pests of significant ecological and economic concern, would need to capture heat fluxes within the host tree trunk. Field observations of whether pests display a preference between available microhabitats, such as height within a canopy, or the sunny or shady sides of a tree trunk (Gent et al. 2017 ), can also inform the choice of model inputs and parameters.

In general, the body temperature of small ectotherms with a high surface area to volume ratio remains closely coupled to surrounding air and/or surface temperatures, particularly where wind speeds are non-negligible. Pest physiology and behavioural responses will nonetheless mediate how microclimate conditions affect their body temperatures. Where necessary, heat fluxes between an organism and its surface and air microclimates can be integrated into microclimate models to capture, for example, convective and radiative heat transfers (Pincebourde and Woods 2020 ; Kearney et al. 2021 ).

Microclimate modelling also renders explicit the selection of biologically relevant spatial and temporal scales used to inform risk modelling. This choice of scale must be informed by the type of pest response and microhabitat that are modelled. Whereas models of insect development or phenology may be driven by daily data, measures of pest exposure to extreme temperatures may require modelling at higher temporal resolutions. Similarly, a lower temporal resolution may prove adequate for modelling buffered, soil microclimates, than for fluctuating microclimates like leaf surfaces. The choice of spatial resolution when modelling microclimates across a landscape will also need to consider variation in the processes and underlying properties determinant of microclimate conditions. A higher spatial resolution may be required to capture the effects of mountainous or hilly terrain.

The integration of mechanistic microclimate modelling must reflect the end-use of pest risk modelling, and tolerance to different kinds of error. If mapping a measure of pest risk across a landscape grid, then it may be appropriate to model the most and least favourable microclimates within individual grid cells. For example, where a pest can occupy habitats throughout a canopy, modelling microclimates at the top and bottom of the canopy can provide a plausible range of risk. Appropriate weighting might be applied to reflect the propensity of different habitats available to a pest or its typical distribution within a canopy. In cases where pest risk mapping does not need to reflect actual landcover, the adoption of canopy parameters typical of a habitat may be appropriate. Conversely, in cases where it is important to reflect actual landcover characteristics, remote sensing observations or in-field sensors can provide estimates of actual vegetative properties.

In all cases, the computational demands of microclimate modelling will also need to be weighed against potential benefits. Even with more efficient algorithms and exploitation of cloud computing, the mapping of multiple microclimates across extensive spatiotemporal extents requires significant resources in terms of time and computational resource.

Microclimate modelling as a heuristic tool

Progress in the development of more reliable pest risk models has been hampered by a dearth of published information about the occurrences and causes of model failure (Magarey and Isard 2017 ), augmented by the use of model systems which conflate different sources of error. This includes the use of meteorological data as a descriptor of pest abiotic conditions. As a result, it has proved difficult to resolve important questions about the constancy of thermal response models derived from controlled environment studies, or the relative importance of pest adaptation or acclimatisation to abiotic conditions (von Schmalensee et al. 2021 ).

We believe the use of mechanistic microclimate models offers a valuable heuristic tool to help discriminate between potential causes of model failure. The validity of thermal response curves continues to be debated when used to predict pest response under fluctuating field conditions (Niehaus et al. 2012 ; Colinet et al. 2015 ; Khelifa et al. 2019 ; Ma et al. 2021 ). It is known that the thermal response of ectotherms can vary between different organisms and populations of the same species, even if the extent of such variation remains uncertain (Chuine and Régnière 2017 ). The speed and duration of temperature fluctuations can affect insect response to high or low temperatures through thermal stress accumulation, temperature-induced or time-dependent acclimation (Sinclair et al. 2016 ; Sears et al. 2019 ). Emerald ash borer pre-pupae, for example, have greater heat tolerance when shifted slowly to high temperatures (Sobek et al. 2011 ). Under certain conditions, some invertebrates may even display semi-permanent endothermy, maintaining higher body temperature than their environment (Dinets 2022 ).

The use of meteorological data as a measure of the conditions to which pests are exposed conflates potential biological explanations of why a model may fail to describe responses in the field, with physical processes mediating effects on microclimate and with mathematical effects such as Jensen’s inequality. Mechanistic microclimate models provide a means to disentangle these different effects and assess different biological explanations of the discrepancies between observed and predicted pest response. Their use complements carefully designed empirical studies and re-analyses of published data which suggest thermal response models are reliable subject to consideration of nonlinearity and spatiotemporal heterogeneity in microclimate conditions (von Schmalensee et al. 2021 ).

A biophysical approach also permits better understanding of how behavioural and physiological adaptations mediate pest response to climate, whether by selecting from available microclimates, modifying their microclimate conditions or altering the effect of their microclimate on body temperatures (Table  2 ). In such cases, the thermal response model remains valid if driven by suitable data. With an adequate knowledge of pest biology, we can adapt mechanistic microclimate models to capture many of these effects.

For example, if it is known that a pest can actively move between different available microclimates, we might assume it will select the microclimate where conditions are closest to their optimal response. Modelling microclimate conditions at contrasting locations within a host plant canopy, at an appropriate temporal resolution that reflects the mobility of the organism and microclimate variability, can provide a range of available microclimate conditions from which the most optimal conditions could be selected. Adaptive responses are likely to reduce exposure to temperatures lying outside the zone of linear response (see Fig.  1 ), but it cannot be inferred that the modelling of microclimate conditions is less important as this also depends on the coupling between microclimate and meteorological conditions.

Modifications of pest microclimates through feeding and/or the creation of specialist structures such as leaf mines may require the integration of additional models (Pincebourde and Casas 2006 ). Integrated biophysical models offer a means of capturing how plant parameters are changed by pest feeding activity (Pincebourde and Casas 2019 ) and their resultant impact on microclimate conditions.

The use of microclimate modelling with dynamic plant canopy and pest developmental models also offers the possibility of better understanding how the correspondence between plant and pest phenology varies, an important aspect of pest risk, or of informing control methods based on the management of canopy structure (Barradas and Fanjul 1986 ; Huber and Gillespie 1992 ; Pangga et al. 2011 ; Caffarra et al. 2012 ; Saudreau et al. 2013 ). A better understanding of the spatiotemporal distribution of abiotic conditions to which pests are exposed can help improve our understanding of the limiting factors that act upon pest populations. Where available microclimate conditions approach optimal conditions, biological limiting factors may exert greater influence (Łaszczyca et al. 2021 ).

The integration of microclimate, plant canopy and pest body temperature models has been undertaken for over a decade (Saudreau et al. 2013 ) and such biophysical model systems have contributed to our understanding of pest biology and control. However, until recently it has been impractical to apply such an approach to risk modelling across entire landscapes and multi-year timeframes. We believe the new generation of microclimate models removes many of these barriers to their use and can make a significant contribution towards more reliable pest risk modelling.

Recent advances in remote sensing and computing resources have greatly expanded the capacity of microclimate modelling to estimate conditions across wide spatiotemporal extents at high resolutions. Although there remain challenges to the modelling of processes determining canopy microclimates, the fundamental principles of mechanistic microclimate modelling have been resolved for several decades.

Mechanistic microclimate models bring significant benefits to pest risk modelling where the widespread use of meteorological data as model inputs conflates various physical, experimental, mathematical and biological mechanisms that can undermine the reliability of risk forecasting. By approximating the conditions experienced by pest organisms and their body temperatures, microclimate estimates, (i) more closely correspond to the data used in the development and fitting of thermal response models under controlled environment conditions; (ii) provide a better estimate of pest exposure to extreme temperatures; (iii) capture spatiotemporal variations and interdependences of landscape, vegetative and climate characteristics that determine the conditions to which pests are exposed; and (iv) help discriminate between possible causes of model failure.

To fully exploit the benefits of microclimate modelling, nonlinear pest models that better reflect thermal responses at extreme temperature need to be adopted. Application to pest risk management also requires an adequate understanding of pest habitat and ecology, including how ecological, behavioural and physiological adaptations mediate the relationships between body temperatures and weather conditions. We believe the adoption of mechanistic microclimate modelling is a key element for building a more robust biophysical approach to pest risk management, informed by pest ecology, the physical determinants of microclimates and the practical requirements of end users.

The application of mechanistic microclimate models helps refocus attention on pest ecology, in particular the availability and occupancy of microhabitats that determine the abiotic conditions to which pests are exposed, and those behavioural or physiological adaptations that mediate the relationship with pest body temperature.

Aghdam HR, Fathipour Y, Radjabi G, Rezapanah M (2009) Temperature-dependent development and temperature thresholds of codling moth (Lepidoptera: Tortricidae) in Iran. Environ Entomol 38:885–895. https://doi.org/10.1603/022.038.0343

Article   PubMed   Google Scholar  

Alford L, Tougeron K, Pierre J-S et al (2018) The effect of landscape complexity and microclimate on the thermal tolerance of a pest insect. Insect Sci 25:905–915. https://doi.org/10.1111/1744-7917.12460

Allen JC (1976) A modified sine wave method for calculating degree days 1. Environ Entomol 5:388–396. https://doi.org/10.1093/ee/5.3.388

Article   Google Scholar  

Allen LH, Sinclair TR, Lemon ER (1976) Radiation and microclimate relationships in multiple cropping systems. Multiple cropping. John Wiley & Sons Ltd, New Jersey, pp 171–200. https://doi.org/10.2134/asaspecpub27.c9

Chapter   Google Scholar  

Angilletta MJ Jr (2009) Looking for answers to questions about heat stress: researchers are getting warmer. Funct Ecol 23:231–232. https://doi.org/10.1111/j.1365-2435.2009.01548.x

Babu A, Cook DR, Caprio MA et al (2014) Prevalence of Helicoverpa zea (Lepidoptera: Noctuidae) on late season volunteer corn in Mississippi: implications on Bt resistance management. Crop Prot 64:207–214. https://doi.org/10.1016/j.cropro.2014.06.005

Baier P, Pennerstorfer J, Schopf A (2007) PHENIPS—a comprehensive phenology model of Ips typographus (L.) (Col., Scolytinae) as a tool for hazard rating of bark beetle infestation. For Ecol Manag 249:171–186. https://doi.org/10.1016/j.foreco.2007.05.020

Baker C (1980) Some problems in using meteorological data to forecast the timing of insect life cycles. EPPO Bull 10:83–91. https://doi.org/10.1111/j.1365-2338.1980.tb02628.x

Baker E, Harper AB, Williamson D, Challenor P (2022) Emulation of high-resolution land surface models using sparse Gaussian processes with application to JULES. Geosci Model Dev 15:1913–1929. https://doi.org/10.5194/gmd-15-1913-2022

Barker BS, Coop L, Wepprich T et al (2020) DDRP: real-time phenology and climatic suitability modeling of invasive insects. PLoS ONE 15:e0244005. https://doi.org/10.1371/journal.pone.0244005

Article   CAS   PubMed   PubMed Central   Google Scholar  

Barradas VL, Fanjul L (1986) Microclimatic chacterization of shaded and open-grown coffee ( Coffea arabica L.) plantations in Mexico. Agric for Meteorol 38:101–112. https://doi.org/10.1016/0168-1923(86)90052-3

Barton M, Porter W, Kearney M (2014) Behavioural thermoregulation and the relative roles of convection and radiation in a basking butterfly. J Therm Biol 41:65–71. https://doi.org/10.1016/j.jtherbio.2014.02.004

Bebber DP, Ramotowski MAT, Gurr SJ (2013) Crop pests and pathogens move polewards in a warming world. Nat Clim Change 3:985–988. https://doi.org/10.1038/nclimate1990

Bebber DP, Holmes T, Gurr SJ (2014) The global spread of crop pests and pathogens. Glob Ecol Biogeogr 23:1398–1407. https://doi.org/10.1111/geb.12214

Bentz BJ, Jönsson AM, Schroeder M, Weed A, Wilcke RAI, Larsson K (2019) Ips typographus and Dendroctonus ponderosae models project thermal suitability for intra- and inter-continental establishment in a changing climate. Front For Glob Change. https://doi.org/10.3389/ffgc.2019.00001

Bezanson J, Edelman A, Karpinski S, Shah VB (2017) Julia: a fresh approach to numerical computing. SIAM Rev 59:65–98. https://doi.org/10.1137/141000671

Bonafoni S, Sekertekin A (2020) Albedo Retrieval from sentinel-2 by new narrow-to-broadband conversion coefficients. IEEE Geosci Remote Sens Lett 17:1618–1622. https://doi.org/10.1109/LGRS.2020.2967085

Bonan GB, Patton EG, Finnigan JJ et al (2021) Moving beyond the incorrect but useful paradigm: reevaluating big-leaf and multilayer plant canopies to model biosphere-atmosphere fluxes – a review. Agric for Meteorol 306:108435. https://doi.org/10.1016/j.agrformet.2021.108435

Bonhomme R (2000) Bases and limits to using ‘degree.day’ units. Eur J Agron 13:1–10. https://doi.org/10.1016/S1161-0301(00)00058-7

Bourguet D, Guillemaud T (2016) The hidden and external costs of pesticide use. In: Lichtfouse E (ed) Sustainable agriculture reviews, vol 19. Springer International Publishing, Cham, pp 35–120

Bradshaw CJA, Leroy B, Bellard C et al (2016) Massive yet grossly underestimated global costs of invasive insects. Nat Commun 7:12986. https://doi.org/10.1038/ncomms12986

Bramer I, Anderson BJ, Bennie J et al (2018) Chapter three-advances in monitoring and modelling climate at ecologically relevant scales. In: Bohan DA, Dumbrell AJ, Woodward G, Jackson M (eds) Advances in ecological research. Academic Press, Cambridge, pp 101–161. https://doi.org/10.1016/bs.aecr.2017.12.005

Briscoe NJ, Morris SD, Mathewson PD et al (2023) Mechanistic forecasts of species responses to climate change: The promise of biophysical ecology. Glob Change Biol 29:1451–1470. https://doi.org/10.1111/gcb.16557

Article   CAS   Google Scholar  

Brunner JF, Hoyt SC, Wright MA (1982) Codling moth control—a new tool for timing sprays. Ext Bull-wash State Univ Coop Ext Serv

Bryant SR, Bale JS, Thomas CD (1999) Comparison of development and growth of nettle-feeding larvae of Nymphalidae (Lepidoptera) under constant and alternating temperature regimes. Eur J Entomol 96:143–148

Google Scholar  

Bütikofer L, Anderson K, Bebber DP et al (2020) The problem of scale in predicting biological responses to climate. Glob Change Biol 26:6657–6666. https://doi.org/10.1111/gcb.15358

Caffarra A, Rinaldi M, Eccel E et al (2012) Modelling the impact of climate change on the interaction between grapevine and its pests and pathogens: European grapevine moth and powdery mildew. Agric Ecosyst Environ Complet. https://doi.org/10.1016/j.agee.2011.11.017

Campbell GS, Norman JM (1998) An introduction to environmental biophysics. Springer, New York, NY

Book   Google Scholar  

Campbell A, Frazer BD, Gilbert N et al (1974) Temperature requirements of some aphids and their parasites. J Appl Ecol 11:431–438. https://doi.org/10.2307/2402197

Cellier P, Ruget F, Chartier M, Bonhomme R (1993) Estimating the temperature of a maize apex during early growth stages. Agric for Meteorol 63:35–54. https://doi.org/10.1016/0168-1923(93)90021-9

Choi YH, Bohan DA, Powers SJ et al (2004) Modelling Deroceras reticulatum (Gastropoda) population dynamics based on daily temperature and rainfall. Agric Ecosyst Environ 103:519–525. https://doi.org/10.1016/j.agee.2003.11.012

Chuang C-L, Yang E-C, Tseng C-L et al (2014) Toward anticipating pest responses to fruit farms: revealing factors influencing the population dynamics of the oriental fruit fly via automatic field monitoring. Comput Electron Agric 109:148–161. https://doi.org/10.1016/j.compag.2014.09.018

Chuine I, Régnière J (2017) Process-based models of phenology for plants and animals. Annu Rev Ecol Evol Syst 48:159–182. https://doi.org/10.1146/annurev-ecolsys-110316-022706

Clarke A (2017) Principles of thermal ecology: temperature, energy and life. Oxford University Press, Oxford

Clench HK (1966) Behavioral thermoregulation in butterflies. Ecology 47:1021–1034. https://doi.org/10.2307/1935649

Colinet H, Sinclair BJ, Vernon P, Renault D (2015) Insects in fluctuating thermal environments. Annu Rev Entomol 60:123–140. https://doi.org/10.1146/annurev-ento-010814-021017

Article   CAS   PubMed   Google Scholar  

Cooke BJ, Roland J (2003) The effect of winter temperature on forest tent caterpillar (Lepidoptera: Lasiocampidae) egg survival and population dynamics in northern climates. Environ Entomol 32:299–311. https://doi.org/10.1603/0046-225X-32.2.299

Coop L, Barker B, Kogan M, Heinrichs E (2020) Advances in understanding species ecology: phenological and life cycle modeling of insect pests. Integrated management of insect pests: current and future developments. Sawston, England, pp 43–96

Crimmins TM, Gerst KL, Huerta DG et al (2020) Short-term forecasts of insect phenology inform pest management. Ann Entomol Soc Am 113:139–148. https://doi.org/10.1093/aesa/saz026

Daly C (2006) Guidelines for assessing the suitability of spatial climate data sets. Int J Climatol 26:707–721. https://doi.org/10.1002/joc.1322

Damos PT, Savopoulou-Soultani M (2010) Development and statistical evaluation of models in forecasting moth phenology of major lepidopterous peach pest complex for Integrated Pest Management programs. Crop Prot 29:1190–1199. https://doi.org/10.1016/j.cropro.2010.06.022

Damos PT, Savopoulou-Soultani M (2011) Temperature-driven models for insect development and vital thermal requirements. Psyche (stuttg) 2012:e123405. https://doi.org/10.1155/2012/123405

De Frenne P, Zellweger F, Rodríguez-Sánchez F et al (2019) Global buffering of temperatures under forest canopies. Nat Ecol Evol 3:744–749. https://doi.org/10.1038/s41559-019-0842-1

De Frenne P, Lenoir J, Luoto M et al (2021) Forest microclimates and climate change: importance, drivers and future research agenda. Glob Change Biol 27:2279–2297. https://doi.org/10.1111/gcb.15569

Deardorff JW (1978) Efficient prediction of ground surface temperature and moisture, with inclusion of a layer of vegetation. J Geophys Res Oceans 83:1889–1903. https://doi.org/10.1029/JC083iC04p01889

Denmead OT, Bradley EF (1985) Flux-Gradient Relationships in a Forest Canopy. In: Hutchison BA, Hicks BB (eds) The Forest-Atmosphere Interaction: Proceedings of the Forest Environmental Measurements Conference held at Oak Ridge, Tennessee, October 23–28, 1983. Springer Netherlands, Dordrecht, pp 421–442

Denmead OT, Bradley EF (1987) On Scalar Transport in Plant Canopies. Irrig Sci 8:131–149. https://doi.org/10.1007/BF00259477

Dinets V (2022) First case of endothermy in semisessile animals. J Exp Zool Part Ecol Integr Physiol 337:111–114. https://doi.org/10.1002/jez.2547

Dubayah R, Blair JB, Goetz S et al (2020) The global ecosystem dynamics investigation: high-resolution laser ranging of the earth’s forests and topography. Sci Remote Sens 1:100002. https://doi.org/10.1016/j.srs.2020.100002

Duffy JP, Anderson K, Fawcett D et al (2021) Drones provide spatial and volumetric data to deliver new insights into microclimate modelling. Landsc Ecol 36:685–702. https://doi.org/10.1007/s10980-020-01180-9

Elliott RH, Mann L, Olfert O (2011) Calendar and degree-day requirements for emergence of adult Macroglenes penetrans (Kirby), an egg-larval parasitoid of wheat midge, Sitodiplosis mosellana (Géhin). Crop Prot 30:405–411. https://doi.org/10.1016/j.cropro.2010.12.007

Faye E, Rebaudo F, Carpio C et al (2017) Does heterogeneity in crop canopy microclimates matter for pests? Evidence from aerial high-resolution thermography. Agric Ecosyst Environ 246:124–133. https://doi.org/10.1016/j.agee.2017.05.027

Ferro DN, Chapman RB, Penman DR (1979) Observations on insect microclimate and insect pest management 12. Environ Entomol 8:1000–1003. https://doi.org/10.1093/ee/8.6.1000

Fey SB, Vasseur DA, Alujević K et al (2019) Opportunities for behavioral rescue under rapid environmental change. Glob Change Biol 25:3110–3120. https://doi.org/10.1111/gcb.14712

Fick SE, Hijmans RJ (2017) WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. Int J Climatol 37:4302–4315. https://doi.org/10.1002/joc.5086

Finnigan JJ, Raupach MR (1987) Transfer processes in plant canopies in relation to stomatal characteristics. Stanford University Press

Fisher JB, Lee B, Purdy AJ et al (2020) ECOSTRESS: NASA’s Next Generation Mission to Measure Evapotranspiration From the International Space Station. Water Resour Res 56:e2019WR026058. https://doi.org/10.1029/2019WR026058

Fones HN, Bebber DP, Chaloner TM et al (2020) Threats to global food security from emerging fungal and oomycete crop pathogens. Nat Food 1:332–342. https://doi.org/10.1038/s43016-020-0075-0

Gardner AS, Maclean IMD, Gaston KJ, Bütikofer L (2021) Forecasting future crop suitability with microclimate data. Agric Syst 190:103084. https://doi.org/10.1016/j.agsy.2021.103084

Gent CA, Wainhouse D, Day K et al (2017) Temperature-dependent development of the great European spruce bark beetle Dendroctonus micans (Kug.) (Coleoptera: Curculionidae: Scolytinae) and its predator Rhizophagus grandis Gyll. (Coleoptera: Monotomidae: Rhizophaginae). Agric For Entomol 19:321–331. https://doi.org/10.1111/afe.12212

Gillooly JF, Brown JH, West GB et al (2001) Effects of size and temperature on metabolic rate. Science 293:2248–2251. https://doi.org/10.1126/science.1061967

Godfray HCJ, Crute IR, Haddad L et al (2010) The future of the global food system. Philos Trans R Soc B Biol Sci 365:2769–2777. https://doi.org/10.1098/rstb.2010.0180

Gómez-Dans JL, Lewis PE, Disney M (2016) Efficient emulation of radiative transfer codes using gaussian processes and application to land surface parameter inferences. Remote Sens 8:119. https://doi.org/10.3390/rs8020119

Goudriaan J (1977) Crop micrometeorology: a simulation study. Wageningen University and Research ProQuest Dissertations Publishing, Netherlands

Guo F, Guénard B, Economo EP et al (2020) Activity niches outperform thermal physiological limits in predicting global ant distributions. J Biogeogr 47:829–842. https://doi.org/10.1111/jbi.13799

Haesen S, Lembrechts JJ, De Frenne P et al (2021) ForestTemp—sub-canopy microclimate temperatures of European forests. Glob Change Biol 27:6307–6319. https://doi.org/10.1111/gcb.15892

Hemming D, Macneill K (2020) Use of meteorological data in biosecurity. Emerg Top Life Sci 4:497–511. https://doi.org/10.1042/ETLS20200078

Article   PubMed   PubMed Central   Google Scholar  

Hersbach H, Bell B, Berrisford P et al (2020) The ERA5 global reanalysis. Q J R Meteorol Soc 146:1999–2049. https://doi.org/10.1002/qj.3803

Higley LG, Pedigo LP, Ostlie KR (1986) Degday: a program for calculating degree-days, and assumptions behind the degree-day approach. Environ Entomol 15:999–1016. https://doi.org/10.1093/ee/15.5.999

Hijmans RJ, Cameron SE, Parra JL et al (2005) Very high resolution interpolated climate surfaces for global land areas. Int J Climatol 25:1965–1978. https://doi.org/10.1002/joc.1276

Hlásny T, König L, Krokene P et al (2021) Bark beetle outbreaks in Europe: state of knowledge and ways forward for management. Curr For Rep 7:138–165. https://doi.org/10.1007/s40725-021-00142-x

Hollis D, McCarthy M, Kendon M et al (2019) HadUK-Grid—a new UK dataset of gridded climate observations. Geosci Data J 6:151–159. https://doi.org/10.1002/gdj3.78

Huber L, Gillespie TJ (1992) Modeling leaf wetness in relation to plant disease epidemiology. Annu Rev Phytopathol 30:553–577. https://doi.org/10.1146/annurev.py.30.090192.003005

Ikemoto T, Egami C (2013) Mathematical elucidation of the Kaufmann effect based on the thermodynamic SSI model. Appl Entomol Zool. https://doi.org/10.1007/s13355-013-0190-6

Jacome L, Schuh W, Stevenson R (1991) Effect of temperature and relative humidity on germination and germ tube development of Mycosphaerella fijiensis var. difformis . Phytopathology 81:1480–1485

Jacquemin G, Chavalle S, De Proft M (2014) Forecasting the emergence of the adult orange wheat blossom midge, Sitodiplosis mosellana (Géhin) (Diptera: Cecidomyiidae) in Belgium. Crop Prot 58:6–13. https://doi.org/10.1016/j.cropro.2013.12.021

Jarvis CH, Baker RHA (2001) Risk assessment for nonindigenous pests: 1. Mapping the outputs of phenology models to assess the likelihood of establishment. Divers Distrib 7:223–235. https://doi.org/10.1046/j.1366-9516.2001.00113.x

Jarvis PG, Stewart J (1979) Evaporation of water from plantation forest. In: Ford ED, Atterson J (eds) The ecology of even-aged forest plantations. Institute of Terrestrial Ecology, Cambridge, pp 327–350

Jensen J (1906) Sur les fonctions convexes et les inégalités entre les valeurs moyennes. Acta Math 30:175–193. https://doi.org/10.1007/BF02418571

Jensen PO, Meddens AJH, Fisher S et al (2021) Broaden your horizon: the use of remotely sensed data for modeling populations of forest species at landscape scales. For Ecol Manag 500:119640. https://doi.org/10.1016/j.foreco.2021.119640

Jönsson AM, Harding S, Bärring L, Ravn HP (2007) Impact of climate change on the population dynamics of Ips typographus in southern Sweden. Agric for Meteorol 146:70–81. https://doi.org/10.1016/j.agrformet.2007.05.006

Jönsson AM, Harding S, Krokene P et al (2011) Modelling the potential impact of global warming on Ips typographus voltinism and reproductive diapause. Clim Change 109:695–718. https://doi.org/10.1007/s10584-011-0038-4

Kasampalis DA, Alexandridis TK, Deva C et al (2018) Contribution of remote sensing on crop models: a review. J Imaging 4:52. https://doi.org/10.3390/jimaging4040052

Kearney MR (2019) MicroclimOz—a microclimate data set for Australia, with example applications. Austral Ecol 44:534–544. https://doi.org/10.1111/aec.12689

Kearney M, Porter W (2009) Mechanistic niche modelling: combining physiological and spatial data to predict species’ ranges. Ecol Lett 12:334–350. https://doi.org/10.1111/j.1461-0248.2008.01277.x

Kearney MR, Porter WP (2017) NicheMapR—an R package for biophysical modelling: the microclimate model. Ecography 40:664–674. https://doi.org/10.1111/ecog.02360

Kearney MR, Isaac AP, Porter WP (2014) microclim: Global estimates of hourly microclimate based on long-term monthly climate averages. Sci Data 1:140006. https://doi.org/10.1038/sdata.2014.6

Kearney MR, Gillingham PK, Bramer I et al (2020) A method for computing hourly, historical, terrain-corrected microclimate anywhere on earth. Methods Ecol Evol 11:38–43. https://doi.org/10.1111/2041-210X.13330

Kearney MR, Porter WP, Huey RB (2021) Modelling the joint effects of body size and microclimate on heat budgets and foraging opportunities of ectotherms. Methods Ecol Evol 12:458–467. https://doi.org/10.1111/2041-210X.13528

Kemp MU, van Loon EE, Shamoun-Baranes J, Bouten W (2012) RNCEP: global weather and climate data at your fingertips. Methods Ecol Amp Evol 3:65–70

Khelifa R, Blanckenhorn WU, Roy J et al (2019) Usefulness and limitations of thermal performance curves in predicting ectotherm development under climatic variability. J Anim Ecol 88:1901–1912. https://doi.org/10.1111/1365-2656.13077

Kingsolver JG (1985) Thermoregulatory significance of wing melanization in Pieris butterflies (Lepidoptera: Pieridae): physics, posture, and pattern. Oecologia 66:546–553. https://doi.org/10.1007/BF00379348

Kingsolver JG, Buckley LB (2020) Ontogenetic variation in thermal sensitivity shapes insect ecological responses to climate change. Curr Opin Insect Sci 41:17–24. https://doi.org/10.1016/j.cois.2020.05.005

Klinges DH, Duffy JP, Kearney MR, Maclean IMD (2022) mcera5: driving microclimate models with ERA5 global gridded climate data. Methods Ecol Evol. https://doi.org/10.1111/2041-210X.13877

Kontodimas DC, Eliopoulos PA, Stathas GJ, Economou LP (2004) Comparative temperature-dependent development of Nephus includens (Kirsch) and Nephus bisignatus (Boheman) (Coleoptera: Coccinellidae) Preying on Planococcus citri (Risso) (Homoptera: Pseudococcidae): evaluation of a linear and various nonlinear models using specific criteria. Environ Entomol 33:1–11. https://doi.org/10.1603/0046-225X-33.1.1

Kriticos DJ, Webber BL, Leriche A et al (2012) CliMond: global high-resolution historical and future scenario climate surfaces for bioclimatic modelling. Methods Ecol Evol 3:53–64. https://doi.org/10.1111/j.2041-210X.2011.00134.x

Lantschner MV, de la Vega G, Corley JC (2019) Predicting the distribution of harmful species and their natural enemies in agricultural, livestock and forestry systems: an overview. Int J Pest Manag 65:190–206. https://doi.org/10.1080/09670874.2018.1533664

Łaszczyca P, Nakonieczny M, Kędziorski A et al (2021) Towards understanding Cameraria ohridella (Lepidoptera: Gracillariidae) development: effects of microhabitat variability in naturally growing horse-chestnut tree canopy. Int J Biometeorol 65:1647–1658. https://doi.org/10.1007/s00484-021-02119-8

Lembrechts JJ, Aalto J, Ashcroft MB et al (2020) SoilTemp: a global database of near-surface temperature. Glob Change Biol 26:6616–6629. https://doi.org/10.1111/gcb.15123

Liu S-S, Zhang G-M, Zhu J (1995) Influence of temperature variations on rate of development in insects: analysis of case studies from entomological literature. Ann Entomol Soc Am 88:107–119. https://doi.org/10.1093/aesa/88.2.107

Ma G, Ma C-S (2022) Potential distribution of invasive crop pests under climate change: incorporating mitigation responses of insects into prediction models. Curr Opin Insect Sci 49:15–21. https://doi.org/10.1016/j.cois.2021.10.006

Ma G, Bai C-M, Wang X-J et al (2018) Behavioural thermoregulation alters microhabitat utilization and demographic rates in ectothermic invertebrates. Anim Behav 142:49–57. https://doi.org/10.1016/j.anbehav.2018.06.003

Ma C-S, Ma G, Pincebourde S (2021) Survive a warming climate: insect responses to extreme high temperatures. Annu Rev Entomol 66:163–184. https://doi.org/10.1146/annurev-ento-041520-074454

MacHattie LB, McCormack RJ (1961) Forest microclimate: a topographic study in Ontario. J Ecol 49:301–323. https://doi.org/10.2307/2257264

Maclean IMD (2020) Predicting future climate at high spatial and temporal resolution. Glob Change Biol 26:1003–1011. https://doi.org/10.1111/gcb.14876

Maclean IMD, Early R (2023) Macroclimate data overestimate range shifts of plants in response to climate change. Nat Clim Change. https://doi.org/10.1038/s41558-023-01650-3

Maclean IMD, Klinges DH (2021) Microclimc: A mechanistic model of above, below and within-canopy microclimate. Ecol Model 451:109567. https://doi.org/10.1016/j.ecolmodel.2021.109567

Maclean IMD, Mosedale JR, Bennie JJ (2019) Microclima: an r package for modelling meso- and microclimate. Methods Ecol Evol 10:280–290. https://doi.org/10.1111/2041-210X.13093

Maclean IMD, Duffy JP, Haesen S et al (2021) On the measurement of microclimate. Methods Ecol Evol 12:1397–1410. https://doi.org/10.1111/2041-210X.13627

Maeno KO, Piou C, Kearney MR et al (2021) A general model of the thermal constraints on the world’s most destructive locust Schistocerca gregaria . Ecol Appl. https://doi.org/10.1002/eap.2310

Magarey RD, Isard SA (2017) A troubleshooting guide for mechanistic plant pest forecast models. J Integr Pest Manag. https://doi.org/10.1093/jipm/pmw015

Magarey RD, Sutton TB (2007) How to create and deploy infection models for plant pathogens. In: Ciancio A, Mukerji KG (eds) General concepts in integrated pest and disease management. Springer, Netherlands, Dordrecht, pp 3–25

Magarey RD, Borchert DM, Engle JS et al (2011) Risk maps for targeting exotic plant pest detection programs in the United States. EPPO Bull 41:46–56. https://doi.org/10.1111/j.1365-2338.2011.02437.x

Maiorano A, Bregaglio S, Donatelli M et al (2012) Comparison of modelling approaches to simulate the phenology of the European corn borer under future climate scenarios. Ecol Model 245:65–74. https://doi.org/10.1016/j.ecolmodel.2012.03.034

Marques da Silva JR, Damásio CV, Sousa AMO et al (2015) Agriculture pest and disease risk maps considering MSG satellite data and land surface temperature. Int J Appl Earth Obs Geoinformation 38:40–50. https://doi.org/10.1016/j.jag.2014.12.016

McGaughran A, Laver R, Fraser C (2021) Evolutionary responses to warming. Trends Ecol Evol 36:591–600. https://doi.org/10.1016/j.tree.2021.02.014

McMaster GS, Wilhelm WW (1997) Growing degree-days: one equation, two interpretations. Agric for Meteorol 87:291–300. https://doi.org/10.1016/S0168-1923(97)00027-0

Meyer AV, Sakairi Y, Kearney MR, Buckley LB (2023) A guide and tools for selecting and accessing microclimate data for mechanistic niche modeling. Ecosphere 14:e4506. https://doi.org/10.1002/ecs2.4506

Meyerson LA, Reaser JK (2002) Biosecurity: moving toward a comprehensive approach: a comprehensive approach to biosecurity is necessary to minimize the risk of harm caused by non-native organisms to agriculture, the economy, the environment, and human health. Bioscience 52:593–600. https://doi.org/10.1641/0006-3568(2002)052[0593:BMTACA]2.0.CO;2

Monin AS, Obukhov AM (1954) Basic laws of turbulent mixing in the surface layer of the atmosphere. Contrib Geophys Inst Acad Sci USSR 24(151):163–187

Monteith JL, Szeicz G (1962) Radiative temperature in the heat balance of natural surfaces. Q J R Meteorol Soc 88:496–507. https://doi.org/10.1002/qj.49708837811

Monteith J, Unsworth M (2013) Principles of environmental physics: plants, animals, and the atmosphere. Academic Press, Cambridge. https://doi.org/10.1016/B978-0-12-386910-4.00001-9

Niehaus AC, Angilletta MJ Jr, Sears MW et al (2012) Predicting the physiological performance of ectotherms in fluctuating thermal environments. J Exp Biol 215:694–701. https://doi.org/10.1242/jeb.058032

Oerke E-C (2006) Crop losses to pests. J Agric Sci 144:31–43. https://doi.org/10.1017/S0021859605005708

Ogée J, Brunet Y, Loustau D et al (2003) MuSICA, a CO 2 , water and energy multilayer, multileaf pine forest model: evaluation from hourly to yearly time scales and sensitivity analysis. Glob Change Biol 9:697–717. https://doi.org/10.1046/j.1365-2486.2003.00628.x

Ogris N, Ferlan M, Hauptman T et al (2019) RITY – A phenology model of Ips typographus as a tool for optimization of its monitoring. Ecol Model 410:108775. https://doi.org/10.1016/j.ecolmodel.2019.108775

Ohsaki N (1986) Body temperatures and behavioural thermoregulation strategies of threePieris butterflies in relation to solar radiation. J Ethol 4:1–9. https://doi.org/10.1007/BF02348247

Pangga IB, Hanan J, Chakraborty S (2011) Pathogen dynamics in a crop canopy and their evolution under changing climate. Plant Pathol 60:70–81. https://doi.org/10.1111/j.1365-3059.2010.02408.x

Penman HL (1948) Natural evaporation from open water, bare soil and grass. Proc R Soc Lond Ser Math Phys Sci 193:120–145. https://doi.org/10.1098/rspa.1948.0037

Pimentel D, Burgess M (2014) Environmental and Economic Costs of the Application of Pesticides Primarily in the United States. In: Pimentel D, Peshin R (eds) Integrated Pest Management: Pesticide Problems. Springer, Netherlands,Dordrecht, pp 47–71

Pincebourde S, Casas J (2006) Multitrophic biophysical budgets: thermal ecology of an intimate herbivore insect-plant interaction. Ecol Monogr 76:175–194. https://doi.org/10.1890/0012-9615(2006)076[0175:MBBTEO]2.0.CO;2

Pincebourde S, Casas J (2019) Narrow safety margin in the phyllosphere during thermal extremes. Proc Natl Acad Sci 116:5588–5596. https://doi.org/10.1073/pnas.1815828116

Pincebourde S, Woods HA (2012) Climate uncertainty on leaf surfaces: the biophysics of leaf microclimates and their consequences for leaf-dwelling organisms. Funct Ecol 26:844–853. https://doi.org/10.1111/j.1365-2435.2012.02013.x

Pincebourde S, Woods HA (2020) There is plenty of room at the bottom: microclimates drive insect vulnerability to climate change. Curr Opin Insect Sci 41:63–70. https://doi.org/10.1016/j.cois.2020.07.001

Pincebourde S, Dillon ME, Woods HA (2021) Body size determines the thermal coupling between insects and plant surfaces. Funct Ecol 35:1424–1436. https://doi.org/10.1111/1365-2435.13801

Poitou L, Robinet C, Suppo C et al (2021) When insect pests build their own thermal niche: the hot nest of the pine processionary moth. J Therm Biol 98:102947. https://doi.org/10.1016/j.jtherbio.2021.102947

Potter KA, Arthur Woods H, Pincebourde S (2013) Microclimatic challenges in global change biology. Glob Change Biol 19:2932–2939. https://doi.org/10.1111/gcb.12257

Pretty JN, Brett C, Gee D et al (2000) An assessment of the total external costs of UK agriculture. Agric Syst 65:113–136. https://doi.org/10.1016/S0308-521X(00)00031-7

Quinn BK (2017) A critical review of the use and performance of different function types for modeling temperature-dependent development of arthropod larvae. J Therm Biol 63:65–77. https://doi.org/10.1016/j.jtherbio.2016.11.013

Raupach MR (1988) Canopy transport processes. In: Steffen WL, Denmead OT (eds) Flow and transport in the natural environment: advances and applications. Springer, Berlin, Heidelberg, pp 95–127

Raupach MR (1989) Applying Lagrangian fluid mechanics to infer scalar source distributions from concentration profiles in plant canopies. Agric for Meteorol 47:85–108. https://doi.org/10.1016/0168-1923(89)90089-0

Rebaudo F, Rabhi V-B (2018) Modeling temperature-dependent development rate and phenology in insects: review of major developments, challenges, and future directions. Entomol Exp Appl 166:607–617. https://doi.org/10.1111/eea.12693

Rebaudo F, Faye E, Dangles O (2016) Microclimate data improve predictions of insect abundance models based on calibrated spatiotemporal temperatures. Front Physiol 7:139. https://doi.org/10.3389/fphys.2016.00139

Richardson LF (1922) Weather prediction by numerical process. Cambridge University Press, Cambridge

Robert PC (2002) Precision agriculture: a challenge for crop nutrition management. Plant Soil 247:143–149. https://doi.org/10.1023/A:1021171514148

Rock GC, Shaffer PL (1983) developmental rates of codling moth (Lepidoptera: Olethreutidae) reared on apple at four constant temperatures1. Environ Entomol 12:831–834. https://doi.org/10.1093/ee/12.3.831

Rodrigues YK, Beldade P (2020) Thermal plasticity in insects’ response to climate change and to multifactorial environments. Front Ecol Evol. https://doi.org/10.3389/fevo.2020.00271

Royer MH, Russo JM, Kelley JGW (1989) Plant disease prediction using a mesoscale weat her forecasting technique. Plant Dis 73:618. https://doi.org/10.1094/PD-73-0618

Ruel JJ, Ayres MP (1999) Jensen’s inequality predicts effects of environmental variation. Trends Ecol Evol 14:361–366. https://doi.org/10.1016/S0169-5347(99)01664-X

Ruesink WG (1976) Status of the systems approach to pest management. Annu Rev Entomol 21:27–44. https://doi.org/10.1146/annurev.en.21.010176.000331

Saha S, Moorthi S, Wu X et al (2014) The NCEP climate forecast system version 2. J Clim 27:2185–2208. https://doi.org/10.1175/JCLI-D-12-00823.1

Samietz J, Graf B, Höhn H et al (2007) Phenology modelling of major insect pests in fruit orchards from biological basics to decision support: the forecasting tool SOPRA*. EPPO Bull 37:255–260. https://doi.org/10.1111/j.1365-2338.2007.01121.x

Sane SP, Ramaswamy SS, Raja SV (2020) Insect architecture: structural diversity and behavioral principles. Curr Opin Insect Sci 42:39–46. https://doi.org/10.1016/j.cois.2020.08.005

Saudreau M, Pincebourde S, Dassot M et al (2013) On the canopy structure manipulation to buffer climate change effects on insect herbivore development. Trees 27:239–248. https://doi.org/10.1007/s00468-012-0791-7

Savary S, Willocquet L, Pethybridge SJ et al (2019) The global burden of pathogens and pests on major food crops. Nat Ecol Evol 3:430–439. https://doi.org/10.1038/s41559-018-0793-y

Scherrer D, Körner C (2011) Topographically controlled thermal-habitat differentiation buffers alpine plant diversity against climate warming. J Biogeogr 38:406–416. https://doi.org/10.1111/j.1365-2699.2010.02407.x

Schultze SR, Campbell MN, Walley S et al (2021) Exploration of sub-field microclimates and winter temperatures: Implications for precision agriculture. Int J Biometeorol 65:1043–1052. https://doi.org/10.1007/s00484-021-02086-0

Sears MW, Riddell EA, Rusch TW, Angilletta MJ Jr (2019) The world still is not flat: lessons learned from organismal interactions with environmental heterogeneity in terrestrial environments. Integr Comp Biol 59:1049–1058. https://doi.org/10.1093/icb/icz130

Shi P-J, Fan M-L, Reddy GVP (2017) Comparison of thermal performance equations in describing temperature-dependent developmental rates of insects: (III) phenological applications. Ann Entomol Soc Am 110:558–564. https://doi.org/10.1093/aesa/sax063

Sinclair BJ, Marshall KE, Sewell MA et al (2016) Can we predict ectotherm responses to climate change using thermal performance curves and body temperatures? Ecol Lett 19:1372–1385. https://doi.org/10.1111/ele.12686

Skocir P, Mandaric K, Kralj I, et al (2021) Analysis of Open Access Data Sources for Application in Precision Agriculture. In: 2021 16th International Conference on Telecommunications (ConTEL). IEEE, Zagreb, Croatia, pp 165–172

Smith K, Strong C, Rassoul-Agha F (2017) A new method for generating stochastic simulations of daily air temperature for use in weather generators. J Appl Meteorol Climatol 56:953–963. https://doi.org/10.1175/JAMC-D-16-0122.1

Sobek S, Rajamohan A, Dillon D et al (2011) High temperature tolerance and thermal plasticity in emerald ash borer Agrilus planipennis . Agric for Entomol 13:333–340. https://doi.org/10.1111/j.1461-9563.2011.00523.x

Srivastava V, Roe AD, Keena MA et al (2021) Oh the places they’ll go: improving species distribution modelling for invasive forest pests in an uncertain world. Biol Invasions 23:297–349. https://doi.org/10.1007/s10530-020-02372-9

Steele-Dunne SC, McNairn H, Monsivais-Huertero A et al (2017) Radar remote sensing of agricultural canopies: a review. IEEE J Sel Top Appl Earth Obs Remote Sens 10:2249–2273. https://doi.org/10.1109/JSTARS.2016.2639043

Suggitt AJ, Wilson RJ, Isaac NJB et al (2018) Extinction risk from climate change is reduced by microclimatic buffering. Nat Clim Change 8:713–717. https://doi.org/10.1038/s41558-018-0231-9

Sutherst RW (2014) Pest species distribution modelling: origins and lessons from history. Biol Invasions 16:239–256. https://doi.org/10.1007/s10530-013-0523-y

Sutherst RW, Maywald GF (1985) A computerised system for matching climates in ecology. Agric Ecosyst Environ 13:281–299. https://doi.org/10.1016/0167-8809(85)90016-7

Tanigoshi LK, Browne RW, Hoyt SC, Lagier RF (1976) Empirical analysis of variable temperature regimes on life stage development and population growth of Tetranychus mcdanieli (Acarina: Tetranychidae)1. Ann Entomol Soc Am 69:712–716. https://doi.org/10.1093/aesa/69.4.712

Taylor F (1981) Ecology and evolution of physiological time in insects. Am Nat 117:1–23. https://doi.org/10.1086/283683

Toda M, Richardson AD (2018) Estimation of plant area index and phenological transition dates from digital repeat photography and radiometric approaches in a hardwood forest in the northeastern United States. Agric For Meteorol 249:457–466. https://doi.org/10.1016/j.agrformet.2017.09.004

Tonelli M, Gomes G, Silva WD et al (2018) Spittlebugs produce foam as a thermoregulatory adaptation. Sci Rep 8:4729. https://doi.org/10.1038/s41598-018-23031-z

Tonnang HEZ, Hervé BDB, Biber-Freudenberger L et al (2017) Advances in crop insect modelling methods—towards a whole system approach. Ecol Model 354:88–103. https://doi.org/10.1016/j.ecolmodel.2017.03.015

Trew BT, Early R, Duffy JP et al (2022) Using near-ground leaf temperatures alters the projected climate change impacts on the historical range of a floristic biodiversity hotspot. Divers Distrib 28:1282–1297. https://doi.org/10.1111/ddi.13540

Trnka M, Muška F, Semerádová D et al (2007) European Corn Borer life stage model: Regional estimates of pest development and spatial distribution under present and future climate. Ecol Model 207:61–84. https://doi.org/10.1016/j.ecolmodel.2007.04.014

Uvarov BP (1931) Insects and Climate. Trans R Entomol Soc Lond 79:1–232. https://doi.org/10.1111/j.1365-2311.1931.tb00696.x

Vasseur DA, DeLong JP, Gilbert B et al (2014) Increased temperature variation poses a greater risk to species than climate warming. Proc R Soc B Biol Sci 281:20132612. https://doi.org/10.1098/rspb.2013.2612

Venette RC, Kriticos DJ, Magarey RD et al (2010) Pest risk maps for invasive alien species: a roadmap for improvement. Bioscience 60:349–362. https://doi.org/10.1525/bio.2010.60.5.5

von Schmalensee L, Hulda Gunnarsdóttir K, Näslund J et al (2021) Thermal performance under constant temperatures can accurately predict insect development times across naturally variable microclimates. Ecol Lett 24:1633–1645. https://doi.org/10.1111/ele.13779

Waggoner PE, Reifsnyder WE (1968) Simulation of the temperature, humidity and evaporation profiles in a leaf canopy. J Appl Meteorol Climatol 7:400–409. https://doi.org/10.1175/1520-0450(1968)007%3c0400:SOTTHA%3e2.0.CO;2

Wellington WG (1950) Effects of radiation on the temperatures of insectan habitats. Sci Agric 30:209–234. https://doi.org/10.4141/sa-1950-0029

Whitman DW (1987) Thermoregulation and daily activity patterns in a black desert grasshopper, Taeniopoda eques . Anim Behav 35:1814–1826. https://doi.org/10.1016/S0003-3472(87)80074-X

Wilson MJ, Digweed AJ, Brown J et al (2015) Invasive slug pests and their parasites—temperature responses and potential implications of climate change. Biol Fertil Soils 51:739–748. https://doi.org/10.1007/s00374-015-1022-3

WMO (2018) Guide to instruments and methods of observation, 2018th edn. World Meteorological Organization, Geneva

Woods HA, Dillon ME, Pincebourde S (2015) The roles of microclimatic diversity and of behavior in mediating the responses of ectotherms to climate change. J Therm Biol 54:86–97. https://doi.org/10.1016/j.jtherbio.2014.10.002

Yang C, Huang Q, Li Z et al (2017) Big Data and cloud computing: innovation opportunities and challenges. Int J Digit Earth 10:13–53. https://doi.org/10.1080/17538947.2016.1239771

Zellweger F, Frenne PD, Lenoir J et al (2019) Advances in microclimate ecology arising from remote sensing. Trends Ecol Evol 34:327–341. https://doi.org/10.1016/j.tree.2018.12.012

Zhao F, Hoffmann AA, Xing K, Ma C (2017) Life stages of an aphid living under similar thermal conditions differ in thermal performance. J Insect Physiol 99:1–7. https://doi.org/10.1016/j.jinsphys.2017.03.003

Zhu L, Hoffmann AA, Li S-M, Ma C-S (2021) Extreme climate shifts pest dominance hierarchy through thermal evolution and transgenerational plasticity. Funct Ecol 35:1524–1537. https://doi.org/10.1111/1365-2435.13774

Download references

The work was supported by funding from the Met Office Hadley Centre Climate Programme (HCCP) funded by UK government departments BEIS and Defra.

Author information

Authors and affiliations.

Environment and Sustainability Institute, Penryn Campus, University of Exeter, Exeter, TR10 9FE, UK

Jonathan R. Mosedale, Brittany Trew & Ilya M. D. Maclean

Defra, York Biotech Campus, Sand Hutton, York, YO41 1LZ, UK

Dominic Eyre, Anastasia Korycinska, Matthew Everatt & Sam Grant

Hadley Centre, Met Office, Exeter, EX1 3PB, UK

Neil Kaye & Deborah Hemming

Birmingham Institute of Forest Research (BIFOR), Birmingham University, Birmingham, B15 2TT, UK

Deborah Hemming

You can also search for this author in PubMed   Google Scholar

Contributions

JRM wrote the manuscript, with contributions from IMDM, DE, AK, SB, ME, NK and DH. The article was informed by explorative, integrated plant-pest-microclimate modelling performed by JRM and BT who also provided Fig.  1 . The opinion piece was conceived jointly by IMDM and D.H.

Corresponding author

Correspondence to Jonathan R. Mosedale .

Ethics declarations

Conflict of interest.

The authors have no relevant financial or non-financial interests to disclose.

Additional information

Communicated by Andrea Battisti .

Publisher's Note

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Mosedale, J.R., Eyre, D., Korycinska, A. et al. Mechanistic microclimate models and plant pest risk modelling. J Pest Sci (2024). https://doi.org/10.1007/s10340-024-01777-y

Download citation

Received : 24 July 2023

Revised : 29 January 2024

Accepted : 25 March 2024

Published : 10 May 2024

DOI : https://doi.org/10.1007/s10340-024-01777-y

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Microclimate
  • Mechanistic modelling
  • Thermal response
  • Find a journal
  • Publish with us
  • Track your research

IMAGES

  1. Mechanistic hypothesis supported by control experiments. a Diagram of

    mechanistic hypothesis biology

  2. Mechanistic Hypotheses

    mechanistic hypothesis biology

  3. Mechanistic hypothesis a, Enthalpic profile. b, Reaction pathway

    mechanistic hypothesis biology

  4. Mechanistic hypothesis supported by control experiments. a Diagram of

    mechanistic hypothesis biology

  5. Mechanistic hypothesis Oxidative addition of anhydride (1) is followed

    mechanistic hypothesis biology

  6. Mechanistic hypothesis. Proposed mechanism of the copper-and

    mechanistic hypothesis biology

VIDEO

  1. Constructing Hypothesis (A'level Biology)

  2. Hypothesis explain biology book

  3. #chemiosmotic_hypothesis #class11th #photosynthesis #atp_synthesis_mechanism #ncertbiology

  4. CHEMIOSMOTIC HYPOTHESIS || HINDI EXPLANATION

  5. notes on hypothesis formulation #biostatistics #msc_zoology #vbu #by_madhuri

  6. Difference Between Null Hypothesis and Alternative Hypothesis

COMMENTS

  1. When are hypotheses useful in ecology and evolution?

    Hypothesis: An explanation for an observed phenomenon. Research Hypothesis: A statement about a phenomenon that also includes the potential mechanism or cause of that phenomenon. Though a research hypothesis doesn't need to adhere to this strict framework it is often best described as the "if" in an "if-then" statement.

  2. Biodiversity and ecosystem functioning: A mechanistic model

    The relationship between biodiversity and ecosystem processes has emerged as a major scientific issue today (1-3).Recent experiments have provided evidence that loss of biodiversity may impair the functioning and sustainability of ecosystems (3-11).The interpretation of these experiments is still debated (12-16), however, and there is some experimental evidence that not all ecosystem ...

  3. Six Theses on Mechanisms and Mechanistic Science

    Given the ubiquity of references to mechanism in biology, and sparseness of reference to laws, it is a curious fact that mechanistic explanation was mostly neglected in the literature of 20th century philosophy of science. ... One mechanism, many models: A distributed theory of mechanistic explanation. Synthese 193(5): 1387-1407. Article ...

  4. The Hierarchy-of-Hypotheses Approach: A Synthesis Method for Enhancing

    Mechanistic hypothesis. Narrowed version of an overarching hypothesis, resulting from specialization or decomposition of the unspecified hypothesis with respect to assumed underlying causes. ... The escalation hypothesis is a prominent hypothesis in evolutionary biology. In response to the question why species often seem to be well adapted to ...

  5. The New Mechanistic Theory of Explanation: A Primer

    The new mechanistic theory of explanation (new mechanism, or neo-mechanism, in short) aims to describe the explanatory practices in the life sciences and provide a normative framework for evaluating purported cases of explanation resulting from these practices. The core of the new mechanistic account is the concept of a mechanism.

  6. Hierarchy-of-Hypotheses Approach: A Synthesis Method for Enhancing

    Mechanistic hypothesis. Narrowed version of an overarching hypothesis, resulting from specialization or decomposition of the unspecified hypothesis with respect to assumed underlying causes. ... The escalation hypothesis is a prominent hypothesis in evolutionary biology. In response to the question why species often seem to be well adapted to ...

  7. The art of mechanistic modeling in biology

    This study provides a fascinating combination of experiment and theory to investigate how individuals age in the wild. ... Sommer, R.J. The art of mechanistic modeling in biology. Nat Comput Sci 2 ...

  8. Mechanistic Science

    For example, in the late 1950s the problem of protein synthesis was central to biology, and "mechanism" to biochemists meant the formation of covalent bonds in polypeptides, ... This in turn may lead to novel hypotheses and predictions that can be tested, thereby completing the transition to "mechanistic" theory-driven models. The ...

  9. Mechanism, autonomy and biological explanation

    The new mechanists and the autonomy approach both aim to account for how biological phenomena are explained. One identifies appeals to how components of a mechanism are organized so that their activities produce a phenomenon. The other directs attention towards the whole organism and focuses on how it achieves self-maintenance. This paper discusses challenges each confronts and how each could ...

  10. Mechanistic approaches to the study of evolution: the functional

    The authors argue that a new approach, the functional synthesis, which combines evolutionary analyses of gene sequences with molecular biology experiments, opens new avenues to the study of the ...

  11. Mechanisms in Science

    Around the turn of the twenty-first century, what has come to be called the new mechanical philosophy (or, for brevity, the new mechanism) emerged as a framework for thinking about the philosophical assumptions underlying many areas of science, especially in sciences such as biology, neuroscience, and psychology.In this entry, we introduce and summarize the distinctive features of this ...

  12. Mechanisms in Science

    Bechtel and Abrahamsen 2005 extends this work to offer a mechanistic theory of explanation opposed to the idea that explanations require knowing the laws of nature. Glennan 1996 also deploys the concept of mechanism, but in the service of a philosophical analysis of causation. For Glennan, mechanisms are the hidden connection Hume sought ...

  13. Chapter 6 Evaluating Evidence of Mechanisms

    6.1. Overview. Evaluating evidence of mechanisms should start with clear formulations of the general mechanistic claim and each specific mechanism hypothesis, for which evidence is gathered via the procedure described in Chap. 5. The general mechanistic claim concerns either the existence of a mechanism (to account for efficacy) or the ...

  14. A mechanistic framework of enemy release

    INTRODUCTION. The enemy release hypothesis (ERH) is the most well-known hypothesis in invasion biology (Enders et al., 2018) and is frequently invoked to explain the success of exotic species in their non-native range (Connolly et al., 2014; Mitchell & Power, 2003; Torchin et al., 2003).According to the ERH, invaders can leave their natural enemies behind when introduced beyond their home ...

  15. Mechanistic Biology in the Next Quarter Century

    The potential of this new biochemical and structural cell biology for twenty-first century therapeutics—molecular, cellular, and genetic—is evident. Just as mechanistic enzymology has changed the directions of drug development during the past 25-30 years, so does mechanistic cell biology promise to redirect it during the next 25 years ...

  16. Mechanistic versus Functional Understanding

    Someone who has an intuitive theory of biology that includes resources for explaining biological adaptations might possess some "functional understanding" of biological adaptations in general. When applying this to the spotted fawn, she comes to possess some functional understanding of why fawns have spots.

  17. The principle of uncertainty in biology: Will machine learning

    Using a systems biology approach, motifs, patterns, and correlations can be identified that in turn raise questions and testable hypothesis. However, the scientific agenda up to this point remained mechanistic comprehension of biological phenomena from first principles.

  18. The concept of mechanism in biology

    The concept of mechanism in biology has three distinct meanings. It may refer to a philosophical thesis about the nature of life and biology ('mechanicism'), to the internal workings of a machine-like structure ('machine mechanism'), or to the causal explanation of a particular phenomenon ('causal m …

  19. Using Diagrams to Reason About Biological Mechanisms

    A common way to identify such lacunae is to compare one mechanistic hypothesis to others that are, in at least one respect, better worked out. In their search for mechanisms, biologists often work comparatively between species. ... similar diagram practices can be found in other areas of molecular biology and biology more generally, along with ...

  20. Mechanism (biology)

    In the science of biology, a mechanism is a system of causally interacting parts and processes that produce one or more effects. Scientists explain phenomena by describing mechanisms that could produce the phenomena. For example, natural selection is a mechanism of biological evolution; other mechanisms of evolution include genetic drift, mutation, and gene flow.

  21. Mechanistic studies, explanation and examples

    EME encourages hypothesis-testing mechanistic studies. These studies can explore the mechanisms of action of the intervention, the causes of differing responses, or promote an understanding of any potential adverse effects and how these could be reduced. They could also contribute to understanding of the disease process.

  22. MECHANISTIC THEORY

    Current State of Research. Modern breakthroughs in mechanistic theory have progressed beyond simple mechanical analogies, incorporating insights from domains as diverse as systems biology, cognitive science, and artificial intelligence research.:. Modern developments in mechanistic theory incorporate insights from various fields.; The current focus is on understanding complex, dynamic systems ...

  23. The Origin of Life: Concept, Approaches and Theories

    By the turn of this century several main approaches have been made on the origin of life. These approaches are the vitalism, special creation, panspermia, mechanistic theory and materialism. 1. Vitalism: This concept attributes the distinctive properties of the living things to a supernatural life force. 2.

  24. mechanistic hypotheses

    Their conclusion is dismal if you think ecological science should make progress in gathering evidence. No change from 1990 to 2015. Multiple alternative hypotheses = 6% of papers, Mechanistic hypotheses = 25% of papers, Descriptive hypotheses = 12%, No hypotheses = 75% of papers. Why should this be after years of recommending the gold standard ...

  25. Biological Mechanism

    One mechanism might be the body's biological response to stress, involving the autonomic nervous system, the hypothalamic-pituitary-adrenal (HPA) axis, and the cardiovascular, metabolic, and immune systems. Many studies have shown the cumulative impact of risk and protective factors on young children's development.

  26. Mechanistic microclimate models and plant pest risk modelling

    Theory and methods were subsequently developed in the agricultural and forest sciences and have been well-established since the 1950s (Penman 1948; ... With an adequate knowledge of pest biology, we can adapt mechanistic microclimate models to capture many of these effects.

  27. Functional coexistence theory: a mechanistic framework ...

    Theory and experiments show that diverse ecosystems often have higher levels of function (for instance, biomass production), yet it remains challenging to identify the biological mechanisms responsible. We synthesize developments in coexistence theory into a general theoretical framework linking community coexistence to ecosystem function. Our framework, which we term functional coexistence ...