Adaptation

In biology, adaptation has three related meanings. Firstly it is the dynamic evolutionary process that fits a population of organisms to their environment, enhancing their evolutionary fitness. Secondly, it is a state reached by the population during that process. Thirdly, it is a phenotypic or adaptive trait, with a functional role in each individual organism, that is maintained and has been evolved by natural selection.

Organisms face a succession of environmental challenges as they grow, and show adaptive plasticity as traits develop in response to the imposed conditions. This gives them resilience to varying environments.

History

Adaptation is a fact of life that has been accepted by many of the great thinkers who have tackled the world of living organisms. It is their explanations of how adaptation arises that separates these thinkers. Empedocles did not believe that adaptation required a final cause (~ purpose), but "came about naturally, since such things survived." Aristotle did believe in final causes or adaptation, but assumed that species were fixed.

In natural theology, adaptation was interpreted as the work of a deity, even as evidence for the existence of God.[1] William Paley believed that organisms were perfectly adapted to the lives they lead, an argument that shadowed Gottfried Wilhelm Leibniz, who had argued that God had brought about "the best of all possible worlds." Voltaire's Dr. Pangloss[2] is a parody of this optimistic idea, and David Hume also argued against design.[3] The Bridgewater Treatises are a product of natural theology, though some of the authors managed to present their work in a fairly neutral manner. The series was lampooned by Robert Knox, who held quasi-evolutionary views, as the Bilgewater Treatises. Charles Darwin broke with the tradition by emphasising the flaws and limitations which occurred in the animal and plant worlds.[4]

Lamarckism is a proto-evolutionary hypothesis of the inheritance of acquired characteristics, whose main purpose is to explain adaptations by natural means.[5] Jean-Baptiste Lamarck proposed a tendency for organisms to become more complex, moving up a ladder of progress, plus "the influence of circumstances," usually expressed as use and disuse.[6]

Many other students of natural history, such as Buffon, accepted adaptation, and some also accepted evolution, without voicing their opinions as to the mechanism. This illustrates the real merit of Darwin and Alfred Russel Wallace, and secondary figures such as Henry Walter Bates, for putting forward a mechanism whose significance had only been glimpsed previously. A century later, experimental field studies and breeding experiments by people such as E. B. Ford and Dobzhansky produced evidence that natural selection was not only the 'engine' behind adaptation, but was a much stronger force than had previously been thought.[7][8][9]

General principles

The significance of an adaptation can only be understood in relation to the total biology of the species.

What adaptation is

Adaptation is, first of all, a process, to rather be with the animal and rather than a part of a body.[11] An internal parasite (such as a liver fluke) can illustrate the distinction: such a parasite may have a very simple bodily structure, but nevertheless the organism is highly adapted to its specific environment. From this we see that adaptation is not just a matter of visible traits: in such parasites critical adaptations take place in the life cycle, which is often quite complex.[12] However, as a practical term, "adaptation" often refers to a product: those features of a species which result from the process. Many aspects of an animal or plant can be correctly called adaptations, though there are always some features whose function remains in doubt. By using the term adaptation for the evolutionary process, and adaptive trait for the bodily part or function (the product), one may distinguish the two different senses of the word.[13][14][15][16]

Adaptation is one of the two main processes that explain the observed diversity of species, such as the different species of Darwin's finches. The other process is speciation, in which new species arise, typically through reproductive isolation.[17][18] A favorite example used today to study the interplay of adaptation and speciation is the evolution of cichlid fish in African lakes, where the question of reproductive isolation is complex.[19][20]

Adaptation is not always a simple matter where the ideal phenotype evolves for a given external environment. An organism must be viable at all stages of its development and at all stages of its evolution. This places constraints on the evolution of development, behavior and structure of organisms. The main constraint, over which there has been much debate, is the requirement that each genetic and phenotypic change during evolution should be relatively small, because developmental systems are so complex and interlinked. However, it is not clear what "relatively small" should mean, for example polyploidy in plants is a reasonably common large genetic change.[21] The origin of eukaryotic symbiosis exemplifies a more dramatic example.[22]

All adaptations help organisms survive in their ecological niches.[23] The adaptive traits may be structural, behavioral or physiological. Structural adaptations are physical features of an organism (shape, body covering, armament, internal organization). Behavioral adaptations are inherited behavior chains and the ability to learn: behaviors may be inherited in detail (instincts), or a capacity for learning may be inherited (see neuropsychology). Examples: searching for food, mating, vocalizations. Physiological adaptations can permit the organism to perform special functions (for instance, making venom, secreting slime, phototropism), but may also involve more general functions such as growth and development, temperature regulation, ionic balance and other aspects of homeostasis. Adaptation affects all aspects of the life of an organism.

The following definitions are given by the evolutionary biologist Theodosius Dobzhansky:

1. Adaptation is the evolutionary process whereby an organism becomes better able to live in its habitat or habitats.[24]
2. Adaptedness is the state of being adapted: the degree to which an organism is able to live and reproduce in a given set of habitats.[25]
3. An adaptive trait is an aspect of the developmental pattern of the organism which enables or enhances the probability of that organism surviving and reproducing.[26]

What adaptation is not

Some generalists, such as birds, have the flexibility to adapt to urban areas.

Adaptation differs from flexibility, acclimatization, and learning. Flexibility deals with the relative capacity of an organism to maintain itself in different habitats: its degree of specialization. Acclimatization describes automatic physiological adjustments during life; learning means improvement in behavioral performance during life. These terms are preferred to adaptation for changes during life which are not inherited by the next generation.

Flexibility stems from phenotypic plasticity, the ability of an organism with a given genotype to change its phenotype in response to changes in its habitat, or to move to a different habitat.[27][28] The degree of flexibility is inherited, and varies between individuals. A highly specialized animal or plant lives only in a well-defined habitat, eats a specific type of food, and cannot survive if its needs are not met. Many herbivores are like this; extreme examples are koalas which depend on Eucalyptus, and giant pandas which require bamboo. A generalist, on the other hand, eats a range of food, and can survive in many different conditions. Examples are humans, rats, crabs and many carnivores. The tendency to behave in a specialized or exploratory manner is inherited—it is an adaptation. Rather different is developmental flexibility: "An animal or plant is developmentally flexible if when it is raised in or transferred to new conditions, it changes in structure so that it is better fitted to survive in the new environment," writes evolutionary biologist John Maynard Smith.[29]

If humans move to a higher altitude, respiration and physical exertion become a problem, but after spending time in high altitude conditions they acclimatize to the reduced partial pressure of oxygen, such as by producing more red blood cells. The ability to acclimatize is an adaptation, but the acclimatization itself is not. Fecundity goes down, but deaths from some tropical diseases also goes down. Over a longer period of time, some people are better able to reproduce at high altitudes than others. They contribute more heavily to later generations, and gradually by natural selection the whole population becomes adapted to the new conditions. This has demonstrably occurred, as the observed performance of long-term communities at higher altitude is significantly better than the performance of new arrivals, even when the new arrivals have had time to acclimatize.[30]

Adaptedness and fitness

In this sketch of a fitness landscape, a population can evolve by following the arrows to the adaptive peak at point B, and the points A and C are local optima where a population could become trapped.

There is a relationship between adaptedness and the concept of fitness used in population genetics. Differences in fitness between genotypes predict the rate of evolution by natural selection. Natural selection changes the relative frequencies of alternative phenotypes, insofar as they are heritable.[31] However, a phenotype with high adaptedness may not have high fitness. Dobzhansky mentioned the example of the Californian redwood, which is highly adapted, but a relict species in danger of extinction.[24] Elliott Sober commented that adaptation was a retrospective concept since it implied something about the history of a trait, whereas fitness predicts a trait's future.[32]

1. Relative fitness. The average contribution to the next generation by a genotype or a class of genotypes, relative to the contributions of other genotypes in the population.[33] This is also known as Darwinian fitness, selection coefficient, and other terms.
2. Absolute fitness. The absolute contribution to the next generation by a genotype or a class of genotypes. Also known as the Malthusian parameter when applied to the population as a whole.[31][34]
3. Adaptedness. The extent to which a phenotype fits its local ecological niche. Researchers can sometimes test this through a reciprocal transplant.

Sewall Wright proposed that populations occupy adaptive peaks on a fitness landscape. In order to evolve to another, higher peak, a population would first have to pass through a valley of maladaptive intermediate stages.[35] A given population might be "trapped" on a peak that is not optimally adapted.

Genetic basis

A large diversity of genome DNAs in a species is the basis for adaptation and differentiation. A large population is needed to carry sufficient diversity. According to the misrepair-accumulation aging theory,[36][37] The misrepair mechanism is important in maintaining a sufficient number of individuals in a species.[38] misrepair is a way of repair for increasing the surviving chance of an organism when it has severe injuries. Without misrepairs, no individual could survive to reproduction age. Thus misrepair mechanism is an essential mechanism for the survival of a species and for maintaining the number of individuals. Although individuals die from aging, genome DNAs are being recopied and transmitted by individuals generation by generation. In addition, the DNA misrepairs in germ cells contribute also to the diversity of genome DNAs.

Types

Adaptation is the heart and soul of evolution.
Niles Eldredge, Reinventing Darwin: The Great Debate at the High Table of Evolutionary Theory[39]

Changes in habitat

Before Darwin, adaptation was seen as a fixed relationship between an organism and its habitat. It was not appreciated that as the climate changed, so did the habitat; and as the habitat changed, so did the biota. Also, habitats are subject to changes in their biota: for example, invasions of species from other areas. The relative numbers of species in a given habitat are always changing. Change is the rule, though much depends on the speed and degree of the change. When the habitat changes, three main things may happen to a resident population: habitat tracking, genetic change or extinction. In fact, all three things may occur in sequence. Of these three effects only genetic change brings about adaptation. When a habitat changes, the resident population typically moves to more suitable places; this is the typical response of flying insects or oceanic organisms, which have wide (though not unlimited) opportunity for movement.[40] This common response is called habitat tracking. It is one explanation put forward for the periods of apparent stasis in the fossil record (the punctuated equilibrium theory).[41]

Genetic change

Genetic change occurs in a population when natural selection and mutations act on its genetic variability.[42] The first pathways of enzyme-based metabolism may have been parts of purine nucleotide metabolism, with previous metabolic pathways being part of the ancient RNA world. By this means, the population adapts genetically to its circumstances.[9] Genetic changes may result in visible structures, or may adjust physiological activity in a way that suits the habitat.

Habitats and biota do frequently change. Therefore, it follows that the process of adaptation is never finally complete.[43] Over time, it may happen that the environment changes little, and the species comes to fit its surroundings better and better. On the other hand, it may happen that changes in the environment occur relatively rapidly, and then the species becomes less and less well adapted. Seen like this, adaptation is a genetic tracking process, which goes on all the time to some extent, but especially when the population cannot or does not move to another, less hostile area. Given enough genetic change, as well as specific demographic conditions, an adaptation may be enough to bring a population back from the brink of extinction in a process called evolutionary rescue. It should be noted that adaptation does affect, to some extent, every species in a particular ecosystem.[44][45]

Leigh Van Valen thought that even in a stable environment, competing species had to constantly adapt to maintain their relative standing. This became known as the Red Queen hypothesis, as seen in host-parasite interaction.[46]

Co-adaptation

Pollinating insects are co-adapted with flowering plants.

In coevolution, where the existence of one species is tightly bound up with the life of another species, new or 'improved' adaptations which occur in one species are often followed by the appearance and spread of corresponding features in the other species. These relationships are intrinsically dynamic, and may continue on a trajectory for millions of years, as has occurred in the relationship between flowering plants and pollinating insects.

Mimicry

A and B show real wasps; the rest are Batesian mimics: three hoverflies and one beetle.

Bates' work on Amazonian butterflies led him to develop the first scientific account of mimicry, especially the kind of mimicry which bears his name: Batesian mimicry.[47] This is the mimicry by a palatable species of an unpalatable or noxious species. A common example seen in temperate gardens is the hoverfly, many of which—though bearing no sting—mimic the warning colouration of hymenoptera (wasps and bees). Such mimicry does not need to be perfect to improve the survival of the palatable species.[48]

Bates, Wallace and Fritz Müller believed that Batesian and Müllerian mimicry provided evidence for the action of natural selection, a view which is now standard amongst biologists.[49] All aspects of this situation can be, and have been, the subject of research.[50] Field and experimental work on these ideas continues to this day; the topic connects strongly to speciation, genetics and evolutionary developmental biology.[51]

Internal adaptations

There are some important adaptations to do with the overall coordination of the systems in the body. Such adaptations may have significant consequences. Examples, in vertebrates, would be temperature regulation, or improvements in brain function, or an effective immune system. Contains an extensive analysis of the evolution of adaptations in the radiation of angiosperms. Such adaptations may make the clade (monophyletic group) more viable in a wide range of habitats. The acquisition of such major adaptations has often served as the spark for adaptive radiation, and huge success over long periods of time for a whole group of animals or plants.

Trade-offs

It is a profound truth that Nature does not know best; that genetical evolution... is a story of waste, makeshift, compromise and blunder.
Peter Medawar, The Future of Man[52]

All adaptations have a downside: horse legs are great for running on grass, but they can't scratch their backs; mammals' hair helps temperature, but offers a niche for ectoparasites; the only flying penguins do is under water. Adaptations serving different functions may be mutually destructive. Compromise and makeshift occur widely, not perfection. Selection pressures pull in different directions, and the adaptation that results is some kind of compromise.[53]

Since the phenotype as a whole is the target of selection, it is impossible to improve simultaneously all aspects of the phenotype to the same degree.

Consider the antlers of the Irish elk, (often supposed to be far too large; in deer antler size has an allometric relationship to body size). Obviously, antlers serve positively for defence against predators, and to score victories in the annual rut. But they are costly in terms of resource. Their size during the last glacial period presumably depended on the relative gain and loss of reproductive capacity in the population of elks during that time.[55] Another example: camouflage to avoid detection is destroyed when vivid colors are displayed at mating time. Here the risk to life is counterbalanced by the necessity for reproduction.

Stream-dwelling salamanders, such as Caucasian salamander or Gold-striped salamander have very slender, long bodies, perfectly adapted to life at the banks of fast small rivers and mountain brooks. Elongated body protects their larvae from being washed out by current. However, elongated body increases risk of desiccation and decreases dispersal ability of the salamanders; it also negatively affects their fecundity. As a result, fire salamander, less perfectly adapted to the mountain brook habitats, is in general more successful, have a higher fecundity and broader geographic range.[56]

An Indian peacock's train
in full display

The peacock's ornamental train (grown anew in time for each mating season) is a famous adaptation. It must reduce his maneuverability and flight, and is hugely conspicuous; also, its growth costs food resources. Darwin's explanation of its advantage was in terms of sexual selection: "This depends on the advantage which certain individuals have over other individuals of the same sex and species, in exclusive relation to reproduction."[57] The kind of sexual selection represented by the peacock is called 'mate choice,' with an implication that the process selects the more fit over the less fit, and so has survival value.[58] The recognition of sexual selection was for a long time in abeyance, but has been rehabilitated.[59]

The conflict between the size of the human foetal brain at birth, (which cannot be larger than about 400 cm3, else it will not get through the mother's pelvis) and the size needed for an adult brain (about 1400 cm3), means the brain of a newborn child is quite immature. The most vital things in human life (locomotion, speech) just have to wait while the brain grows and matures. That is the result of the birth compromise. Much of the problem comes from our upright bipedal stance, without which our pelvis could be shaped more suitably for birth. Neanderthals had a similar problem.[60][61][62]

As another example, the long neck of a giraffe is a burden and a blessing. The neck of a giraffe can be up to 2 m (6 ft 7 in) in length.[63] This neck can be used for inter-species competition or for foraging on tall trees where shorter herbivores cannot reach. However, as previously stated, there is always a trade-off. This long neck is heavy and it adds to the body mass of a giraffe, so the giraffe needs an abundance of nutrition to provide for this costly adaptation.[64]

Shifts in function

Adaptation and function are two aspects of one problem.
Julian Huxley, Evolution: The Modern Synthesis[65]

Pre-adaptations

This occurs when a species or population has characteristics which (by chance) are suited for conditions which have not yet arisen. For example, the polyploid cordgrass Spartina townsendii is better adapted than either of its parent species to their own habitat of saline marsh and mud-flats.[66] White Leghorn chicken are markedly more resistant to vitamin B1 deficiency than other breeds.[67] On a plentiful diet there is no difference, but on a restricted diet this preadaptation could be decisive.

Pre-adaptation may occur because a natural population carries a huge quantity of genetic variability.[68] In diploid eukaryotes, this is a consequence of the system of sexual reproduction, where mutant alleles get partially shielded, for example, by the selective advantage of heterozygotes. Microorganisms, with their huge populations, also carry a great deal of genetic variability.

The first experimental evidence of the pre-adaptive nature of genetic variants in microorganisms was provided by Salvador Luria and Max Delbrück who developed Fluctuation Test, a method to show the random fluctuation of pre-existing genetic changes that conferred resistance to bacteriophage in the bacterium Escherichia coli.

Co-option of existing traits: exaptation

Sinosauropteryx fossil, a dinosaur with feathers, used for insulation, not flight

The classic example is the ear ossicles of mammals, which we know from paleontological and embryological studies originated in the upper and lower jaws and the hyoid bone of their synapsid ancestors, and further back still were part of the gill arches of early fish.[69] We owe this esoteric knowledge to the comparative anatomists, who, a century ago, were at the cutting edge of evolutionary studies.[70] The word exaptation was coined to cover these shifts in function, which are surprisingly common in evolutionary history.[71] The origin of wings from feathers that were originally used for temperature regulation is a more recent discovery (see feathered dinosaurs).

Non-adaptive traits

Some traits do not appear to be adaptive, that is, they have a neutral or deleterious effect on fitness in the current environment. Because genes have pleiotropic effects, not all traits may be functional: they may be what Stephen Jay Gould and Richard Lewontin called spandrels, features brought about by neighbouring adaptations, like the triangular areas under neighbouring arches in architecture which began as functionless features.[72]

Another possibility is that a trait may have been adaptive at some point in an organism's evolutionary history, but a change in habitats caused what used to be an adaptation to become unnecessary or even a hindrance (maladaptations). Such adaptations are termed vestigial. Many organisms have vestigial organs, which are the remnants of fully functional structures in their ancestors. As a result of changes in lifestyle the organs became redundant, and are either not functional or reduced in functionality. Since any structure represents some kind of cost to the general economy of the body, an advantage may accrue from their elimination once they are not functional. Examples: wisdom teeth in humans; the loss of pigment and functional eyes in cave fauna; the loss of structure in endoparasites.[73]

Extinction and coextinction

If a population cannot move or change sufficiently to preserve its long-term viability, then obviously, it will become extinct, at least in that locale. The species may or may not survive in other locales. Species extinction occurs when the death rate over the entire species exceeds the birth rate for a long enough period for the species to disappear. It was an observation of Van Valen that groups of species tend to have a characteristic and fairly regular rate of extinction.[74]

Just as we have co-adaptation, there is also coextinction.[75] Coextinction is the loss of a species due to the extinction of another with which it is coadapted, as with the extinction of parasitic insects following the loss of their hosts, or when a flowering plant loses its pollinator, or when a food chain is disrupted.[76]

Philosophical issues

Adaptation raises philosophical issues concerning how biologists speak of function and purpose, as this carries implications of evolutionary history – that a feature evolved by natural selection for a specific reason – and potentially of supernatural intervention – that features and organisms exist because of a deity's conscious intentions.[77][78] Aristotle introduced teleology to describe the adaptedness of organisms, but without accepting the supernatural intention built into Plato's thinking, which Aristotle rejected.[79][80] Modern biologists continue to face the same difficulty.[81][82][83][84][85][86] On the one hand, adaptation is obviously purposeful: natural selection chooses what works and eliminates what does not. On the other hand, biologists want to deny conscious purpose in evolution. The dilemma gave rise to a famous joke by the evolutionary biologist Haldane: "Teleology is like a mistress to a biologist: he cannot live without her but he's unwilling to be seen with her in public.'" David Hull commented that Haldane's mistress "has become a lawfully wedded wife. Biologists no longer feel obligated to apologize for their use of teleological language; they flaunt it."[87]

See also

References

  1. Desmond 1989, pp. 31–32, fn 18
  2. Voltaire (1759). Candide. Cramer et al.
  3. Sober 1993, chpt. 2
  4. Darwin 1872, p. 397: "Rudimentary, Atrophied, and Aborted Organs"
  5. See, for example, the discussion in Bowler 2003, pp. 86–95: "Whatever the true nature of Lamark's theory, it was his mechanism of adaptation that caught the attention of later naturalists." (p. 90)
  6. Bowler, Peter J. (1989) [1983]. Evolution The History of an Idea (Revised ed.). University of California Press. p. 86. ISBN 0-520-06386-4.
  7. Provine 1986
  8. Ford 1975
  9. 1 2 Orr, H. Allen (February 2005). "The genetic theory of adaptation: a brief history". Nature Reviews Genetics. London: Nature Publishing Group. 6 (2): 119–127. ISSN 1471-0056. PMID 15716908. doi:10.1038/nrg1523.
  10. Huxley 1942, p. 449
  11. Mayr 1982, p. 483: "Adaptation... could no longer be considered a static condition, a product of a creative past, and became instead a continuing dynamic process."
  12. Price 1980
  13. Daintith, John; Martin, Elizabeth A., eds. (2010) [First published 1984 as Concise Science Dictionary]. "adaptation". A Dictionary of Science. Oxford Paperback Reference (6th ed.). Oxford; New York: Oxford University Press. p. 13. ISBN 978-0-19-956146-9. LCCN 2010287468. OCLC 444383696. Any change in the structure or functioning of successive generations of a population that makes it better suited to its environment.
  14. Bowler 2003, p. 10
  15. Patterson 1999, p. 1
  16. Williams 1966, p. 5: "Evolutionary adaptation is a phenomenon of pervasive importance in biology."
  17. Mayr 1963
  18. Mayr 1982, pp. 562–566
  19. Salzburger, Walter; Mack, Tanja; Verheyen, Erik; Meyer, Axel (February 21, 2005). "Out of Tanganyika: Genesis, explosive speciation, key-innovations and phylogeography of the haplochromine cichlid fishes" (PDF). BMC Evolutionary Biology. London: BioMed Central. 5 (17). ISSN 1471-2148. PMC 554777Freely accessible. PMID 15723698. doi:10.1186/1471-2148-5-17. Retrieved 2015-08-15.
  20. Kornfield, Irv; Smith, Peter F. (November 2000). "African Cichlid Fishes: Model Systems for Evolutionary Biology". Annual Review of Ecology and Systematics. Palo Alto, CA: Annual Reviews. 31: 163–196. ISSN 1545-2069. doi:10.1146/annurev.ecolsys.31.1.163.
  21. Stebbins 1950, chpts. 8 and 9
  22. Margulis & Fester 1991
  23. Hutchinson 1965. The niche is the central concept in evolutionary ecology; see especially part II: "The niche: an abstractly inhabited hypervolume." (pp. 26–78)
  24. 1 2 Dobzhansky 1968, pp. 1–34
  25. Dobzhansky 1970, pp. 4–6; 79–82
  26. Dobzhansky, Theodosius (March 1956). "Genetics of Natural Populations. XXV. Genetic Changes in Populations of Drosophila pseudoobscura and Drosophila persimilis in Some Localities in California". Evolution. Hoboken, NJ: John Wiley & Sons for the Society for the Study of Evolution. 10 (1): 82–92. JSTOR 2406099. doi:10.2307/2406099.
  27. Price, Trevor D.; Qvarnström, Anna; Irwin, Darren E. (July 2003). "The role of phenotypic plasticity in driving genetic evolution". Proceedings of the Royal Society B. London: Royal Society. 270 (1523): 1433–1440. ISSN 0962-8452. PMC 1691402Freely accessible. PMID 12965006. doi:10.1098/rspb.2003.2372.
  28. Price, Trevor D. (June 2006). "Phenotypic plasticity, sexual selection and the evolution of colour patterns". The Journal of Experimental Biology. Cambridge, UK: The Company of Biologists. 209 (12): 2368–2376. ISSN 0022-0949. PMID 16731813. doi:10.1242/jeb.02183.
  29. Maynard Smith 1993, p. 33
  30. Moore, Lorna G.; Regensteiner, Judith G. (October 1983). "Adaptation to High Altitude". Annual Review of Anthropology. Palo Alto, CA: Annual Reviews. 12: 285–304. doi:10.1146/annurev.an.12.100183.001441.
  31. 1 2 Endler 1986, pp. 33–51
  32. Sober 1984, p. 210
  33. Futuyma 1986, p. 552
  34. Fisher 1930, p. 25
  35. Wright 1932, pp. 356–366
  36. Wang, Jicun; Michelitsch, Thomas M.; Wunderlin, Arne; Mahadeva, Ravi (2009). "Aging as a consequence of misrepair—A novel theory of aging". arXiv:0904.0575Freely accessible [q-bio.TO].
  37. Wang-Michelitsch, Jicun; Michelitsch, Thomas M. (2015). "Aging as a process of accumulation of misrepairs". arXiv:1503.07163Freely accessible [q-bio.TO].
  38. Wang-Michelitsch, Jicun; Michelitsch, Thomas M. (2015). "Misrepair mechanism: a mechanism essential for individual adaptation, species adaptation and species evolution". arXiv:1505.03900Freely accessible [q-bio.TO].
  39. Eldredge 1995, p. 33
  40. Eldredge 1985, p. 136: "Of glaciers and beetles"
  41. Eldredge 1995, p. 64
  42. Hogan, C. Michael (October 12, 2010). "Mutation". In Monosson, Emily. Encyclopedia of Earth. Washington, D.C.: Environmental Information Coalition, National Council for Science and the Environment. OCLC 72808636. Retrieved 2015-08-18.
  43. Mayr 1982, pp. 481–483: This sequence tells how Darwin's ideas on adaptation developed as he came to appreciate it as "a continuing dynamic process."
  44. Sterelny & Griffiths 1999, p. 217
  45. Freeman & Herron 2007, p. 364
  46. Rabajante, J; et al. (2016). "Host-parasite Red Queen dynamics with phase-locked rare genotypes". Science Advances. 2: e1501548. doi:10.1126/sciadv.1501548.
  47. Carpenter & Ford 1933
  48. Wickler 1968
  49. Moon 1976
  50. Ruxton, Sherratt & Speed 2004
  51. Mallet, James (November 2001). "The speciation revolution" (PDF). Journal of Evolutionary Biology. Hoboken, NJ: Wiley-Blackwell on behalf of the European Society for Evolutionary Biology. 14 (6): 887–888. ISSN 1010-061X. doi:10.1046/j.1420-9101.2001.00342.x.
  52. Medawar 1960
  53. Jacob, François (June 10, 1977). "Evolution and Tinkering". Science. Washington, D.C.: American Association for the Advancement of Science. 196 (4295): 1161–1166. ISSN 0036-8075. PMID 860134. doi:10.1126/science.860134.
  54. Mayr 1982, p. 589
  55. Gould, Stephen Jay (June 1974). "The Origin and Function of 'Bizarre' Structures: Antler Size and Skull Size in the 'Irish Elk,' Megaloceros giganteus". Evolution. Hoboken, NJ: John Wiley & Sons for the Society for the Study of Evolution. 28 (2): 191–220. ISSN 0014-3820. JSTOR 2407322. doi:10.2307/2407322.
  56. Tarkhnishvili, David N. (1994). "Interdependences between Populational, Developmental and Morphological Features of the Caucasian salamander, Mertensiella caucasica" (PDF). Mertensiella. Bonn, Germany: Deutsche Gesellschaft für Herpetologie und Terrarienkunde. 4: 315–325. ISSN 0934-6643. Retrieved 2015-08-18.
  57. Darwin 1871, p. 256
  58. The case was treated by Fisher 1930, pp. 134–139
  59. Cronin 1991
  60. Rosenberg, Karen R. (1992). "The evolution of modern human childbirth". American Journal of Physical Anthropology. Hoboken, NJ: John Wiley & Sons for the American Association of Physical Anthropologists. 35 (Supplement S15): 89–124. ISSN 0002-9483. doi:10.1002/ajpa.1330350605.
  61. Friedlander, Nancy J.; Jordan, David K. (October–December 1994). "Obstetric implications of Neanderthal robusticity and bone density". Human Evolution. Kluwer Academic Publishers. 9 (4): 331–342. ISSN 0393-9375. doi:10.1007/BF02435519.
  62. Miller 2007
  63. Williams 2010, p. 29
  64. Altwegg, Robert E.; Simmons, Res (September 2010). "Necks-for-sex or competing browsers? A critique of ideas on the evolution of giraffe". Journal of Zoology. Hoboken, NJ: Wiley-Blackwell. 282 (1): 6–12. ISSN 0952-8369. doi:10.1111/j.1469-7998.2010.00711.x.
  65. Huxley 1942, p. 417
  66. Huskins, C. Leonard (1930). "The origin of Spartina Townsendii". Genetica. Martinus Nijhoff, The Hague/Kluwer Academic Publishers. 12 (6): 531–538. ISSN 0016-6707. doi:10.1007/BF01487665.
  67. Lamoreux, Wilfred F.; Hutt, Frederick B. (February 15, 1939). "Breed differences in resistance to a deficiency in vitamin B1 in the fowl" (PDF). Journal of Agricultural Research. Washington, D.C.: United States Department of Agriculture. 58 (4): 307–316. ISSN 0095-9758. Retrieved 2015-08-20.
  68. Dobzhansky 1981
  69. Allin & Hopson 1992, pp. 587–614
  70. Panchen 1992, chpt. 4, "Homology and the evidence for evolution"
  71. Gould, Stephen Jay; Vrba, Elizabeth S. (Winter 1982). "Exaptation–A Missing Term in the Science of Form". Paleobiology. Boulder, CO: Paleontological Society. 8 (1): 4–15. ISSN 0094-8373. JSTOR 2400563.
  72. Wagner, Günter P., Homology, Genes, and Evolutionary Innovation. Princeton University Press. 2014. Chapter 1: The Intellectual Challenge of Morphological Evolution: A Case for Variational Structuralism. Page 7
  73. Barrett et al. 1987. Charles Darwin was the first to put forward such ideas.
  74. Van Valen, Leigh (July 1973). "A New Evolutionary Law" (PDF). Evolutionary Theory. Chicago, IL: University of Chicago. 1: 1–30. ISSN 0093-4755. Retrieved 2015-08-22.
  75. Koh, Lian Pin; Dunn, Robert R.; Sodhi, Navjot S.; et al. (September 2004). "Species Coextinctions and the Biodiversity Crisis". Science. Washington, D.C.: American Association for the Advancement of Science. 305 (5690): 1632–1634. ISSN 0036-8075. PMID 15361627. doi:10.1126/science.1101101.
  76. Darwin 1872, pp. 57–58. Darwin in tells the story of "a web of complex relations" involving heartsease (Viola tricolor), red clover (Trifolium pratense), humble-bees (bumblebees), mice and cats.
  77. Sober 1993, pp. 85–86
  78. Williams 1966, pp. 8–10
  79. Nagel, Ernest (May 1977). "Goal-Directed Processes in Biology". The Journal of Philosophy. New York: The Journal of Philosophy, Inc. 74 (5): 261–279. ISSN 0022-362X. JSTOR 2025745. doi:10.2307/2025745. Teleology Revisisted: The Dewy Lectures 1977 (first lecture)
  80. Nagel, Ernest (May 1977). "Functional Explanations in Biology". The Journal of Philosophy. 74 (5): 280–301. ISSN 0022-362X. JSTOR 2025746. doi:10.2307/2025746. Teleology Revisisted: The Dewy Lectures 1977 (second lecture)
  81. Pittendrigh 1958
  82. Mayr 1965, pp. 33–50
  83. Mayr 1988, chpt. 3, "The Multiple Meanings of Teleological"
  84. Williams 1966, "The Scientific Study of Adaptation"
  85. Monod 1971
  86. Allaby, Michael, ed. (2003). "teleonomy". A Dictionary of Zoology. Oxford Paperback Reference (Reissued with new cover and corrections ed.). Oxford; New York: Oxford University Press. ISBN 0-19-860758-X. LCCN 2003278285. OCLC 444678726. Retrieved 2015-08-24.
  87. Hull 1982

Sources

This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.