Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-p2v8j Total loading time: 0.001 Render date: 2024-06-03T02:56:48.167Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  07 August 2009

Raphael Falk
Affiliation:
Hebrew University of Jerusalem
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Genetic Analysis
A History of Genetic Thinking
, pp. 293 - 320
Publisher: Cambridge University Press
Print publication year: 2009

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Allen, G. E. (1966). Thomas Hunt Morgan and the problem of sex determination, 1903–1910. Proceedings of the American Philosophical Society, 110, 48–57.Google Scholar
Allen, G. E. (1975 [1978]). Life Science in the Twentieth Century. Cambridge:Cambridge University Press.Google Scholar
Allen, G. E. (1978). Thomas Hunt Morgan: The Man and His Science. Princeton, NJ:Princeton University Press.Google Scholar
Allen, G. E. (1979). Naturalists and experimentalists: The genotype and the phenotype. In Coleman, W. and Limoges, C. (eds.), Studies in the History of Biology (Vol. III, pp. 179–209). Baltimore, MD: Johns Hopkins University Press.Google Scholar
Allen, G. E. (1984). Thomas Hunt Morgan: Materialism and experimentalism in the development of modern genetics. Social Research, 51(3), 709–738.Google Scholar
Altenburg, E. (1933). The production of mutations by ultra-violet light. Science, 78, 587.CrossRefGoogle ScholarPubMed
Altenburg, E., and Muller, H. J. (1920). The genetic basis of truncate wing – an inconstant and modifiable character in Drosophila. Genetics, 5, 1–59.Google Scholar
Amundson, R. (2006). EvoDevo as cognitive psychology. Biological Theory, 1(1), 10–11.CrossRefGoogle Scholar
Anfinsen, C. B., and Redfield, R. R. (1956). Protein structure in relation to function and biosynthesis. Advances in Protein Chemistry, 11, 1–100.CrossRefGoogle ScholarPubMed
Ankeny, R. A. (2000). Marvelling at the marvel: The supposed conversion of A. D. Darbishire to Mendelism. Journal of the History of Biology, 33(2), 315–347.CrossRefGoogle Scholar
Artavanis-Tsakonas, S., Rand, M. D., and Lake, R. J. (1999). Notch signaling: Cell fate control and signal integration in development. Science, 284(5415), 770–776.CrossRefGoogle ScholarPubMed
Auerbach, C. (1967). The chemical production of mutations. Science, 158, 1141–1147.CrossRefGoogle ScholarPubMed
Auerbach, C., and Robson, J. M. (1947). The production of mutations by chemical substances. Proceedings of the Royal Society of Edinburgh, series B, 62, 271–283.Google ScholarPubMed
Auerbach, C., Robson, J. M., and Carr, J. G. (1947). The chemical production of mutations. Science, 105, 243–247.CrossRefGoogle ScholarPubMed
Avery, O. T., MacLeod, C. M., and McCarty, M. (1944). Studies on the chemical nature of the substance inducing transformation of Pneumococcal types. Journal of Experimental Medicine, 79, 137–158.CrossRefGoogle ScholarPubMed
Baltimore, D. (1970). RNA-dependent polymerase in virions of RNA tumor viruses. Nature and System, 226, 1209–1211.CrossRefGoogle Scholar
Barkow, J. H., Cosmides, L., and Tooby, , J. (1992). The Adapted Mind: Evolutionary Psychology and the Generation of Culture. New York: Oxford University Press.Google Scholar
Bacq, Z. M., and Alexander, P. (1955). Fundamentals of Radiobiology. London: Butterworths Scientific Publications.Google Scholar
Bateson, W. (1894). Materials for the Study of Variation. London: Macmillan.Google Scholar
Bateson, W. (1905 [1928]). A suggestion as to the nature of the “walnut” comb in fowls. In Punnett, R. C. (ed.), Scientific Papers of William Bateson (Vol. II, pp. 135–138). Cambridge:Cambridge University Press.Google Scholar
Bateson, W. (1907). Facts limiting the theory of heredity. Science N.S., 26(672), 649–660.CrossRefGoogle ScholarPubMed
Bateson, W. (1913 [1979]). Problems of Genetics. New Haven: Yale University Press.Google Scholar
Bateson, W. (1926 [1928]). Segregation. In Punnett, R. C. (ed.), Scientific Papers of William Bateson (Vol. II, pp. 405–440). Cambridge:Cambridge University Press.Google Scholar
Bateson, W., and Punnett, R. C. (1911). On gametic series involving reduplication of certain terms. Journal of Genetics, 1(4), 293–302.CrossRefGoogle Scholar
Bateson, W., and Saunders, E. R. (1902). Reports to the Evolution Committee of the Royal Society (pp. 1–160) (Report I. – Experiments undertaken by W. Bateson, F.R.S. and Miss E. R. Saunders). London: Royal Society.Google Scholar
Baur, E. (1912). Ein Fall von geschlechtsbegrenzter Vererbung bei Melandrium album. Zeitschrift für induktive Abstammungs- und Vererbungslehre, 8, 335–336.Google Scholar
Beadle, G. W. (1945). Biochemical genetics. Chemical Review, 37, 15–96.CrossRefGoogle Scholar
Beadle, G. W., and Ephrussi, B. (1936). The differentiation of eye pigments in Drosophila as studied by transplantation. Genetics, 21(3), 225–247.Google ScholarPubMed
Beadle, G. W., and Tatum, E. L. (1941a). Experimental control of development and differentiation: Genetic control of developmental reactions. The American Naturalist, 75, 107–116.CrossRefGoogle Scholar
Beadle, G. W., and Tatum, E. L. (1941b). Genetic control of biochemical reaction in Neurospora. Proceedings of the National Academy of Sciences of the USA, 27, 499–506.CrossRefGoogle Scholar
Belling, J. (1928). A working hypothesis for segmental interchange between homologous chromosomes in flowering plants. University of California Publications in Botany, 14(8), 283–291.Google Scholar
Ben-Chetrit, E., and Levy, M. (1998). Familial Mediterranean fever. Lancet, 351, 659–664.CrossRefGoogle ScholarPubMed
Bender, W., Akam, M., Karch, F., Beachy, P., Pfeifer, M., Spierer, P., et al. (1983). Molecular genetics of the bithorax complex in Drosophila melanogaster. Science, 221, 23–29.CrossRefGoogle ScholarPubMed
Benzer, S. (1957). The elementary units of heredity. In Glass, B. and McElroy, W. D. (eds.), The Chemical Basis of Heredity (pp. 70–93). Baltimore, MD: Johns Hopkins Press.Google Scholar
Benzer, S. (1973). Genetic dissection of behavior. Scientific American, 229(6), 24–37.CrossRefGoogle ScholarPubMed
Benzer, S., and Champe, S. P. (1961). Ambivalent rII mutants of phage T4. Proceedings of the National Academy of Sciences of the USA, 47, 1025–1038.CrossRefGoogle ScholarPubMed
Benzer, S., and Champe, S. P. (1962). A change from nonsense to sense in the genetic code. Proceedings of the National Academy of Sciences of the USA 48, 1114–1121.CrossRefGoogle ScholarPubMed
Benzer, S., and Freese, E. (1958). Induction of specific mutations with 5-bromouracil. Proceedings of the National Academy of Sciences of the USA, 44(2), 112–119.CrossRefGoogle ScholarPubMed
Berry, A. J., Ajioka, J. W., and Kreitman, M. (1991). Lack of polymorphism on the Drosophila fourth chromosome resulting from selection. Genetics, 129, 1111–1117.Google ScholarPubMed
Bishop, B. E. (1996). Mendel's opposition to evolution and to Darwin. Journal of Heredity, 87(3), 205–213.CrossRefGoogle Scholar
Blakeslee, A. F. (1922). Variations in Datura due to changes in chromosome number. The American Naturalist, 56, 16–31.CrossRefGoogle Scholar
Blixt, S. (1975). Why didn't Mendel find linkage?Nature, 256, 206.CrossRefGoogle ScholarPubMed
Bolton, E. T., and McCarthy, B. J. (1962). A general method for the isolation of RNA complementary to DNA. Proceedings of the National Academy of Sciences of the USA, 48(8), 1390–1397.CrossRefGoogle Scholar
Boveri, T. (1902). Über mehrpolige Mitosen als Mittel zur Analyse des Zellkerns. Verhandlungen der physikalische-medizinische Gesellschaft, Würzburg, N.F., 35, 67–90.Google Scholar
Boyce, R. P., and Howard-Flanders, P. (1964). Release of ultraviolet light-induced thymine dimers from DNA in E. coli K12. Proceedings of the National Academy of Sciences of the USA, 51, 293–300.CrossRefGoogle Scholar
Brannigan, A. (1979). The reification of Mendel. Social Studies of Science, 9, 423–454.CrossRefGoogle Scholar
Brenner, S. (1957). On the impossibility of all overlapping triplet codes in information transfer from nucleic acid to proteins. Proceedings of the National Academy of Sciences of the USA, 43(8), 687–694.CrossRefGoogle ScholarPubMed
Brenner, S. (1974). The genetics of Caenorhabditis elegans. Genetics, 77, 71–94.Google ScholarPubMed
Brenner, S. (2000). The end of the beginning. Science, 287(5461), 2173–2174.CrossRefGoogle ScholarPubMed
Brenner, S., Dove, W., Herskowitz, I., and Thomas, R. (1990). Genes and development: Molecular and logical themes. Genetics, 126(3), 479–486.Google ScholarPubMed
Bridges, C. B. (1913). Non-disjunction of the sex chromosomes of Drosophila. Journal of Experimental Zoology, 15, 587–605.CrossRefGoogle Scholar
Bridges, C. B. (1914). Direct proof through non-disjunction that the sex-linked genes of Drosophila are borne by the X-chromosome. Science, 40, 107–109.CrossRefGoogle ScholarPubMed
Bridges, C. B. (1916). Non-disjunction as proof of the chromosome theory of heredity. Genetics, 1, 1–52 and 107–163.Google ScholarPubMed
Bridges, C. B. (1917). Deficiency. Genetics, 2, 445–465.Google ScholarPubMed
Bridges, C. B. (1925). Sex in relation to chromosomes and genes. The American Naturalist, 59, 127–137.CrossRefGoogle Scholar
Bridges, C. B. (1935). Salivary chromosome maps. The Journal of Heredity, 26, 60–64.CrossRefGoogle Scholar
Bridges, C. B. (1938). A revised map of the salivary gland X-chromosome. The Journal of Heredity, 29, 11–13.CrossRefGoogle Scholar
Bridges, C. B., and Anderson, E. G. (1925). Crossing over in the X chromosome of triploid females of Drosophila melanogaster. Genetics, 10, 418–441.Google Scholar
Bridges, C. B., and Bridges, P. N. (1939). A new map of the second chromosome. The Journal of Heredity, 30, 475–477.CrossRefGoogle Scholar
Britten, R. J., and Davidson, E. H. (1969). Gene regulation for higher cells: A theory. Science, 165, 349–357.CrossRefGoogle ScholarPubMed
Britten, R. J., and Kohne, D. E. (1968). Repeated sequences in DNA. Science, 161(3841), 529–540.CrossRefGoogle ScholarPubMed
Brown, S. W., and Zohary, D. (1955). The relationship of chiasmata and crossing over in Lilium formosanum. Genetics, 40, 850–873.Google ScholarPubMed
Bruce, A. B. (1910). The Mendelian theory of heredity and the augmentation of vigor. Science, 32, 627–628.CrossRefGoogle ScholarPubMed
Bull, A. (1966). Bicaudal, a genetic factor which affects the polarity of the embryo in Drosophila melanogaster. Journal of Experimental Zoology, 161, 221–241.CrossRefGoogle Scholar
Cairns, J. (1963). The bacterial chromosome and its manner of replication as seen in autoradiography. Journal of Molecular Biology, 6(3), 208–213.CrossRefGoogle ScholarPubMed
Cairns, J., Overbaugh, J., and Miller, S. (1988). The origin of mutants. Nature, 335, 142–145.CrossRefGoogle ScholarPubMed
Campbell, M. (1982). Mendel's theory: Its context and plausibility. Centaurus, 26, 38–69.CrossRefGoogle ScholarPubMed
Carlson, E. A. (1959). Allelism, pseudoallelism, and complementation at the dumpy locus in D. melanogaster. Genetics, 44, 347–373.Google Scholar
Carlson, E. A. (1966 [1989]). The Gene: A Critical History. Philadelphia: W. B. Saunders; Ames, IA: Iowa State University Press.Google Scholar
Carlson, E. A. (2004). Mendel's Legacy: The Origin of Classical Genetics. Cold Spring Harbor, NY:Cold Spring Harbor Laboratory Press.Google Scholar
Carothers, E. E. (1913). The Mendelian ratio in relation to certain Orthopteran chromosomes. Journal of Morphology, 24, 487–511.CrossRefGoogle Scholar
Casperson, T., and Schultz, J. (1938). Nucleic acid metabolism of the chromosomes in relation to gene reproduction. Nature, 142, 294–295.CrossRefGoogle Scholar
Castle, W. E. (1906). Yellow mice and gametic purity. Science N.S., 24, 275–281.CrossRefGoogle ScholarPubMed
Castle, W. E. (1913a). Heredity. New York: D. Appleton and Comp.Google Scholar
Castle, W. E. (1913b). Simplification of Mendelian formulae. The American Naturalist, 47(555), 170–182.CrossRefGoogle Scholar
Castle, W. E. (1915). Mr. Muller on the constancy of Mendelian factors. The American Naturalist, 49, 37–42.CrossRefGoogle Scholar
Castle, W. E. (1919a). Is the arrangement of the genes in the chromosome linear?Proceedings of the National Academy of Sciences of the USA, 5(2), 25–32.CrossRefGoogle ScholarPubMed
Castle, W. E. (1919b). Piebald rats and selection. The American Naturalist, 53, 370–376.CrossRefGoogle Scholar
Castle, W. E. (1919c). Piebald rats and the theory of genes. Proceedings of the National Academy of Sciences of the USA, 5, 126–130.CrossRefGoogle ScholarPubMed
Chovnick, A. (1961). The garnet locus in Drosophila melanogaster. I. Pseudoallelism. Genetics, 46, 493–507.Google ScholarPubMed
Chovnick, A. (1989). Intragenic recombination in Drosophila: The rosy locus. Genetics, 123(4), 621–624.Google ScholarPubMed
Chovnick, A., Schalet, A., Kernaghan, R. P., and Kraus, M. (1964). The rosy cistron in Drosophila melanogaster genetic fine structure analysis. Genetics, 50, 1254–1259.Google ScholarPubMed
Churchill, F. B. (1974). William Johannsen and the genotype concept. Journal of the History of Biology, 7, 5–30.CrossRefGoogle ScholarPubMed
Churchill, F. B. (1987). From heredity theory to Vererbung. The transmission problem, 1850–1915. Isis, 78(3), 337–364.CrossRefGoogle ScholarPubMed
Cohen, J. (2007). DNA duplications and deletions help determine health. Science, 317(5843), 1315–1317.CrossRefGoogle ScholarPubMed
Coleman, W. (1970). Bateson and chromosomes: Conservative thought in science. Centaurus, 15(3–4), 228–314.CrossRefGoogle Scholar
Cooper, K. W. (1948). A new theory of secondary non-disjunction in female Drosophila melanogaster. Proceedings of the National Academy of Sciences of the USA, 34, 179–187.CrossRefGoogle ScholarPubMed
Correns, C. (1902 [1924]). Scheinbare Ausnahmen von der Mendel'schen Spaltungregel für Bastarde. In Carl Correns: Gesammelte Abhandlungen zur Vererbungswissenschaft aus periodischen Schrifren, 1899–1924 (pp. 287–299). Berlin: Julius Springer.Google Scholar
Correns, C. (1903 [1924a]). Über die dominierenden Merkmale der Bastarde. In Carl Correns: Gesammelte Abhandlungen zur Vererbungswissenschaft aus periodischen Schrifren, 1899–1924 (pp. 329–343). Berlin: Julius Springer.CrossRefGoogle Scholar
Correns, C. (1903 [1924b]). Weitere beiträge zur Kenntnis der dominierenden Merkmale und der Mosaikbildung der Bastarde. In Carl Correns: Gesammelte Abhandlungen zur Vererbungswissenschaft aus periodischen Schrifren, 1899–1924 (pp. 342–349). Berlin: Julius Springer.CrossRefGoogle Scholar
Creager, A. N. H. (2004). Mapping genes in microorganisms. In Gaudillière, J.-P. and Rheinberger, H.-J. (eds.), From Molecular Genetics to Genomics: The Mapping Cultures of Twentieth-Century Genetics (Vol. XXII, pp. 9–41). London: Routledge.Google Scholar
Creighton, H. B., and McClintock, B. (1931). A correlation of cytological and genetical crossing-over in Zea mays. Proceedings of the National Academy of Sciences of the USA, 17, 492–497.CrossRefGoogle ScholarPubMed
Crick, F. H. C. (1958). On protein synthesis. In Symposium of the Society for Experimental Biology. The Biological Replication of Macromolecules (Vol. 12, pp. 138–163). Cambridge: Cambridge University Press.Google Scholar
Crick, F. H. C. (1966). On Molecules and Men. Seattle, WA:University of Washington Press.Google Scholar
Crick, F. H. C. (1970). Central dogma of molecular biology. Nature, 227(5258), 561–563.CrossRefGoogle ScholarPubMed
Crick, F. H. C., Barnett, L., Brenner, S., and Watts-Tobin, R. J. (1961). General nature of the genetic code for proteins. Nature, 192(4809), 1227–1232.CrossRefGoogle ScholarPubMed
Crick, F. H. C., and Lawrence, P. A. (1975). Compartments and polyclones in insect development. Science, 189, 340–347.CrossRefGoogle ScholarPubMed
Crow, J. F., and Kimura, M. (1970). An Introduction to Population Genetics Theory. New York: Harper and Row.Google Scholar
Crow, J. F., and Temin, R. G. (1964). Evidence for the partial dominance of recessive lethal genes in natural populations of Drosophila. The American Naturalist, 98, 21–33.CrossRefGoogle Scholar
Danchin, A., and He'enaut, A. (1997). The map of the cell is in the chromosome. Current Opinions in Genetics and Development, 7(6), 852–854.CrossRefGoogle ScholarPubMed
Darbishire, A. D. (1902). Note on the results of crossing Japanese Waltzing mice with European Albino races. Biometrika, 2, 101–104.Google Scholar
Darden, L. (1991). Theory Change in Science: Strategies from Mendelian Genetics. New York: Oxford University Press.Google Scholar
Darlington, C. D. (1931). Meiosis (precocity theory). Biological Review, 6, 221–264.Google Scholar
Darlington, C. D. (1937). Recent Advances in Cytology (2nd edn.). London: J and A Churchill.Google Scholar
Darnell, J. E., Jr. (1999). E.B. Wilson Lecture, 1998. Eukaryotic RNAs: Once more from the beginning. Molecular Biology of the Cell, 10(6), 1685–1692.CrossRefGoogle Scholar
Davis, R. H., and Perkins, D. D. (2002). Neurospora: a model of model microbes. Nature Reviews: Genetics, 3(5), 397–403.CrossRefGoogle ScholarPubMed
Davis, R. L. (2000). Neurofibromin progress on the fly. Nature, 403(6772), 846–847.CrossRefGoogle ScholarPubMed
Darwin, C. (1868). The Variation of Animals and Plants under Domestication (Vol. I and II). London: Murray.Google Scholar
Darwin, C. (1981 [1871]). The Descent of Man and Selection in Relation to Sex. New York: D. Appleton.CrossRefGoogle Scholar
Dawkins, R. (1976). The Selfish Gene. New York: Oxford University Press.Google Scholar
Vries, H. (1889). Intracelluläre Pangenesis. Jena: Gustav Fischer.CrossRefGoogle Scholar
Vries, H. (1889 [1910]). Intracellular Pangenesis. Chicago, IL: Open Court.CrossRefGoogle Scholar
Vries, H. (1902–3). Die Mutationstheorie: Versuche und Beobachtungen Über die Entstehung von Arten im Pflanzenreich. Leipzig: Von Veit.Google Scholar
Vries, H. (1912 [c. 1904]). Species and Varieties: Their Origin by Mutation (3rd, corrected and revised edn.). Chicago, IL: The Open Court Publishing Company.Google Scholar
Vries, H. (1950 [1900]). Concerning the law of segregation of hybrids. Genetics, 35 Supplement (The birth of Genetics) (5, part 2), 30–32.Google Scholar
Deichmann, U. (2004). Early responses to Avery et al.'s paper on DNA as hereditary material. Historical Studies in the Physical and Biological Sciences, 34(2), 207–232.CrossRefGoogle Scholar
Demerec, M., and Hoover, M. E. (1936). Three related X-chromosome deficiencies in Drosophila. The Journal of Heredity, 27, 206–212.CrossRefGoogle Scholar
Dennis, C. (2002a). Gene regulation: The brave new world of RNA. Nature, 418(6894), 122–124.CrossRefGoogle ScholarPubMed
Dennis, C. (2002b). Small RNAs: The genome's guiding hand?Nature, 420(6917), 732.CrossRefGoogle ScholarPubMed
Depew, D. J., and Weber, B. H. (1995). Darwinism Evolving. Systems Dynamics and the Genealogy of Natural Selection. Cambridge, MA: The MIT Press.Google Scholar
Di Trocchio, F. (1991). Mendel's experiments: a reinterpretation. Journal of the History of Biology, 24, 485–519.CrossRefGoogle Scholar
Dobzhansky, T. (1929). Genetical and cytological proof of translocations involving the third and the fourth chromosomes of Drosophila melanogaster. Biologisches Zentralblat, 49, 408–419.Google Scholar
Dobzhansky, T. (1931). Translocations involving the second and the fourth chromosomes of Drosophila melanogaster. Genetics, 16, 629–658.Google ScholarPubMed
Dobzhansky, T. (1937). Genetics and the Origin of Species (1st edn.). New York: Columbia University Press.Google Scholar
Dobzhansky, T. (1970). Genetics of the Evolutionary Process. New York: Columbia University Press.Google Scholar
Dobzhansky, T. (1973). Nothing in biology makes sense except in the light of evolution. The American Biology Teacher, 35, 125–129.CrossRefGoogle Scholar
Duhem, P. (1914). The Aim and Structure of Physical Theory. Princeton, NJ: Princeton University Press.Google Scholar
Dunn, L. C. (1965). A Short History of Genetics. New York: McGraw Hill.Google Scholar
East, E. M. (1910). A Mendelian interpretation of variation that is apparently continuous. The American Naturalist, 44(518), 65–82.CrossRefGoogle Scholar
East, E. M. (1911). The genotype hypothesis and hybridization. The American Naturalist, 45, 160–174.CrossRefGoogle Scholar
East, E. M. (1912). The Mendelian notation as a description of physiological facts. The American Naturalist, 46, 633–655.CrossRefGoogle Scholar
East, E. M. (1923). Mendel and his contemporaries. The Scientific Monthly, 6, 225–237.Google Scholar
Eldredge, N., and Gould, S. J. (1972). Punctuated equilibrium: An alternative to phyletic gradualism. In Schopf, T. J. M. (ed.), Models of Paleobiology (pp. 82–115). San Francisco, CA: Freeman, Cooper and Co.Google Scholar
Emerson, R. A. (1914). The inheritance of a recurring somatic variation in variegated ears of maize. The American Naturalist, 48, 87–115.CrossRefGoogle Scholar
Ereshefsky, M. (2001). The Poverty of the Linnaean Hierarchy: A Philosophical Study of Biological Taxonomy. New York: Cambridge University Press.CrossRefGoogle Scholar
Falk, R. (1955). Some preliminary observations on crossing-over and non-disjunction. Hereditas, 41, 376–383.CrossRefGoogle Scholar
Falk, R. (1961). Are induced mutations in Drosophila overdominant? II. Experimental results. Genetics, 46(7), 737–757.Google ScholarPubMed
Falk, R. (1967). Fitness of heterozygotes for irradiated chromosomes in Drosophila. Mutation Research, 4, 805–819.CrossRefGoogle ScholarPubMed
Falk, R. (1986). What is a gene?Studies in the History and Philosophy of Science, 17(2), 133–173.CrossRefGoogle ScholarPubMed
Falk, R. (1988). Species as individuals. Biology and Philosophy, 3(4), 455–462.CrossRefGoogle Scholar
Falk, R. (1991). The dominance of traits in genetic analysis. Journal of the History of Biology, 24(3), 457–484.CrossRefGoogle ScholarPubMed
Falk, R. (1995). The struggle of genetics for independence. Journal of the History of Biology, 28(2), 219–246.CrossRefGoogle ScholarPubMed
Falk, R. (2000a). Can the norm of reaction save the gene concept? In Singh, R. S., Krimbas, C. B., Paul, D. B. and Beatty, J. (eds.), Thinking About Evolution: Historical, Philosophical and Political Perspectives (Vol. II, pp. 119–140). Cambridge: Cambridge University Press.Google Scholar
Falk, R. (2000b). The gene – a concept in tension. In Beurton, P. J., Falk, R. and Rheinberger, H.-J. (eds.), The Concept of the Gene in Development and Evolution: Historical and Epistemological Perspectives (pp. 317–348). Cambridge: Cambridge University Press.Google Scholar
Falk, R. (2001a). Mendel's hypothesis. In Allen, G. E. and MacLeod, R. M. (eds.), Science, History and Social Activism: A Tribute to Everett Mendelsohn (pp. 77–86). Dordrecht: Kluwer.Google Scholar
Falk, R. (2001b). The rise and fall of dominance. Biology and Philosophy, 16(3), 285–323.CrossRefGoogle Scholar
Falk, R. (2003). Linkage: From particulate to interactive genetics. Journal of the History of Biology, 36(1), 87–117.CrossRefGoogle ScholarPubMed
Falk, R. (2004). Long Live the Genome! So should the gene. History and Philosophy of the Life Sciences, 26, 105–121.CrossRefGoogle ScholarPubMed
Falk, R. (2006). Mendel's impact. Science in Context, 19(2), 215–236.CrossRefGoogle Scholar
Falk, R. (2007). Genetic analysis. In Matthen, M. and Stephens, C. (eds.), Philosophy of Biology (Vol. III, pp. 249–308). Dordrecht: Elsevier.Google Scholar
Falk, R. (2008). Wilhelm Johannsen: A rebel or a diehard? In Harman, O. and Dietrich, M. R. (eds.), Rebels, Mavericks, and Heretics in Biology (pp. 66–83). New Haven, CT: Yale University Press.Google Scholar
Falk, R. (in press). Genetic regulation before the 1960s. In Laubichler, M., Rheinberger, H.-J. and Hammerstein, P. (eds.), Regulation: Historical and Current Themes in Theoretical Biology.
Falk, R., and Sarkar, S. (1991). The real objective of Mendel's paper. Biology and Philosophy, 6, 447–451.CrossRefGoogle Scholar
Falk, R., and Sarkar, S. (2006). Genetics. In Sarkar, S. and Pfeifer, J. (eds.), The Philosophy of Science: An Encyclopedia (Vol. I, pp. 330–339). New York: Routledge.Google Scholar
Falk, R., and Schwartz, S. (1993). Morgan's hypothesis of the genetic control of development. Genetics, 134(3), 671–674.Google ScholarPubMed
Farber, P. L. (1976). The type-concept in zoology during the first half of the nineteenth century. Journal of the History of Biology, 9(1), 93–119.Google Scholar
Feldman, M. (2001). The origin of cultivated wheat. In Bonjean, A. and Angus, W. (eds.), The World of Wheat Book (pp. 3–56). Paris: Lavoisier Tech. and Doc.Google Scholar
Fincham, J. R. S. (1998). Fungal genetics – past and present. Journal of Genetics, 77(2 and 3), 55–63.CrossRefGoogle Scholar
Fincham, J. R. S., Day, P. R., and Radford, A. (1979). Fungal Genetics (4th edn.). London: Blackwell.Google Scholar
Fisher, R. A. (1918). The correlation between relatives on the supposition of Mendelian inheritance. Transactions of the Royal Society, Edinburgh, 52, 399–433.CrossRefGoogle Scholar
Fisher, R. A. (1922). On the dominance ratio. Proceedings of the Royal Society of Edinburgh, series B, 42, 321–341.CrossRefGoogle Scholar
Fisher, R. A. (1928). The possible modification of the response of the wild type to recurrent mutation. The American Naturalist, 62, 115–126.CrossRefGoogle Scholar
Fisher, R. A. (1930). The Genetical Theory of Natural Selection. Oxford: Clarendon.CrossRefGoogle Scholar
Fisher, R. A. (1936). Has Mendel's work been rediscovered?Annals of Science, 1, 115–137.CrossRefGoogle Scholar
Ford, E. B. (1964). Ecological Genetics. London/New York: Methuen/John Wiley.Google Scholar
Franklin, I., and Lewontin, R. C. (1970). Is the gene the unit of selection?Genetics, 65(4), 707–734.Google ScholarPubMed
Freese, E. (1957). The correlation effect for a histidine locus of Neurospora crassa. Genetics, 42, 671–684.Google ScholarPubMed
Freese, E. (1959). The specific mutagenic effect of base analogues. Journal of Molecular Biology, 1, 87–105.CrossRefGoogle Scholar
Fuerst, J. A. (1982). The role of reductionism in the development of molecular biology: Peripheral or central?Social Studies of Science, 12, 241–278.CrossRefGoogle ScholarPubMed
Fuller, S. (2000). Thomas Kuhn: A philosophical history for our times. Chicago, IL: University of Chicago Press.Google Scholar
Galton, F. (1875). A theory of heredity (Revised version). Journal of the Anthropological Institute, 5, 329–348.Google Scholar
Galton, F. (1908). Memories of My Life (2nd edn.). London: Methuen.CrossRefGoogle Scholar
Gamow, G. (1954). Possible relation between deoxyribonucleic acid and protein structure. Nature, 173, 318.CrossRefGoogle Scholar
Garcia-Bellido, A. (1998). The engrailed story. Genetics, 148(2), 539–544.Google ScholarPubMed
Garcia-Bellido, A., Lawrence, P. A., and Morata, G. (1979). Compartments in animal development. Scientific American, 241(1), 102–110.CrossRefGoogle Scholar
Garcia-Bellido, A., and Merriam, J. R. (1969). Cell lineage of the imaginal disc in Drosophila gynandromorphs. Journal of Experimental Zoology, 170, 61–76.CrossRefGoogle ScholarPubMed
Garcia-Bellido, A., and Merriam, J. R. (1971). Parameters of the wing imaginal disc development of Drosophila melanogaster. Developmental Biology, 24, 61–87.CrossRefGoogle ScholarPubMed
Garcia-Bellido, A., Rippoll, P., and Morata, G. (1973). Developmental compartmentalisation of the wing disc of Drosophila. Nature: New Biology, 245, 251–253.Google Scholar
Garcia-Bellido, A., and Santamaria, P. (1972). Developmental analysis of the wing disc in the mutant engrailed of Drosophila melanogaster. Genetics, 72, 87–104.Google ScholarPubMed
Garen, A. (1992). Looking for the homunculus in Drosophila. Genetics, 131(1), 5–7.Google ScholarPubMed
Garrod, A. E. (1908). Inborn Errors of Metabolism (2nd edn. 1923). Oxford: Oxford University Press.Google Scholar
Gasper, P. (1992). Reduction and instrumentalism in genetics. Philosophy of Science, 59, 655–670.CrossRefGoogle Scholar
Gayon, J. (2000). From measurement to organization: A philosophical scheme for the history of the concept of heredity. In Beurton, P. J., Falk, R. and Rheinberger, H.-J. (eds.), The Concept of the Gene in Development and Evolution: Historical and Epistemological Perspectives (pp. 69–90). Cambridge: Cambridge University Press.Google Scholar
Gehring, W. J. (1967). Clonal analysis of determination dynamics in cultures of imaginal discs of Drosophila melanogaster. Developmental Biology, 16, 438–456.CrossRefGoogle Scholar
Gehring, W. J. (1998). Master Control Genes in Development and Evolution: The Homeobox Story. New Haven, CT: Yale University Press.Google Scholar
Gehring, W. J., and Nöthiger, R. (1973). The imaginal discs of Drosophila. In Counce, S. J. and Waddington, C. H. (eds.), Developmental Systems (Vol. II, pp. 211–290). London: Academic Press.Google Scholar
Gerstein, M. B., Lan, N., and Jansen, R. (2002). Integrating interactomes. Science, 295(5553), 284–287.CrossRefGoogle ScholarPubMed
Gerstein, M. B., Bruce, C., Rozowsky, J. S., Zheng, D., Du, J., Korbel, J. O., et al. (2007). What is a gene, post-ENCODE? History and updated definition. Genome Research, 17(6), 669–681.CrossRefGoogle ScholarPubMed
Ghiselin, M. T. (1971). The individual in the Darwinian revolution. New Literary History, 3, 113–134.CrossRefGoogle Scholar
Ghiselin, M. T. (1987). Species concepts, individuality, and objectivity. Biology and Philosophy, 2(2), 127–143.Google Scholar
Gibbs, W. W. (2003). The unseen genome: Gems among the junk. Scientific American, 289(5), 46–53.CrossRefGoogle ScholarPubMed
Gilbert, S. F. (1978). The embryological origins of the gene theory. Journal of the History of Biology, 11(2), 307–351.CrossRefGoogle ScholarPubMed
Gilbert, S. F. (1991a). Developmental Biology (3rd edn.). Sunderland, MA: Sinauer.Google Scholar
Gilbert, S. F. (1991b). Induction and the origins of developmental genetics. In Gilbert, S. (ed.), Developmental Biology. A Comprehensive Synthesis. A Conceptual History of Modern Embryology (pp. 181–206). Baltimore, MD: Johns Hopkins University Press.CrossRefGoogle Scholar
Gilbert, S. F. (1998). Bearing crosses: A historiography of genetics and embryology. American Journal of Medical Genetics, 76(2), 168–182.3.0.CO;2-J>CrossRefGoogle ScholarPubMed
Gilbert, S. F., Opitz, J. M., and Raff, R. A. (1996). Resynthesizing evolutionary and developmental biology. Developmental Biology, 173, 357–372.CrossRefGoogle ScholarPubMed
Gingeras, T. R. (2007). Origin of phenotypes: Genes and transcripts. Genome Research, 17(6), 682–690.CrossRefGoogle ScholarPubMed
Glass, B. (1963). The establishment of modern genetical theory as an example of the interaction of different models, techniques, and inferences. In Crombie, A. C. (ed.), Scientific Change: Historical studies in the intellectual, social and technical conditions for scientific discovery and technical invention, from antiquity to the present (pp. 521–541). London: Heinemann.Google Scholar
Gliboff, S. (1999). Gregor Mendel and the law of evolution. History of Science, 37, 217–235.CrossRefGoogle Scholar
Golding, B. (1994). Non-Neutral Evolution: Theories and Molecular Data. New York: Chapman and Hall.CrossRefGoogle Scholar
Goldschmidt, R. (1917). Crossing over Ohne Chiasmatypie?Genetics, 2(1), 82–95.Google ScholarPubMed
Goldschmidt, R. (1938a). The theory of the gene. Scientific Monthly, 45, 268–273.Google Scholar
Goldschmidt, R. (1938b). Physiological Genetics. New York: McGraw-Hill.Google Scholar
Goldschmidt, R. (1950). Fifty years of Genetics. The American Naturalist, 84, 313–340.CrossRefGoogle Scholar
Goldschmidt, R. (1954). Different philosophies of genetics. Science, 119, 703–710.CrossRefGoogle ScholarPubMed
Goldschmidt, R. (1955). Theoretical Genetics. Berkeley, CA: University of California Press.CrossRefGoogle Scholar
Gould, S. J. (1977). Ontogeny and Phylogeny. Cambridge, MA: Harvard University Press.Google Scholar
Gould, S. J., and Lewontin, R. C. (1979). The spandrels of San Marco and the Panglossian paradigm: A critique of the adaptationist programme. Proceedings of the Royal Society of London, series B, 205, 581–598.CrossRefGoogle Scholar
Gowen, J. W. (1952). Heterosis. Ames, IA:Iowa State College Press.Google Scholar
Gray, J. (1969). Near Eastern Mythology, Mesopotamia, Syria, Palestine. London: The Hamlyn Publishing Group.Google Scholar
Green, M. M. (1963). Interallelic complementation and recombination at the rudimentary wing locus in Drosophila melanogaster. Genetica, 34, 242–253.CrossRefGoogle Scholar
Greider, C. W., and Blackburn, E. H. (1985). Identification of a specific telomere terminal transferase activity in Tetrahymena extracts. Cell, 43, 405–413.CrossRefGoogle ScholarPubMed
Grell, R. F. (1976). Distributive pairing. In Ashburner, M. and Novitski, E. (eds.), The Genetics and Biology of Drosophila (Vol. Ia, pp. 435–486). London: Academic Press.Google Scholar
Griesemer, J. R. (2000). Reproduction and reduction of genetics. In Beurton, P. J., Falk, R. and Rheinberger, H.-J. (eds.), The Concept of the Gene in Development and Evolution: Historical and Epistemological Perspectives (pp. 240–285). Cambridge: Cambridge University Press.Google Scholar
Griffiths, A. J. F., Gelbart, W. M., Miller, J. H., and Lewontin, R. C. (1999). Modern Genetic Analysis. New York: Freeman.Google Scholar
Griffiths, P. E. (2006). The fearless vampire conservator: Philip Kitcher, genetic determinism, and the informational gene. In Neumann-Held, E. M. and Rehmann-Sutter, C. (eds.), Genes in Development: Re-reading the Molecular Paradigm (pp. 175–198). Durham: Duke University Press.Google Scholar
Griffiths, P. E., and Stotz, K. (2006). Genes in the postgenomic era. Theoretical Medicine and Bioethics, 27(6), 499–521.CrossRefGoogle ScholarPubMed
Gruneberg, H. (1938). An analysis of the “pleiotropic” effects of a new lethal mutation in the rat (Mus norvegicus). Proceedings of the Royal Society of London, series B, 125, 123–144.CrossRefGoogle Scholar
Hadorn, E. (1961). Developmental Genetics and Lethal Factors. New York: John Wiley.Google Scholar
Hadorn, E. (1968). Transdetermination in cells. Scientific American, 219 (5), 110–118.CrossRefGoogle ScholarPubMed
Haldane, J. B. S. (1930). A note on Fisher's theory of the origin of dominance, and a correlation between dominance and linkage. The American Naturalist, 64, 87–90.CrossRefGoogle Scholar
Haldane, J. B. S. (1954). The Biochemistry of Genetics. London: George Allen and Unwin.Google Scholar
Haldane, J. B. S. (1957). The cost of natural selection. Journal of Genetics, 55, 511–524.CrossRefGoogle Scholar
Hamilton, W. D. (1963). The evolution of altruistic behavior. The American Naturalist, 97, 354–356.CrossRefGoogle Scholar
Hamilton, W. D. (1964a). The genetical evolution of social behavior. I. Journal of Theoretical Biology, 7, 1–16.CrossRefGoogle Scholar
Hamilton, W. D. (1964b). The genetical evolution of social behavior. II. Journal of Theoretical Biology, 7, 17–52.CrossRefGoogle Scholar
Hardy, G. H. (1908). Mendelian proportions in a mixed population. Science, 28, 49–50.CrossRefGoogle Scholar
Hardy, G. H. (1940). A Mathematician's Apology. Cambridge: Cambridge University Press.Google Scholar
Harman, O. S. (2004). The Man Who Invented the Chromosome: A Life of Cyril Darlington. Cambridge, MA: Harvard University Press.Google Scholar
Harper, L., Golubovskaya, I., and Cande, W. Z. (2004). A bouquet of chromosomes. Journal of Cell Science, 117(18), 4025–4032.CrossRefGoogle Scholar
Harris, H. (1966). Enzyme polymorphism in man. Proceedings of the Royal Society of London, series B, 164, 298–310.CrossRefGoogle Scholar
Hartman, P. E., Loper, J. C., and Serman, D. (1960). Fine structure mapping by complete transduction between histidine-requiring Salmonella mutants. Journal of General Microbiology, 22, 323–353.CrossRefGoogle ScholarPubMed
Harwood, J. (1987). National style in science: Genetics in Germany and the United Stated between the world wars. Isis, 78, 390–414.CrossRefGoogle Scholar
Harwood, J. (1993). Styles of Scientific Thought. The German Genetics Community 1900–1933. Chicago, IL: The University of Chicago Press.Google Scholar
Harwood, J. (1996). Weimar culture and biological theory: A study of Richard Woltereck (1877–1944). History of Science, 34, 347–377.CrossRefGoogle Scholar
Hastings, P. J., and Whitehouse, H. L. K. (1964). A polaron model of genetic recombination by the formation of hybrid deoxyribonucleic acid. Nature, 201, 1052–1054.CrossRefGoogle ScholarPubMed
Hayes, W. (1953). The mechanism of genetic recombination in E. coli. Cold Spring Harbor Symposia on Quantitative Biology, 18, 75–93.CrossRefGoogle Scholar
Hayes, W. (1964). The Genetics of Bacteria and Their Viruses: Studies in Basic Genetics and Molecular Biology. New York: John Wiley.Google Scholar
Heitz, E. (1929). Heterochromatin, Chromozentern, Chromomeren. Berichte der deutschen botanischen Gesellschaft, 47, 274–284.Google Scholar
Heitz, E., and Bauer, H. (1933). Beweise für die Chromosomennatur der Kernschleifen in den Knäuelkernen von Bibio hortulanus. Zeitschrift für Zellforschung und mikroskopische Anatomie, 17, 76–82.CrossRefGoogle Scholar
Helinski, D. R., and Yanofsky, C. (1962). Correspondence between genetic data and the position of amino acid alteration in a protein. Proceedings of the National Academy of Sciences of the USA, 48, 173–183.CrossRefGoogle Scholar
Hershey, A. D., and Chase, M. (1952). Independent functions of viral protein and nucleic acid in growth of bacteriophage. Journal of General Physiology, 36(1), 39–56.CrossRefGoogle ScholarPubMed
Holliday, R. (1964). A mechanism for gene conversion in fungi. Genetical Research, 5, 282–304.CrossRefGoogle Scholar
Holmes, F. L. (2000). Seymour Benzer and the definition of the gene. In Beurton, P. J., Falk, R. and Rheinberger, H.-J. (eds.), The Concept of the Gene in Development and Evolution (pp. 115–155). Cambridge: Cambridge University Press.Google Scholar
Holmes, F. L. (2001). Meselson, Stahl, and the Replication of DNA: A History of “The Most Beautiful Experiment in Biology”. New Haven, CT: Yale University Press.CrossRefGoogle Scholar
Hornstein, E., and Shomron, N. (2006). Canalization of development by microRNAs. Nature Genetics, 38 (Supplement), S20–S24.CrossRefGoogle ScholarPubMed
Horowitz, N. H. (1991). Fifty years ago: The Neurospora revolution. Genetics, 127(4), 631–635.Google ScholarPubMed
Hotta, Y., and Benzer, S. (1972). Mapping of behavior in Drosophila melanogaster. Nature, 240, 527–535.CrossRefGoogle Scholar
Hotta, Y., Ito, M., and Stern, H. (1966). Synthesis of DNA during meiosis. Proceedings of the National Academy of Sciences of the USA, 56, 1184–1191.CrossRefGoogle ScholarPubMed
Howard, A., and Pelc, S. R. (1951). Nuclear incorporation of P32 as demonstrated by autoradiographs. Experimental Cell Research, 2, 178–197.CrossRefGoogle Scholar
Hoyer, B. H., McCarthy, B. J., and Bolton, E. T. (1964). A molecular approach in the systematics of higher organisms. Science, 144, 959–967.CrossRefGoogle Scholar
Hubby, J. L., and Lewontin, R. C. (1966). 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.Google Scholar
Huberman, J. A., and Riggs, D. A. (1968). On the mechanism of DNA replication in mammalian chromosomes. Journal of Molecular Biology, 32, 327–341.CrossRefGoogle ScholarPubMed
Hull, D. L. (1973). Darwin and His Critics. Chicago, IL: University of Chicago Press.Google Scholar
Hull, D. L. (1974). Philosophy of Biological Science. Englewood Cliffs, NJ:Prentice-Hall.Google Scholar
Hull, D. L. (1976). Are species really individuals?Systematic Zoology, 25, 174–191.CrossRefGoogle Scholar
Hurst, C. C. (1906). Mendelian characters in plants and animals. In Report of Conference on Genetics (pp. 114–129). London: Royal Horticultural Society.Google Scholar
Hutchinson, G. E., and Rachootin, S. (1979). Historical introduction. In William Bateson: Problems of Genetics (pp. vii–xx). New Haven, CT: Yale University Press.Google Scholar
Huxley, J. (1943). Evolution: The Modern Synthesis. New York: Harper and Row.Google Scholar
Ingram, V. M. (1957). Gene mutations in human haemoglobin: the chemical difference between normal and sickle cell haemoglobin. Nature, 180, 326–328.CrossRefGoogle ScholarPubMed
Ingram, V. M. (1963). The Hemoglobin in Genetics and Evolution. New York: Columbia University Press.Google Scholar
Jacob, F., and Monod, J. (1961). Genetic regulatory mechanisms in the synthesis of proteins. Journal of Molecular Biology, 3, 318–356.CrossRefGoogle ScholarPubMed
Jacob, F., and Wollman, E. L. (1961). Sexuality and the Genetics of Bacteria. New York: Academic Press.Google Scholar
Jahn, I. (1957/58). Zur Geschichte der Wiederentdeckung der Mendelschen Gesetze. Wissenschaftliche Zeitschrift der Friedrich-Schiller Universität Jena, Mathematisch-naturwissenschaftliche Reihe, 7(2/3), 215–227.Google Scholar
Janssens, F. A. (1909). La Théorie de la chiasmatypie. Nouvelle interprétation des cinèses de maturation. La Cellule, 25, 389–411.Google Scholar
Johannsen, W. (1903). Über Erblichkeit in Populationen und reinen Linien. Jena: Gustav Fischer.Google Scholar
Johannsen, W. (1903 [1955]). Concerning heredity in populations (Über Erblichkeit in Populationen und in reinen Linien) (H. Gall and E. Putschar, Trans.). By The Staff of Natural Science 3 (ed.), Selected Readings in Biology for Natural Sciences (Vol. III, pp. 172–215). Chicago, IL: University of Chicago Press.Google Scholar
Johannsen, W. (1909). Elemente der exakten Erblichkeitslehre. Jena: Gustav Fischer.Google Scholar
Johannsen, W. (1911). The genotype conception of heredity. The American Naturalist, 45(531), 129–159.CrossRefGoogle Scholar
Johannsen, W. (1923). Some remarks about units in heredity. Hereditas, 4, 133–141.CrossRefGoogle Scholar
Johannsen, W. (1926). Elemente der exakten Erblichkeitslehre (3rd edn.). Jena: Gustav Fischer.Google Scholar
Judson, H. F. (1979). The Eighth Day of Creation: Makers of Revolution in Biology. New York: Simon and Schuster.Google Scholar
Kacser, H. (1987). Dominance not inevitable but very likely. Journal of theoretical Biology, 126, 505–506.CrossRefGoogle Scholar
Kacser, H., and Burns, J. A. (1981). The molecular basis of dominance. Genetics, 97, 639–666.Google ScholarPubMed
Kauffman, S. A. (1973). Control circuits for determination and transdetermination. Science, 181, 310–318.CrossRefGoogle ScholarPubMed
Kavenoff, R., and Zimm, B. H. (1973). Chromosome-sized DNA molecules from Drosophila. Chromosoma, 41, 1–27.CrossRefGoogle ScholarPubMed
Kay, L. E. (2000). Who Wrote the Book of Life? A History of the Genetic Code. Stanford, CA: Stanford University Press.Google Scholar
Keeble, F., and Pellew, C. (1910). The mode of inheritance of stature and of time of flowering in peas (Pisum sativum). Journal of Genetics, 1, 47–56.CrossRefGoogle Scholar
Keightley, P. D. (1996). A metabolic basis for dominance and recessivity. Genetics, 143(2), 621–625.Google ScholarPubMed
Keller, E. F. (1995). Refiguring Life: Metaphors of Twentieth-Century Biology. New York: Columbia University Press.Google Scholar
Keller, E. F. (2000). The Century of the Gene. Cambridge, MA: Harvard University Press.Google Scholar
Keller, E. F. (2002). Making Sense of Life: Explaining Biological Development with Models, Metaphors, and Machines. Cambridge, MA: Harvard University Press.Google Scholar
Kelner, A. (1949). Effect of visible light on the recovery of Streptomyces griseus conidia from ultraviolet irradiation injury. Proceedings of the National Academy of Sciences of the USA, 35, 73–79.CrossRefGoogle Scholar
Kettlewell, H. B. D. (1955). Selection experiments on industrial melanism in the Lepidoptera. Heredity, 9, 323–342.CrossRefGoogle Scholar
Kettlewell, H. B. D. (1956). Further selection experiments on industrial melanism in the Lepidoptera. Heredity, 10, 287–301.CrossRefGoogle Scholar
Kimura, M. (1968). Evolutionary rate at the molecular level. Nature, 217, 624–626.CrossRefGoogle ScholarPubMed
King, J. L., and Jukes, T. H. (1969). Non-Darwinian evolution. Science, 164, 788–798.CrossRefGoogle ScholarPubMed
Kitcher, P. (1984). 1953 and all that. A tale of two sciences. The Philosophical Review, 93(3), 335–373.CrossRefGoogle ScholarPubMed
Kohler, R. E. (1994). Lords of the Fly. Drosophila Genetics and Experimental Practice. Chicago, IL: University of Chicago Press.Google Scholar
Kostoff, D. (1930). Discoid structure of the spireme. The Journal of Heredity, 21(7), 323–324.CrossRefGoogle Scholar
Kottler, M. J. (1979). Hugo de Vries and the rediscovery of Mendel's laws. Annals of Science, 36, 517–538.CrossRefGoogle Scholar
Kreitman, M. (1983). Nucleotide polymorphism at the alcohol dehydrogenase locus of Drosophila melanogaster. Nature, 304, 412–417.CrossRefGoogle ScholarPubMed
Kreitman, M., and Antezana, M. (2000). The population and evolutionary genetics of codon bias. In Singh, R. S., Lewontin, R. C., and Krimbas, C. B. (eds.), Evolutionary Genetics: From Molecules to Morphology (Vol. I, pp. 82–101). Cambridge: Cambridge University Press.Google Scholar
Kühn, A. (1941). Über eine Gen-Wirkkette der Pigmentbildung bei Insekten. Nachrichten der Akademie der Wissenschaften in Göttingen, Mathematisch-Physkalische Klasse, 1941, 231–261.Google Scholar
Laubichler, M., and Rheinberger, H.-J. (2004). Alfred Kühn (1885–1968) and developmental evolution. Journal of Experimental Zoology, 302B, 103–110.CrossRefGoogle Scholar
Lea, D. E. (1962). Actions of Radiations on Living Cells. Cambridge: Cambridge University Press.Google Scholar
Lederberg, J. (1947). Gene recombination and linked segregation in Escherichia coli. Genetics, 32, 505–525.Google Scholar
Lederberg, J. (ed.). (1990). The Excitement and Fascination of Science: Reflections by Eminent Scientists (Vol. III, part 1). Palo Alto, CA: Annual Reviews.Google Scholar
Lederberg, J. (1994). The transformation of genetics by DNA: An anniversary celebration of Avery, MacLeod, and McCarty (1944). Genetics, 136(2), 423–426.Google Scholar
Lederberg, J. (1996). Genetic recombination in Escherichia coli: Disputation at Cold Spring Harbor, 1946–1996. Genetics, 144(2), 439–443.Google ScholarPubMed
Lederberg, J., and Lederberg, E. M. (1952). Replica plating and indirect selection of bacterial mutants. Journal of Bacteriology, 63, 399–406.Google ScholarPubMed
Lederberg, J., Lederberg, E. M., Zinder, N. D., and Lively, E. R. (1951). Recombination analysis of bacterial heredity. Cold Spring Harbor Symposia on Quantitative Biology, 16, 413–443.CrossRefGoogle ScholarPubMed
Lederberg, J., and Tatum, E. L. (1946). Gene recombination in Escherichia coli. Nature, 158, 558.CrossRefGoogle ScholarPubMed
Lenoir, T. (1982). The Strategy of Life. Chicago, IL: University of Chicago Press.Google Scholar
Lerner, I. M. (1954). Genetic Homeostasis. New York: John Wiley.Google Scholar
Levine, M. (1988). Molecular analysis of dorsal-ventral polarity in Drosophila. Cell, 52, 785–786.CrossRefGoogle ScholarPubMed
Lewis, E. B. (1951). Pseudoallelism and gene evolution. Cold Spring Harbor Symposia on Quantitative Biology, 16, 159–172.CrossRefGoogle ScholarPubMed
Lewis, E. B. (1963). Genes and developmental pathways. American Zoologist, 3, 33–56.CrossRefGoogle Scholar
Lewis, E. B. (1967). Genes and gene complexes. In Brink, R. A. (ed.), Heritage from Mendel (pp. 17–47). Madison, WI: University of Wisconsin Press.Google Scholar
Lewis, E. B. (1978). A gene complex controlling segmentation in Drosophila. Nature, 276, 565–570.CrossRefGoogle ScholarPubMed
Lewis, E. B. (1998). Antonio Garcia-Bellido in Caltech. International Journal of Developmental Biology, 42, 523–524.Google Scholar
Lewontin, R. C. (1974). The Genetic Basis of Evolutionary Change. New York: Columbia University Press.Google Scholar
Lewontin, R. C. (1991). Twenty-five years ago in Genetics: Electrophoresis in the development of evolutionary genetics: Milestone or millstone?Genetics, 128(4), 657–662.Google ScholarPubMed
Lewontin, R. C. (1992). Genotype and phenotype. In Keller, E. F. and Lloyd, E. A. (eds.), Keywords in Evolutionary Biology (pp. 137–144). Cambridge, MA: Harvard University Press.Google Scholar
Lewontin, R. C., and Berlan, J.-P. (1986). Technology, research, and the penetration of capital: The case of U.S. agriculture. Monthly Review, 38, 21–34.Google Scholar
Lewontin, R. C., and Berlan, J.-P. (1990). The political economy of agricultural research: The case of hybrid corn. In Carroll, C. R., Vandermer, J. H. and Rosset, P. (eds.), Agroecology (pp. 613–628). New York: McGraw-Hill.Google Scholar
Lewontin, R. C., and Hubby, J. L. (1966). A molecular approach to the study of genic heterozygosity in natural populations. II. Amount of variation and degree of heterozygosity in natural populations of Drosophila pseudoobscura. Genetics, 54, 595–609.Google Scholar
Lewontin, R. C., and Kojima, K.-I. (1960). The evolutionary dynamics of complex polymorphism. Evolution, 14(4), 458–472.Google Scholar
Lewontin, R. C., Moore, J. A., Provine, W. B., and Wallace, B. (1981). Dobzhansky's Genetics of Natural Populations I-XLIII. New York: Columbia University Press.Google Scholar
Lewontin, R. C., Paul, D., Beatty, J., and Krimbas, C. B. (2000). Interview of R. C. Lewontin. In R. S. Singh, Krimbas, C. B., Paul, D. B., and Beatty, J. (eds.), Thinking About Evolution: Historical, Philosophical and Political Perspectives (Vol. II, pp. 22–61). Cambridge: Cambridge University Press.Google Scholar
Liebe, B., Alsheimer, M., Höög, C., Benavente, R., and Scherthan, H. (2004). Telomere attachment, meiotic chromosome condensation, pairing, and bouquet stage duration are modified in spermatocytes lacking axial elements. Molecular Biology of the Cell, 15, 827–837.CrossRefGoogle ScholarPubMed
Lifschytz, E., and Falk, R. (1969a). Fine structure analysis of a chromosome segment in Drosophila melanogaster. Analysis of ethyl methanesulphonate-induced lethals. Mutation Research, 8, 147–155.CrossRefGoogle ScholarPubMed
Lifschytz, E., and Falk, R. (1969b). A genetic analysis of the Killer-prune (K-pn) locus of Drosophila melanogaster. Genetics, 62, 353–358.Google ScholarPubMed
Lindegren, C. C. (1953). Gene conversion in Saccharomyces. Journal of Genetics, 51, 625–637.CrossRefGoogle Scholar
Lindsley, D. L., and Grell, E. H. (1968). Genetic Variations of Drosophila melanogaster. Washington, DC: Carnegie Institute of Washington Publication No. 627.Google Scholar
Loeb, J. (1912). The Mechanistic Conception of Life: Biological Essays. Chicago, IL: The University of Chicago Press.CrossRefGoogle Scholar
Lovejoy, A. O. (1936 [1950]). The Great Chain of Being. Cambridge, MA: Harvard University Press.Google Scholar
Lüning, K. G. (1952a). Studies on the origin of apparent gene mutations in Drosophila melanogaster. Acta Zoologica, 33, 193–207.CrossRefGoogle Scholar
Lüning, K. G. (1952b). X-ray induced chromosome breaks in Drosophila melanogaster. Hereditas, 38, 321–338.CrossRefGoogle Scholar
Luria, S. E., and Delbrück, M. (1943). Mutations of bacteria from virus sensitivity to virus resistance. Genetics, 28(6), 491–511.Google ScholarPubMed
MacCorquodale, K., and Meehl, P. E. (1948). On distinction between hypothetical constructs and intervening variables. Psychological Review, 55, 95–107.CrossRefGoogle ScholarPubMed
Malling, H. V., and Serres, F. J. (1968). Identification of genetic alterations induced by ethyl methanesulfonate in Neurospora crassa. Mutation Research, 6, 181–193.CrossRefGoogle ScholarPubMed
Marmur, J., and Lane, D. (1960). Strand separation and specific recombination in deoxyribonucleic acids: biological studies. Proceedings of the National Academy of Sciences of the USA, 46(4), 453–461.CrossRefGoogle ScholarPubMed
Mattick, J. S. (2003). Challenging the dogma: the hidden layer of non-protein-coding RNAs in complex organisms. BioEssays, 25(10), 930–939.CrossRefGoogle ScholarPubMed
Mattick, J. S. (2005). The functional genomics of noncoding RNA. Science, 309, 1527–1528.CrossRefGoogle ScholarPubMed
Mayr, E. (1987). The ontological status of species: scientific progress and philosophical terminology. Biology and Philosophy, 2(2), 145–166.Google Scholar
McCarthy, B. J., and Bolton, E. T. (1963). An approach to the measurement of genetic relatedness among organisms. Proceedings of the National Academy of Sciences of the USA, 50(1), 156–164.CrossRefGoogle ScholarPubMed
McCarthy, B. J., and Hoyer, B. H. (1964). Identity of DNA and diversity of messenger RNA molecules in normal mouse tissues. Proceedings of the National Academy of Sciences of the USA, 52(4), 915–922.CrossRefGoogle ScholarPubMed
McClintock, B. (1938). The production of homozygous deficient tissues with mutant characteristics by means of the aberrant mitotic behavior of ring-shaped chromosomes. Genetics, 23, 315–376.Google ScholarPubMed
McClintock, B. (1951). Chromosome organization and genic expression. Cold Spring Harbor Symposia on Quantitative Biology, 16, 13–47.CrossRefGoogle ScholarPubMed
McGinnis, W. (1994). A century of homeosis; A decade of homeoboxes. Genetics, 137(3), 607–611.Google ScholarPubMed
McGinnis, W., Garber, R., Wirz, J., Kuroiwa, A., and Gehring, W. (1984). A homologous protein-coding sequence in Drosophila homeotic genes and its conservation in other metazoans. Cell, 37, 403–408.CrossRefGoogle ScholarPubMed
McLaughlin, P. (2002). Naming Biology. Journal of the History of Biology, 35(1), 1–4.CrossRefGoogle ScholarPubMed
Meijer, O. G. (1985). Hugo de Vries no Mendelian?Annals of Science, 42, 189–232.CrossRefGoogle Scholar
Mendel, G. (1866). Versuche über Pflanzenhybriden. Verhandlungen Naturforscher Verein, Brunn, 4, 3–47.Google Scholar
Meselson, M., and Stahl, F. W. (1958a). The replication of DNA. Cold Spring Harbor Symposia on Quantitative Biology, 23, 9–12.CrossRefGoogle ScholarPubMed
Meselson, M., and Stahl, F. W. (1958b). The replication of DNA in Escherichia coli. Proceedings of the National Academy of Sciences of the USA, 44, 671–682.CrossRefGoogle ScholarPubMed
Meselson, M. S., and Radding, C. M. (1975). A general model for genetic recombination. Proceedings of the National Academy of Sciences of the USA, 72, 358–361.CrossRefGoogle ScholarPubMed
Metz, C. W. (1935). Structure of the salivary gland chromosomes in Sciara. The Journal of Heredity, 26, 177–188.CrossRefGoogle Scholar
Mitchell, M. B. (1955). Aberrant recombination of pyridoxine mutants of Neurospora. Proceedings of the National Academy of Sciencess of the USA, 41, 215–220.CrossRefGoogle ScholarPubMed
Moberg, C. L. (2005). René Dubos: Friend of the Good Earth. Microbiologist, Medical Scientist, Environmentalist. Washington, DC: ASM Press.Google Scholar
Mohr, O. L. (1923). Das Defizienz-Phänomen bei Drosophila melanogaster. Zeitschrift für induktive Abstammungs- und Vererbungslehre, 30, 279–283.Google Scholar
Mohr, O. L. (1924). A genetic and cytological analysis of a section deficiency involving four units of the X-chromosome in Drosophila melanogaster. Zeitschrift für induktive Abstammungs- und Vererbungslehre, 32, 108–232.Google Scholar
Monaghan, F. V., and Corcos, A. F. (1990). The real objective of Mendel's paper. Biology and Philosophy, 5, 267–292.CrossRefGoogle Scholar
Monod, J. (1972). Chance and Necessity. New York: Vintage Books.Google Scholar
Morange, M. (1994). Histoire de la biologie moléculaire. Paris: La Découverte.Google Scholar
Morange, M. (1998). A History of Molecular Biology (M. Cobb, Trans.). Cambridge, MA: Harvard University Press.Google Scholar
Morgan, L. V. (1922). Non-criss-cross inheritance in Drosophila melanogaster. Biological Bulletin, Woods Hole, 42, 267–274.CrossRefGoogle Scholar
Morgan, T. H. (1910a). Chromosomes and heredity. The American Naturalist, 44, 449–498.CrossRefGoogle Scholar
Morgan, T. H. (1910b). Sex limited inheritance in Drosophila. Science, 32, 120–122.CrossRefGoogle ScholarPubMed
Morgan, T. H. (1911). Random segregation versus coupling in Mendelian inheritance. Science, 34(873), 384.CrossRefGoogle ScholarPubMed
Morgan, T. H. (1912). A modification of the sex ratio, and of other ratios, in Drosophila through linkage. Zeitschrift für induktive Abstammungs- und Vererbungslehre, 7, 323–345.Google Scholar
Morgan, T. H. (1913a). Factors and unit characters in Mendelian heredity. The American Naturalist, 47(553), 5–16.CrossRefGoogle Scholar
Morgan, T. H. (1913b). Heredity and Sex. New York: Columbia University Press.Google Scholar
Morgan, T. H. (1913c). Simplicity versus adequacy in Mendelian formulae. The American Naturalist, 47(558), 372–374.CrossRefGoogle Scholar
Morgan, T. H. (1919). The Physical Basis of Heredity. Philadelphia and London: Lippincott.CrossRefGoogle Scholar
Morgan, T. H. (1934a). Embryology and Genetics. New York: Columbia University Press.Google Scholar
Morgan, T. H. (1934b). The relation of genetics to physiology and medicine. In Nobel Lectures, Physiology or Medicine 1922–1941. Amsterdam: Elsevier, 1965.Google Scholar
Morgan, T. H., Sturtevant, A. H., Muller, H., and Bridges, C. B. (1915). The Mechanism of Mendelian Heredity. New York: Henry Holt and Company.CrossRefGoogle Scholar
Moss, L. (2003). What Genes Can't Do. Cambridge, MA: The MIT Press.Google Scholar
Muller, H. J. (1914a). The bearing of the selection experiments of Castle and Philips on the variability of Genes. The American Naturalist, 48, 567–576.CrossRefGoogle Scholar
Muller, H. J. (1914b). A gene for the fourth chromosome of Drosophila. The Journal of Experimental Zoology, 17, 325–336.CrossRefGoogle Scholar
Muller, H. J. (1916). The mechanism of crossing over. The American Naturalist, 50(592–595), 193–221, 284–305, 350–366, 421–434.CrossRefGoogle Scholar
Muller, H. J. (1918). Genetic variability, twin hybrids and constant hybrids, in a case of balanced lethal factors. Genetics, 3, 422–499.Google Scholar
Muller, H. J. (1920). Are the factors of heredity arranged in a line?The American Naturalist, 54(631), 97–121.CrossRefGoogle Scholar
Muller, H. J. (1922). Variation due to change in the individual gene. The American Naturalist, 56, 32–50.CrossRefGoogle Scholar
Muller, H. J. (1927a). Artificial transmutation of the gene. Science, 66, 84–87.CrossRefGoogle ScholarPubMed
Muller, H. J. (1927b). Quantitative methods in gene research. The American Naturalist, 61, 407–419.CrossRefGoogle Scholar
Muller, H. J. (1928). The measurement of gene mutation rate in Drosophila, its high variability, and its dependence upon temperature. Genetics, 13, 279–357.Google ScholarPubMed
Muller, H. J. (1929). The gene as the basis of life. Proceedings of the International Congress of Plant Sciences, Ithaca 1926, 1, 897–921.Google Scholar
Muller, H. J. (1932). Further studies on the nature and causes of gene mutations. Proceedings of the 6th International Congress of Genetics, Ithaca, 1, 213–255.Google Scholar
Muller, H. J. (1940). An analysis of the process of structural change in chromosomes of Drosophila. Journal of Genetics, 40, 1–66.CrossRefGoogle Scholar
Muller, H. J. (1947). The gene. Proceedings of the Royal Society of London, series B, 134, 1–37.CrossRefGoogle ScholarPubMed
Muller, H. J. (1950a). Evidence of the precision of genetic adaptation. The Harvey Lectures, 43, 165–229.Google Scholar
Muller, H. J. (1950b). Our load of mutations. American Journal of Human Genetics, 2(2), 111–176.Google ScholarPubMed
Muller, H. J. (1954). The manner of production of mutations by radiation. In Hollaender, A. (ed.), Radiation Biology (Vol. I, pp. 475–626). New York: McGraw Hill.Google Scholar
Muller, H. J. (1956). On the relation between chromosome changes and gene mutations. Brookhaven Symposia in Biology, 8 (591), 126–147.Google Scholar
Muller, H. J., and Falk, R. (1961). Are induced mutations in Drosophila overdominant? I. Experimental design. Genetics, 46(7), 727–735.Google ScholarPubMed
Muller, H. J., and Herskowitz, I. H. (1954). Concerning the healing of chromosome ends produced by breakage in Drosophila melanogaster. The American Naturalist, 88, 177–208.CrossRefGoogle Scholar
Muller, H. J., and Painter, T. S. (1929). The cytological expression of changes in gene alignment produced by X-rays in Drosophila. The American Naturalist, 63(686), 193–200.CrossRefGoogle Scholar
Müller-Hill, B. (1996). The lac Operon: A Short History of a Genetic Paradigm. Berlin, New York: Walter Gruyer.CrossRefGoogle Scholar
Müller-Wille, S. (1998). “Reducing varieties to their species”. The Linnaean research program and its significance for modern biology. In Ruis, E. M. and Corrales, M. de P. P. (eds.), Carl Linnaeus and Enlightened Science in Spain (pp. 113–126). Madrid: Fundacion Berndt Wistedt (Spain and Sweden: Encounters throughout History).Google Scholar
Müller-Wille, S., and Orel, V. (2007). From Linnaean species to Mendelian factors: Elements of hybridism, 1751–1870. Annals of Science, 64(2), 171–215.CrossRefGoogle Scholar
Neel, J. V., and Schull, W. J. (1954). Human Heredity. Chicago, IL: The University of Chicago Press.Google Scholar
Nelkin, D., and Lindee, M. S. (1995). The DNA Mystique. The Gene as a Cultural Icon. New York: Freeman.Google Scholar
Neumann-Held, E. M., and Rehmann-Sutter, C. (eds.). (2006). Genes in Development: Re-reading the Molecular Paradigm. Durham: Duke University Press.CrossRefGoogle Scholar
Newcombe, H. B. (1949). The origin of bacterial variants. Nature, 164, 150–151.CrossRefGoogle ScholarPubMed
Nicklas, R. B. (1997). How cells get the right chromosomes. Science, 275(5300), 632–637.CrossRefGoogle ScholarPubMed
Nicklas, R. B., and Koch, C. A. (1969). Chromosome micromanipulation: 3. Spindle Fiber Tension and the Reorientation of Mal-Oriented Chromosomes. Journal of Cell Biology, 43(1), 40–50.CrossRefGoogle ScholarPubMed
Nöthiger, R. (2002). Ernst Hadorn, a Pioneer of Developmental Genetics. International Journal of Developmental Biology, 46, 23–27.Google Scholar
Novitski, C. E. (2004). Revision of Fisher's analysis of Mendel's garden peas experiments. Genetics, 166(3), 1139–1140.CrossRefGoogle Scholar
Novitski, E. (2004). On Fisher's criticism of Mendel's results with the garden pea. Genetics, 166(3), 1133–1136.CrossRefGoogle ScholarPubMed
Novitski, E., and Blixt, S. (1978). Mendel, linkage and synteny. BioScience, 28(1), 34–35.CrossRefGoogle ScholarPubMed
Nüsslein-Volhard, C. (1977). Genetic analysis of pattern-form organization in the embryo of Drosophila melanogaster. Characterization of the maternal effect mutant bicaudal. Roux's Archiv für Entwicklungsmechanik, 183, 249–268.Google Scholar
Nüsslein-Volhard, C., and Wieschaus, E. (1980). Mutations affecting segment number and polarity in Drosophila. Nature, 287, 795–801.CrossRefGoogle ScholarPubMed
Ohno, S. (1970). Evolution by Gene Duplication. Berlin: Springer.CrossRefGoogle Scholar
Okazaki, R., Okazaki, T., Sakabe, K., Sugimoto, K., and Sugino, A. (1968). Mechanism of DNA chain growth. I. Possible discontinuity and unusual secondary structure of newly synthesized chains. Proceedings of the National Academy of Sciences of the USA, 59, 598–605.CrossRefGoogle ScholarPubMed
Olby, R. C. (1974). The Path to the Double Helix: The Discovery of DNA. London: Macmillan.Google Scholar
Olby, R. C. (1979). Mendel no Mendelian?History of Science, 17, 53–72.CrossRefGoogle Scholar
Olby, R. C. (1985). Origins of Mendelism (2nd edn.). Chicago, IL: University of Chicago Press.Google Scholar
Olby, R. C. (1987). William Bateson's introduction of Mendelism to England: A reassessment. British Journal for the History Science, 20, 399–420.CrossRefGoogle ScholarPubMed
Olby, R. C. (1990). The molecular revolution in biology. In Olby, R. C., Cantor, G. N., Christie, J. R. R. and Hodge, M. J. S. (eds.), Companion to the History of Modern Science (pp. 503–520). London: Routledge.Google Scholar
Oliver, B., and Leblanc, B. (2003). How many genes in a genome?Genome Biology, 5(1), 204.1–3.CrossRefGoogle ScholarPubMed
Orel, V. (1996). Gregor Mendel: The First Geneticist (S. Finn, Trans.). Oxford: Oxford University Press.Google Scholar
Orel, V. (2005). Contested memory: debates over the nature of Mendel's paradigm. Hereditas, 142, 98–102.CrossRefGoogle ScholarPubMed
Orel, V. (personal communication a). Science studies and nature of Mendel's paradigm.
Orel, V. (unpubl. ms. b). The useful research question of heredity in the background of the origin of heredity and evolution.
Orel, V., and Hartl, D. (1994). Controversies in the interpretation of Mendel's discovery. History and Philosophy of the Life Sciences, 16, 423–464.Google Scholar
Orel, V., and Wood, R. J. (1998). Empirical genetic laws published in Brno before Mendel was born. Journal of Heredity, 89(1), 79–82.CrossRefGoogle Scholar
Orr-Weaver, T. L., Szostak, J. W., and Rothstein, R. J. (1981). Yeast transformation: A model system for the study of recombination. Proceedings of the National Academy of Sciences of the USA, 78, 6354–6358.CrossRefGoogle Scholar
Orr, H. A. (1991). A test of Fisher's theory of dominance. Proceedings of the National Academy of Sciences of the USA, 88, 11413–11415.CrossRefGoogle ScholarPubMed
Oyama, S. (2000). The Ontogeny of Information: Development Systems and Evolution (2nd edn.). Durham: Duke University Press.CrossRefGoogle Scholar
Oyama, S., Griffiths, P. E., and Gray, R. D. (eds.). (2001). Cycles of Contingency: Developmental Systems and Evolution. Cambridge, MA: MIT Press.Google Scholar
Painter, T. S. (1934a). A new method for the study of chromosome aberrations and the plotting of chromosome maps in Drosophila melanogaster. Genetics, 19, 175–188.Google Scholar
Painter, T. S. (1934b). Salivary chromosomes and the attack on the gene. The Journal of Heredity, 25, 465–476.CrossRefGoogle Scholar
Panshin, I. B. (1936). A demonstration of the specific nature of position effect. Compts Rendus (Doklady) Academie Science U.R.S.S., N.S., 1(10), 83–86.Google Scholar
Panshin, I. B. (1938). The cytogenetic nature of the position effect of the genes white (mottled) and cubitus interuptus (translated from Russian). Biologicheskii zhurnal, 7(4), 837–865.Google Scholar
Pardee, A. B., Jacob, F., and Monod, J. (1959). The genetic control and cytoplasmic expression of inducibility in the synthesis of β-galactosidase by E. coli. Journal of Molecular Biology, 1, 165–178.CrossRefGoogle Scholar
Pardue, M. L., and Gall, J. G. (1970). Chromosomal localization of mouse satellite DNA. Science, 168, 1356–1358.CrossRefGoogle ScholarPubMed
Parnas, O. S. (2006). On the shoulders of generations: The new epistemology of heredity in the nineteenth century. In Müller-Wille, S. and Rheinberger, H.-J. (eds.), Heredity Produced: At the Crossroads of Biology, Politics, and Culture 1500–1870 (Vol. I, pp. 315–346). Cambridge, MA: MIT Press.Google Scholar
Paul, D. (1992). Heterosis. In Keller, E. Fox and Lloyd, E. A. (eds.), Keywords in Evolutionary Biology (pp. 167–169). Cambridge, MA: Harvard University Press.Google Scholar
Pearson, H. (2006). What is a gene?Nature, 441, 399–401.CrossRefGoogle ScholarPubMed
Pearson, K. (1900). The Grammar of Science (2nd edn.). London: Macmillan.Google Scholar
Peaslee, M. H., and Orel, V. (2007). The evolutionary ideas of F. M. (Ladimir) Klácel, teacher of Gregor Mendel. Biomedical Papers of the Medical Faculty, University Palacky Olomouc Czech Republic, 151(1), 151–156.CrossRefGoogle Scholar
Pick, D. (1989). Faces of Degeneration. A European Disorder, c. 1848 – c. 1918. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Plunkett, C. R. (1932). Temperature as a tool of research in phenogenetics: Methods and results. Proceedings of the 6th International Congress of Genetics, Ithaca, 2, 158–169.Google Scholar
Polanyi, M. (1968). Life's irreducible structure. Science, 160, 1308–1312.CrossRefGoogle ScholarPubMed
Pontecorvo, G. (1952). Genetic formulation of gene structure and gene action. Advances in Enzymology, 13, 121–149.Google ScholarPubMed
Pontecorvo, G. (1958). Trends in Genetic Analysis. New York: Columbia University Press.Google Scholar
Portin, P. (1993). The concept of the gene: Short history and the present status. The Quarterly Review of Biology, 68(2), 173–223.CrossRefGoogle ScholarPubMed
Pritchard, R. H. (1955). The linear arrangement of a series of alleles of Aspergillus nidulans. Heredity, 9, 343–371.CrossRefGoogle Scholar
Provine, W. B. (1971). The Origins of Theoretical Population Genetics. Chicago, IL: University of Chicago Press.Google Scholar
Punnett, R. C. (1909). Mendelism (American Edition of 2nd edn., 1907). New York: Wilshire.Google Scholar
Punnett, R. C. (1911). Mendelism (3rd edn.). London: Macmillan.CrossRefGoogle Scholar
Punnett, R. C. (1913). Reduplication series in sweet peas. Journal of Genetics, 3(2), 77–103.CrossRefGoogle Scholar
Punnett, R. C. (1950). Early days of genetics. Heredity, 4(1), 1–10.CrossRefGoogle Scholar
Raff, R. A., and Kaufman, T. C. (1983). Embryos, Genes, and Evolution: The Developmental-Genetic Basis of Evolutionary Change. New York: Macmillan.Google Scholar
Raffel, D., and Muller, H. J. (1940). Position effect and gene divisibility considered in connection with three strikingly similar scute mutations. Genetics, 25, 541–583.Google ScholarPubMed
Rheinberger, H.-J. (1997). Toward a History of Epistemic Things: Synthesizing Proteins in the Test Tube. Stanford, CA: Stanford University Press.Google Scholar
Rheinberger, H.-J. (2000). Ephestia: The experimental design of Alfred Kühn's physiological developmental genetics. Journal of the History of Biology, 33(3), 535–576.CrossRefGoogle ScholarPubMed
Rheinberger, H.-J. (2006). Epistemologie des Konkreten. Studies zur Geschichte der modernen Biologie. Frankfurt am Mein: Suhrkamp.Google Scholar
Rieger, R., Michaelis, A., and Green, M. M. (1976). Glossary of Genetics and Cytogenetics (4th edn.). Berlin: Springer.CrossRefGoogle Scholar
Roll-Hansen, N. (1978). The genotype theory of Wilhelm Johannsen and its relation to plant breeding and the study of evolution. Centaurus, 22, 201–235.CrossRefGoogle Scholar
Rosenberg, A. (1979). From reductionism to instrumentalism? In Ruse, M. (ed.), What the Philosophy of Biology Is: Essays Dedicated to David Hull (pp. 245–262). Dordrecht: Kluwer.Google Scholar
Rosenberg, A. (1985). The Structure of Biological Science. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Roux, W. (1894). The problems, methods, and scope of developmental mechanics. In Maienschein, J. (ed.), Defending Biology: Lectures from the 1890s (pp. 104–148). Cambridge, MA: Harvard University Press.Google Scholar
Sandler, I., and Sandler, L. (1985). A conceptual ambiguity that contributed to the neglect of Mendel's paper. History and Philosophy of the Life Sciences, 7, 3–70.Google ScholarPubMed
Sapp, J. (1983). The struggle for authority in the field of heredity, 1900–1932: New perspectives on the rise of genetics. Journal of the History of Biology, 16, 311–342.CrossRefGoogle ScholarPubMed
Sapp, J. (1990). Where the Truth Lies: Franz Moewus and the Origins of Molecular Biology. Cambridge: Cambridge University Press.Google Scholar
Sarkar, S. (1998). Genetics and Reductionism. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Sarkar, S. (1999). From the Reaktionsnorm to the adaptive norm: The norm of reaction, 1909–1960. Biology and Philosophy, 14(2), 235–252.CrossRefGoogle Scholar
Schaffner, K. F. (1969). The Watson-Crick model and reductionism. British Journal for the Philosophy of Science, 20, 325–348.CrossRefGoogle Scholar
Schaffner, K. F. (1976). Reduction in biology: prospects and problems. In Cohen, R. S., Hooker, C. A., Pearce, G., Michalos, A. C. and Evra, J. W. (eds.), Proceedings of the 1974 Biennial Meeting of the Philosophy of Science Association (pp. 613–632). Dordrecht: Reidel.Google Scholar
Schaffner, K. F. (1993). Discovery and Explanation in Biology and Medicine. Chicago, IL: University of Chicago Press.Google Scholar
Scherthan, H. (2001). A bouquet makes ends meet. Nature Review: Molecular Cell Biology, 2, 621–627.CrossRefGoogle ScholarPubMed
Scherthan, H. (2007). Telomere attachment and clustering during meiosis. Cellular and Molecular Life Sciences, 64(2), 117–124.CrossRefGoogle ScholarPubMed
Schmitt, J., Benavente, R., Hodzic, D., Höög, C., Stewart, C. L., and Alsheimer, M. (2007). Transmembrane protein Sun2 is involved in tethering mammalian meiotic telomeres to the nuclear envelope. Proceedings of the National Academy of Sciences of the USA, 104(18), 7426–7431.CrossRefGoogle ScholarPubMed
Schrödinger, E. (1962 [1944]). What Is Life? The Physical Aspect of the Living Cell. Cambridge: Cambridge University Press.Google Scholar
Schwartz, S. (1998). The significance of the trait in genetics, 1900–1945. Unpublished Ph.D. Dissertation, The Hebrew University of Jerusalem.Google Scholar
Schwartz, S. (2000). The differential concept of the gene. In Beurton, P. J., Falk, R. and Rheinberger, H.-J. (eds.), The Concept of the Gene in Development and Evolution: Historical and Epistemological Perspectives (pp. 26–39). Cambridge: Cambridge University Press.Google Scholar
Schwartz, S. (2002). Characters as units and the case of the presence and absence hypothesis. Biology and Philosophy, 17(3), 369–388.CrossRefGoogle Scholar
Selzer, J. (ed.). (1993). Understanding Scientific Prose. Madison, WI: University of Wisconsin Press.Google Scholar
Setlow, R. B., and Carrier, W. L. (1964). The disappearance of thymine dimers from DNA: An error-correcting mechanism. Proceedings of the National Academy of Sciences of the USA, 51, 226–231.CrossRefGoogle Scholar
Shull, G. H. (1909). The “presence and absence” hypothesis. The American Naturalist, 43, 410–419.CrossRefGoogle Scholar
Simpson, G. G. (1964). This View of Life. New York: Harcourt, Brace and World.Google Scholar
Sinnott, E. W., Dunn, L. C., and Dobzhansky, T. (1950). Principles of Genetics (4th edn.). New York: McGraw-Hill.Google Scholar
Sinnott, E. W., Dunn, L. C., and Dobzhansky, T. (1958). Principles of Genetics (5th edn.). New York: McGraw-Hill.Google Scholar
Sirks, M. J., and Zirkle, C. (1964). The Evolution of Biology. New York: The Ronald Press Company.Google Scholar
Sloan, P. R. (1976). The Buffon-Linnaeus controversy. Isis, 67(3), 358–375.CrossRefGoogle ScholarPubMed
Snyder, M., and Gerstein, M. (2003). Defining genes in the genomics era. Science, 300(5617), 258–260.CrossRefGoogle Scholar
Sober, E. (1984). The Nature of Selection. Cambridge, MA: The MIT Press.Google Scholar
Sober, E., and Wilson, D. S. (1998). Unto Others: The Evolution and Psychology of Unselfish Behavior. Cambridge, MA: Harvard University Press.Google Scholar
Srb, A. M., and Owen, R. D. (1953). General Genetics. San Francisco: Freeman.Google Scholar
Stadler, D. (1997). Ultraviolet-induced mutation and the chemical nature of the gene. Genetics, 145(4), 863–865.Google Scholar
Stadler, L. J. (1928a). Genetic effects of X-rays in maize. Proceedings of the National Academy of Sciences of the USA, 14, 69–75.CrossRefGoogle ScholarPubMed
Stadler, L. J. (1928b). Mutations in barley induced by X-rays and radium. Science, 68, 186–187.CrossRefGoogle ScholarPubMed
Stadler, L. J. (1954). The gene. Science, 120, 811–819.CrossRefGoogle ScholarPubMed
Stadler, L. J., and Sprague, G. F. (1937). Contrasts in the genetic effects of ultra-violet treatment. Science, 85, 57–58.Google Scholar
Stahl, F. W. (1961). A chain model for chromosomes. Journal de Chemie Physique, 58, 1072–1077.CrossRefGoogle Scholar
Stahl, F. W. (1979). Genetic Recombination: Thinking About It in Phage and Fungi. San Francisco: Freeman.Google Scholar
Stahl, F. W. (1988). A unicorn in the garden. Nature, 335, 112–113.CrossRefGoogle ScholarPubMed
Stahl, F. W. (1994). The Holliday junction on its thirtieth anniversary. Genetics, 138(2), 241–246.Google ScholarPubMed
Stamhuis, I. H., Meijer, O. G., and Zevenhuizen, E. J. A. (1999). Hugo de Vries on heredity, 1889–1903: Statistics, Mendelian laws, pangenes, mutations. Isis, 90(2), 238–267.CrossRefGoogle Scholar
Stent, G. S. (1968). That was the molecular biology that was. Science, 160, 390–395.CrossRefGoogle Scholar
Stent, G. S. (1969). The Coming of the Golden Age: A View of the End of Progress. New York: Natural History Press.Google Scholar
Stent, G. S. (1971). Molecular Genetics: An Introductory Narrative. San Francisco: Freeman.Google Scholar
Sterelny, K., and Griffiths, P. E. (1999). Sex and Death: An Introduction to Philosophy of Biology. Chicago, IL: Chicago University Press.Google Scholar
Stern, C. (1929). Über die additive Wirkung multipler allele. Biologisches Zentralblat, 49, 261–290.Google Scholar
Stern, C. (1931). Zytologisch-genetische Untersuchungen als Beweise fur die Morgansche Theorie des Faktorenaustauschs. Biologisches Zentralblat, 51(10), 547–587.Google Scholar
Stern, C. (1936). Somatic crossing-over and segregation in Drosophila melanogaster. Genetics, 21, 625–730.Google ScholarPubMed
Stern, C. (1943). The Hardy-Weinberg law. Science, 97(2510), 137–138.CrossRefGoogle ScholarPubMed
Stern, C. (1955). Two or three bristles. Science in Progress, 9, 41–84; 327–328.Google ScholarPubMed
Stern, C. (1959). Use of the term “superfemale”. Lancet, 12, 1088.CrossRefGoogle Scholar
Stern, C. (1960a). Dosage compensation – development of a concept and new facts. Canadian Journal of Genetics and Cytology, 2(2), 105–118.CrossRefGoogle Scholar
Stern, C. (1960b). Principles of Human Genetics (2nd edn.). San Francisco, CA: Freeman.Google Scholar
Stern, C., and Sherwood, E. (1966). The Origin of Genetics: A Mendel Sourcebook. San Francisco, CA: Freeman.Google Scholar
Stern, H., and Hotta, Y. (1973). Biochemical control of meiosis. Annual Review of Genetics, 7, 37–66.CrossRefGoogle Scholar
Streisinger, G., Okada, Y., Emrich, J., Newton, J., Tsugita, A., Terzaghi, E., et al. (1967). Frameshift mutations and the genetic code. Cold Spring Harbor Symposia on Quantitative Biology, 31, 77–84.CrossRefGoogle Scholar
Strickberger, M. W. (1976). Genetics (2nd edn.). New York: Macmillan.Google Scholar
Sturtevant, A. H. (1913a). The Himalayan rabbit case, with some considerations on multiple allelomorphs. The American Naturalist, 47, 234–238.CrossRefGoogle Scholar
Sturtevant, A. H. (1913b). The linear arrangement of six sex-linked factors in Drosophila, as shown by their mode of association. Journal of Experimental Zoology, 14, 43–59.CrossRefGoogle Scholar
Sturtevant, A. H. (1914). The reduplication hypothesis as applied to Drosophila. The American Naturalist, 48, 535–549.CrossRefGoogle Scholar
Sturtevant, A. H. (1923). Inheritance of the direction of coiling in Limnaea. Science, 58, 269–270.CrossRefGoogle Scholar
Sturtevant, A. H. (1925). The effects of unequal crossing over at the Bar locus in Drosophila. Genetics, 10, 117–147.Google ScholarPubMed
Sturtevant, A. H. (1929). The claret mutant type of Drosophila simulans: a study of chromosome elimination and cell-lineage. Zeitschrift für Wissenschaftliche Zoology, 135, 323–356.Google Scholar
Sturtevant, A. H. (1932). The use of mosaics in the study of the developmental effects of genes. In Proceedings of the 6th International Congress of Genetics, Ithaca, 1, 304–307.
Sturtevant, A. H. (1965). A Short History of Genetics. New York: Harper and Row.Google Scholar
Sturtevant, A. H., and Beadle, G. W. (1936). The relations of inversions in the X chromosome of Drosophila melanogaster to crossing-over and disjunction. Genetics, 21, 554–604.Google ScholarPubMed
Sturtevant, A. H., and Beadle, G. W. (1962 [1939]). An Introduction to Genetics. New York: Dover.Google Scholar
Sturtevant, A. H., and Dobzhansky, T. (1936). Inversions in the third chromosome of wild races of Drosophila pseudoobscura, and their use in the study of the history of the species. Proceedings of the National Academy of Sciences of the USA, 22, 448–450.CrossRefGoogle Scholar
Sutton, W. (1903). The chromosomes in heredity. Biological Bulletin, 4, 231–251.CrossRefGoogle Scholar
Swift, H. H. (1950). The desoxyribose nucleic acid content of animal nuclei. Physiological Zoology, 23, 169–198.CrossRefGoogle ScholarPubMed
Swinburne, R. G. (1962). The presence-and-absence theory. Annals of Science, 18(3), 131–145.CrossRefGoogle Scholar
Tabery, J. G. (2004). The “evolutionary synthesis” of George Udny Yule. Journal of the History of Biology, 37(1), 73–101.CrossRefGoogle Scholar
Taylor, J. H. (1953). Autoradiographic detection of incorporation of P32 into chromosomes during meiosis and mitosis. Experimental Cell Research, 4, 164–173.CrossRefGoogle Scholar
Taylor, J. H. (1959). The organization and duplication of genetic material. Proceedings of the Tenth International congress of Genetics, Montreal 1958, 1, 63–78.Google Scholar
Taylor, J. H., Woods, P. S., and Hughes, W. I. (1957). The organization and duplication of chromosomes as revealed by autoradiographic studies using tritium-labeled thymidine. Proceedings of the National Academy of Sciences of the USA, 43, 122–128.CrossRefGoogle Scholar
Temin, H. M., and Mizutani, S. (1970). RNA-dependent DNA polymerase in virions of Rous sarcoma virus. Nature, 226, 1211–1213.CrossRefGoogle ScholarPubMed
Theunissen, B. (1994). Closing the door on Hugo de Vries' Mendelism. Annals of Science, 51, 225–248.CrossRefGoogle ScholarPubMed
Thoday, J. M., and Read, J. (1947). Effect of oxygen on the frequency of chromosome aberrations produced by X-rays. Nature, 160, 608.CrossRefGoogle ScholarPubMed
Timofeéff-Ressovsky, H. A., and Timofeéff-Ressovsky, N. W. (1926). Über das phaenotypische Manifestieren des Genotyps. II Über idio-somatische Variationsgruppen beiDrosophila funebris. Wilhelm Roux's Archives of Developmental Biology, 108, 146–170.CrossRefGoogle Scholar
Timofeéff-Ressovsky, N. W., Zimmer, E. G., and Delbrück, M. (1935). Über die Natur der Genmutation und der Genstruktur. Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, 1, 189–245.Google Scholar
Tjio, J. H., and Levan, A. (1956). The chromosome number of man. Hereditas, 42, 1–6.CrossRefGoogle Scholar
Toulmin, S. (1972). Human Understanding. The Collective Use and Evolution of Concepts. Princeton, NJ: Princeton University Press.Google Scholar
Troland, L. T. (1917). Biological enigmas and the theory of enzyme action. The American Naturalist, 51(606), 321–350.CrossRefGoogle Scholar
Tschermak, E. (1950 [1900]). Concerning artificial crossing in Pisum sativum. Genetics, 35 Supplement (The birth of Genetics)(5, part 2), 42–47.Google Scholar
Vega, J. M., and Feldman, M. (1998). Effect of the pairing gene Ph1 on centromere misdivision in common wheat. Genetics, 148, 1285–1294.Google ScholarPubMed
Verdun, R. E., and Karlseder, J. (2007). Replication and protection of telomeres. Nature, 447, 924–931.CrossRefGoogle Scholar
Vogelstein, B., and Kinzler, K. W. (1992). P53 function and disfunction. Cell, 70, 523–526.CrossRefGoogle Scholar
Vogt, O. (1926). Psychiatrisch wichtige Tatsachen der zoologisch-botanischen Systematik. Zeitschrift für die gesamte Neurologie und Psychiatrie, 101, 805–832.CrossRefGoogle Scholar
Volkin, E., and Astrachan, L. (1957). RNA metabolism in T2-infected Escherichia coli. In Glass, B. and McElroy, W. D. (eds.), The Chemical Basis of Heredity (pp. 686–695). Baltimore, MD: Johns Hopkins Press.Google Scholar
Waddington, C. H. (1942). Canalization of development and the inheritance of acquired characters. Nature, 150, 563–565.CrossRefGoogle Scholar
Waddington, C. H. (1957). The Strategy of the Genes. A Discussion of Some Aspects of Theoretical Biology. London: George Allen and Unwin.Google Scholar
Wallace, B. (1958). The average effect of radiation-induced mutations on viability in Drosophila melanogaster. Evolution, 12, 532–556.CrossRefGoogle Scholar
Watson, J. D. (1968). The Double Helix: A Personal Account of the Discovery of the Structure of DNA. New York: Atheneum.Google Scholar
Watson, J. D., and Crick, F. H. C. (1953a). Genetical implications of the structure of deoxyribose nucleic acid. Nature, 171, 964–967.CrossRefGoogle Scholar
Watson, J. D., and Crick, F. H. C. (1953b). Molecular structure of nucleic acids. Nature, 171, 737–738.CrossRefGoogle ScholarPubMed
Watson, J. D., and Crick, F. H. C. (1953c). The structure of DNA. Cold Spring Harbor Symposia on Quantitative Biology, 18, 123–131.CrossRefGoogle ScholarPubMed
Weaver, W. (1970). Molecular biology: Origin of the term. Science, 170(3958), 581–582.CrossRefGoogle ScholarPubMed
Weinberg, R. A. (2007). The Biology of Cancer. New York: Garland.Google Scholar
Weiner, J. (1999). Time, Love, Memory: A Great Biologist and His Quest for the Origins of Behavior. New York: Alfred A. Knopf.Google Scholar
Weinstein, A. (1918). Coincidence of crossing over in Drosophila melanogaster (ampelophila). Genetics, 3, 135–172.Google Scholar
Weinstein, A. (1977). How unknown was Mendel's paper?Journal of the History of Biology, 10(2), 341–364.CrossRefGoogle Scholar
Weinstock, G. M. (2007). ENCODE: More genomic improvement. Genome Research, 17, 667–668.CrossRefGoogle Scholar
Westergaard, M. (1958). The mechanism of sex determination in dioecious plants. Advances in Genetics, 9, 217–281.Google Scholar
White, M. J. D. (1954). Animal Cytology and Evolution (2nd edn.). Cambridge: Cambridge University Press.Google Scholar
White, M. J. D. (1973). The Chromosomes (6th edn.). London: Chapman and Hall.Google Scholar
Whitehouse, H. L. K. (1965). Towards an Understanding of the Mechanism of Heredity. London: Edward Arnold.Google Scholar
Wilkie, A. O. M. (1994). The molecular basis of genetic dominance. Journal of Medical Genetics, 31, 89–98.CrossRefGoogle ScholarPubMed
Wilkins, A. S. (1997). Canalization: a molecular genetic perspective. BioEssays, 19(3), 257–262.CrossRefGoogle ScholarPubMed
Wilkins, A. S. (2002). The Evolution of Developmental Pathways. Sunderland, MA: Sinauer.Google Scholar
Williams, G. C. (1974 [1966]). Adaptation and Natural Selection: A Critique of Some Current Evolutionary Thought. Princeton, NJ: Princeton University Press.Google Scholar
Wilson, D. S. (1983). The group selection controversy: History and current status. Annual Review of Ecology and Systematics, 14, 159–187.CrossRefGoogle Scholar
Wilson, D. S. (1992). Group selection. In Keller, E. F. and Lloyd, E. A. (eds.), Keywords in Evolutionary Biology (pp. 145–148). Cambridge, MA: Harvard University Press.Google Scholar
Wilson, D. S., and Sober, E. (1989). Reviving the superorganism. Journal of Theoretical Biology, 136, 337–356.CrossRefGoogle ScholarPubMed
Wilson, E. B. (1893 [1986]). The mosaic theory of development. In Maienschein, J. (ed.), Defending Biology: Lecture from the 1890s (pp. 67–80). Cambridge, MA: Harvard University Press.Google Scholar
Wilson, E. B. (1896). The Cell in Development and Inheritance. New York: Macmillan.Google Scholar
Wilson, E. B. (1924). The Cell in Development and Inheritance (3rd edn.). New York: Macmillan.Google Scholar
Wilson, E. O. (1975). Sociobiology: The New Synthesis. Cambridge, MA: Harvard University Press.Google Scholar
Winkler, H. (1930). Die Konversion der Gene. Jena: Fischer.Google Scholar
Wollman, E. L., Jacob, F., and Hayes, W. (1956). Conjugation and genetic recombination in Escherichia coli K-12. Cold Spring Harbor Symposia on Quantitative Biology, 21, 141–162.CrossRefGoogle ScholarPubMed
Woltereck, R. (1909). Weitere experimentelle Untersuchungen über Artveränderung, speziell über des Wesen quantitativer Artunterschiede bei Daphnien. Verhandlungen der Deutschen Zoologischen Gesellschaft, 19, 110–173.Google Scholar
Wood, R. J., and Orel, V. (2005). Scientific breeding in central Europe during the early nineteenth century: Background to Mendel's later work. Journal of the History of Biology, 38, 239–272.CrossRefGoogle Scholar
Woodger, J. H. (1967). Biological Principles. London: Routledge and Kegan Paul.Google Scholar
Wright, S. (1929a). The evolution of dominance: Comment on Dr. Fisher's reply. The American Naturalist, 63, 556–561.CrossRefGoogle Scholar
Wright, S. (1929b). Fisher's theory of dominance. The American Naturalist, 63, 274–279.CrossRefGoogle Scholar
Wright, S. (1931). Evolution in Mendelian populations. Genetics, 16, 97–159.Google ScholarPubMed
Wright, S. (1945). Tempo and mode in evolution: A critical review. Ecology, 26, 415–419.CrossRefGoogle Scholar
Wright, S. (1958). Systems of Mating and Other Papers. Ames, IA: The Iowa State College Press.Google Scholar
Wynne-Edwards, V. C. (1962). Animal Dispersion in Relation to Social Behaviour. Edinburgh: Oliver and Boyd.Google Scholar
Yanofsky, C., Carlton, B. C., Guest, J. R., Helinski, D. R., and Henning, U. (1964). On the colinearity of gene structure and protein structure. Proceedings of the National Academy of Sciences of the USA, 51(2), 266–272.CrossRefGoogle ScholarPubMed
Yanofsky, C., Drapeau, G. R., Guest, J. R., and Carlton, B. C. (1967). The complete amino acid sequence of the tryptophan synthetase A protein (α subunit) and its colinear relationship with the genetic map of the A gene. Proceedings of the National Academy of Sciences of the USA, 57(2), 296–298.CrossRefGoogle ScholarPubMed
Yanofsky, C., and Lennox, E. S. (1959). Transduction and recombination study of linkage relationships among the genes controlling tryptophan synthesis in Escherichia coli. Virology, 8, 425–447.CrossRefGoogle ScholarPubMed
Yule, G. U. (1902). Mendel's laws and their probable relations to intra-racial heredity. The New Phytologist, 1(9 and 10), 193–207 and 222–238.CrossRefGoogle Scholar
Yule, G. U. (1903). Professor Johannsen's experiments in heredity: A review. The New Phytologist, 2, 235–242.CrossRefGoogle Scholar
Zacharias, H. (1995). Emil Heitz (1892–1965): Chloroplasts, heterochromatin and polytene chromosomes. Genetics, 141(1), 7–14.Google ScholarPubMed
Zevenhuizen, E. (1998). The hereditary statistics of Hugo de Vries. Acta Botanica Neerlandica, 47(4), 427–463.Google Scholar
Zirkle, C. (1935a). The Beginnings of Plant Hybridization (Morris Arboretum Monographs). Philadelphia, PA: University of Pennsylvania Press.CrossRefGoogle Scholar
Zirkle, C. (1935b). The inheritance of acquired characters and the provisional hypothesis of pangenesis. The American Naturalist, 69, 417–445.CrossRefGoogle Scholar
Zirkle, C. (1945). The early history of the idea of the inheritance of acquired characters and of pangenesis. Transactions of the American Philosophical Society, 35(part II), 91–150.CrossRefGoogle Scholar
Zuckerman, H., and Lederberg, J. (1986). Postmature scientific discovery?Nature, 324(6098), 629–631.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Bibliography
  • Raphael Falk, Hebrew University of Jerusalem
  • Book: Genetic Analysis
  • Online publication: 07 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511581465.030
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • Bibliography
  • Raphael Falk, Hebrew University of Jerusalem
  • Book: Genetic Analysis
  • Online publication: 07 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511581465.030
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • Bibliography
  • Raphael Falk, Hebrew University of Jerusalem
  • Book: Genetic Analysis
  • Online publication: 07 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511581465.030
Available formats
×