Hostname: page-component-848d4c4894-ndmmz Total loading time: 0 Render date: 2024-06-06T19:43:44.784Z Has data issue: false hasContentIssue false

Micro- and macroevolution: Scale and hierarchy in evolutionary biology and paleobiology

Published online by Cambridge University Press:  26 February 2019

David Jablonski*
Affiliation:
Department of Geophysical Sciences, University of Chicago, 5734 South Ellis Avenue, Chicago, Illinois 60637. E-mail: djablons@midway.uchicago.edu

Abstract

The study of evolution has increasingly incorporated considerations of history, scale, and hierarchy, in terms of both the origin of variation and the sorting of that variation. Although the macroevolutionary exploration of developmental genetics has just begun, considerable progress has been made in understanding the origin of evolutionary novelty in terms of the potential for coordinated morphological change and the potential for imposing uneven probabilities on different evolutionary directions. Global or whole-organism heterochrony, local heterochrony (affecting single structures, regions, or organ systems) and heterotopies (changes in the location of developmental events), and epigenetic mechanisms (which help to integrate the developing parts of an organism into a functional whole) together contribute to profound nonlinearities between genetic and morphologic change, by permitting the generation and accommodation of evolutionary novelties without pervasive, coordinated genetic changes; the limits of these developmental processes are poorly understood, however. The discordance across hierarchical levels in the production of evolutionary novelties through time, and among latitudes and environments, is an intriguing paleontological pattern whose explanation is controversial, in part because separating effects of genetics and ecology has proven difficult. At finer scales, species in the fossil record tend to be static over geologic time, although this stasis—to which there are gradualistic exceptions—generally appears to be underlain by extensive, nondirectional change rather than absolute invariance. Only a few studies have met the necessary protocols for the analysis of evolutionary tempo and mode at the species level, and so the distribution of evolutionary patterns among clades, environments, and modes of life remains poorly understood. Sorting among taxa is widely accepted in principle as an evolutionary mechanism, but detailed analyses are scarce; if geographic range or population density can be treated as traits above the organismic level, then the paleontological and mac̀roecological literature abounds in potential raw material for such analyses. Even if taxon sorting operates on traits that are not emergent at the species level, the differential speciation and extinction rates can shape large-scale evolutionary patterns in ways that are not simple extrapolations from short-term evolution at the organismal level. Changes in origination and extinction rates can evidently be mediated by interactions with other clades, although such interactions need to be studied in a geographically explicit fashion before the relative roles of biotic and physical factors can be assessed. Incumbency effects are important at many scales, with the most dramatic manifestation being the postextinction diversifications that follow the removal of incumbents. However, mass extinctions are evolutionarily important not only for the removal of dominant taxa, which can occur according to rules that differ from those operating during times of lower extinction intensity, but also for the dramatic diversifications that follow upon the removal or depletion of incumbents. Mass extinctions do not entirely reset the evolutionary clock, so survivors can exhibit unbroken evolutionary continuity, trends that suffer setbacks but then resume, or failure to participate in the recovery.

Type
Research Article
Copyright
Copyright © 2000 by The Paleontological Society 

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

Literature Cited

Adams, M. D., et al. 2000. The genome sequence of Drosophila melanogaster. Science 287:21852195.Google Scholar
Ahn, D.-G., and Gibson, G. 1999a. Axial variation in the threespine stickleback: relationship to Hox gene expression. Development, Genes and Evolution 209:473481.Google Scholar
Ahn, D.-G., and Gibson, G. 1999b. Expression patterns of the threespine stickleback Hox genes and insights into the evolution of the vertebrate body axis. Development, Genes and Evolution 209:482494.Google Scholar
Akam, M. 1998. Hox genes, homeosis and the evolution of segment identity: no need for hopeless monsters. International Journal of Developmental Biology 42:445451.Google Scholar
Alberch, P., Gould, S. J., Oster, G. F., and Wake, D. B. 1979. Size and shape in ontogeny and phylogeny. Paleobiology 5:296317.Google Scholar
Arneberg, P., Skorping, A., and Read, A. F. 1997. Is population density a species character? Comparative analyses of the nematode parasites of mammals. Oikos 80:289300.Google Scholar
Arnold, H. 1966. Grundsätzliche Schwierigkeiten bei der biostratigraphischen Deutung phyletischer Reihen. Senckenbergiana Lethaea 47:537547.Google Scholar
Arnold, S. J., and Phillips, P. C. 1999. Hierarchical comparison of genetic variance-covariance matrices. II. Coastal-inland divergence in the garter snake, Thamnophis elegans. Evolution 53:15161527.Google Scholar
Arnone, M. I., and Davidson, E. H. 1997. The hardwiring of development: organization and function of genomic regulatory systems. Development 124:18511864.Google Scholar
Aronson, R. B. 1992. Biology of a scale-independent predatorprey interaction. Marine Ecology Progress Series 80:113.Google Scholar
Aronson, R. B. 1994. Scale-independent biological processes in the marine environment. Oceanography and Marine Biology, an Annual Review 32:435460.Google Scholar
Arthur, W. 1988. A theory of the evolution of development. Wiley, Chichester, England.Google Scholar
Arthur, W. 1997. The origin of animal body plans. Cambridge University Press, Cambridge.Google Scholar
Askin, R. A., and Spicer, R. A. 1995. The Late Cretaceous and Cenozoic history of vegetation and climate at northern and southern high latitudes: a comparison. Pp. 156173 in National Research Council Board on Earth Sciences and Resources, Effects of past global change on life. National Academy Press, Washington, D.C. Google Scholar
Atchley, W. R., and Hall, B. K. 1991. A model for development and evolution of complex morphological structures. Biological Reviews 66:101157.Google Scholar
Atchley, W. R., Xu, S. Z., and Vogl, C. 1994. Developmental quantitative genetic models of evolutionary change. Developmental Genetics 15:92103.Google Scholar
Averof, M., and Patel, N. H. 1997. Crustacean appendage evolution associated with changes in Hox gene expression. Nature 388:682686.Google Scholar
Avise, J. 2000. Phylogeography. Harvard University Press, Cambridge.Google Scholar
Bambach, R. K. 1983. Ecospace utilization and guilds in marine communities through the Phanerozoic. Pp. 719746 in Tevesz, M. J. S. and McCall, P. L., eds. Biotic interactions in recent and fossil benthic communities. Plenum, New York. Barnosky, A. D. 1987. Punctuated equilibria and phyletic gradualism: some facts from the Quaternary fossil record. Current Mammalogy 1:109147.Google Scholar
Barraclough, T. G., and Vogler, A. P. 2000. Detecting geographical pattern of speciation from species-level phylogenies. American Naturalist 155:419434.Google Scholar
Bateman, R. M., Crane, P. R., DiMichele, W. A., Kenrick, P. R., Rowe, N. P., Speck, T., and Stein, W. E. 1998. Early evolution of land plants: phylogeny, physiology, and ecology of the primary terrestrial radiation. Annual Review of Ecology and Systematics 29:263292.Google Scholar
Behrensmeyer, A. K., and Hook, R. W. 1992. Paleoenvironmental contexts and taphonomic modes. Pp. 15136 in Behrensmeyer, A. K. et al., eds. Terrestrial ecosystems through time. University of Chicago Press, Chicago.Google Scholar
Behrensmeyer, A. K., Kidwell, S. M., and Gastaldo, R. A. 2000. Taphonomy and paleobiology. Paleobiology 26:103147.Google Scholar
Bell, M. A., Sadagursky, M. S., and Baumgartner, J. V. 1987. Utility of lacustrine deposits for the study of variation within fossil samples. Palaios 2:455466.Google Scholar
Belliure, J., Sorci, G., M⊘ller, A. P., and Cloubert, J. 2000. Dispersal distances predict subspecies richness in birds. Journal of Evolutionary Biology 13:480487.Google Scholar
Belting, H. G., Shashikant, C. S., and Ruddle, F. H. 1998. Modification of expression and cis-regulation of Hoxc8 in the evolution of diverged axial morphology. Proceedings of the National Academy of Sciences USA 95:23552360.Google Scholar
Bennett, K. D. 1997. Evolution and ecology: the pace of life. Cambridge University Press, Cambridge.Google Scholar
Bennett, P. M., and Owens, I. P. F. 1997. Variation in extinction risk among birds: chance or evolutionary predisposition? Proceedings of the Royal Society of London B 264:401408.Google Scholar
Benton, M. J. 1987. Progress and competition in macroevolution. Biological Reviews 62:305338.Google Scholar
Benzie, J. A. H. 1999a. Genetic structure of coral reef organisms: ghosts of dispersal past. American Zoologist 39:131145.Google Scholar
Benzie, J. A. H. 1999b. Major genetic differences between crown-of-thorns starfish (Acanthaster planci) populations in the Indian and Pacific Oceans. Evolution 53:17821795.Google Scholar
Benzie, J. A. H., and Williams, S. T. 1997. Genetic structure of giant clam (Tridacna maxima) populations in the west Pacific is not consistent with dispersal by present-day ocean currents. Evolution 51:768783.Google Scholar
Blackburn, T. M., and Gaston, K. J. 1997. A critical assessment of the form of the interspecific relationship between abundance and body size in animals. Journal of Animal Ecology 66:233249.Google Scholar
Blackburn, T. M., and K. J. 1999. The relationship between animal abundance and body size: a review of the mechanisms. Advances in Ecological Research 28:181210.Google Scholar
Bohonak, A. J. 1999. Dispersal, gene flow, and population structure. Quarterly Review of Biology 74:2145.Google Scholar
Bookstein, F. L. 1987. Random walk and the existence of evolutionary rates. Paleobiology 11:258271.Google Scholar
Bookstein, F. L. 1988. Random walk and the biometrics of morphological characters. Evolutionary Biology 23:369398.Google Scholar
Bradshaw, H. D., Otto, K. G., Frewen, B. E., McKay, J. K., and Schemske, D. W. 1998. Quantitative trait loci affecting differences in floral morphology between two species of monkeyflower (Mimulus). Genetics 149:367382.Google Scholar
Bralower, T. J., and Parrow, M. 1996. Morphometrics of the Paleocene coccolith genera Cruciplacolithus, Chiasmolithus, and Sullivania: a complex evolutionary history. Paleobiology 22:352385.Google Scholar
Brandon, R. N. 1982. The levels of selection. Pp. 315323 in Asquith, P. and Nichols, T., eds. PSA 1982, Vol. 1. Philosophy of Science Association, East Lansing, Mich.Google Scholar
Brandon, R. N. 1988. The levels of selection: a hierarchy of interactors. Pp. 5171 in Plotkin, H., ed. The role of behavior in evolution. MIT Press, Cambridge.Google Scholar
Brown, J. H. 1995. Macroecology. University of Chicago Press, Chicago.Google Scholar
Brown, J. H. 1999. Macroecology: progress and prospect. Oikos 87:314.Google Scholar
Brown, J. H., and Davidson, D. W. 1977. Competition between seed-eating rodents and ants in desert ecosystems. Science 196:880882.Google Scholar
Brown, J. H., Davidson, D. W., and Reichman, O. J. 1979a. An experimental study of competition between seed-eating desert rodents and ants. American Zoologist 19:11291143.Google Scholar
Brown, J. H., Reichman, O. J., and Davidson, D. W. 1979b. Granivory in desert ecosystems. Annual Review of Ecology and Systematics 10:201227.Google Scholar
Brown, J. H., Stevens, G. C., and Kaufman, D. M. 1996. The geographic range: size, shape, boundaries, and internal structure. Annual Review of Ecology and Systematics 27:597623.Google Scholar
Budd, A. F., Johnson, K. G., and Potts, D. C. 1994. Recognizing morphospecies in colonial reef corals. 1. Landmark-based methods. Paleobiology 20:484505 Google Scholar
Burke, A. C. 1989. Development of the turtle carapace: implications for the evolution of a novel Bauplan. Journal of Morphology 199:363378.Google Scholar
Burke, A. C. 1991. The development and evolution of the turtle body plan: inferring intrinsic aspects of the evolutionary process from experimental embryology. American Zoologist 31:616627.Google Scholar
Burke, A. C., Nelson, C. E., Morgan, B. A., and Tabin, C. 1995. Hox genes and the evolution of vertebrate axial morphology. Development 121:333346.Google Scholar
Buzas, M. A., and Culver, S. J. 1986. Geographic origin of benthic foraminiferal species. Science 232:775776.Google Scholar
Byrne, K., and Nichols, R. A. 1999. Culex pipiens in London Underground tunnels: differentiation between surface and subterranean populations. Heredity 82:715.Google Scholar
Cardillo, M. 1999. Latitude and rates of diversification in birds and butterflies. Proceedings of the Royal Society of London B 266:12211225.Google Scholar
Carroll, R. L. 1997. Patterns and processes of vertebrate evolution. Cambridge University Press, Cambridge.Google Scholar
Carter, D. R., Miki, B., and Padian, K. 1998. Epigenetic mechanical factors in the evolution of long bone epiphyses. Zoological Journal of the Linnean Society 123:163178.Google Scholar
Case, T. J. 1996. Global patterns in the establishment and distribution of exotic birds. Biological Conservation 78:6996.Google Scholar
Chaline, J., and Laurin, B. 1986. Phyletic gradualism in a European Plio-Pleistocene Mimomys lineage (Arvicolidae, Rodentia). Paleobiology 12:203216.Google Scholar
Chaline, J., Brunet-Lecomte, P., Montuire, S., Viriot, L., and Courant, F. 1993. Morphological trends and rates of evolution in arvicolids (Arvicolidae, Rodentia): towards a punctuated equilibria/disequilibria model. Quaternary International 19:2739.Google Scholar
Charlesworth, B., Lande, R., and Slatkin, M. 1982. A neo-Darwinian commentary on macroevolution. Evolution 36:474498.Google Scholar
Cheetham, A. H., and Jackson, J. B. C. 1995. Process from pattern: tests for selection versus random change in punctuated bryozoan speciation. Pp. 184207 in Erwin and Anstey 1995b.Google Scholar
Cheetham, A. H., and Jackson, J. B. C. 1996. Speciation, extinction, and the decline of arborescent growth in Neogene and Quaternary cheilostome Bryozoa of tropical America. Pp. 205233 in Jackson, J. B. C., Budd, A. F., and Coates, A. G., eds. Evolution and environment in tropical America. University of Chicago Press, Chicago.Google Scholar
Chesser, R. T., and Zink, R. M. 1994. Modes of speciation in birds: a test of Lynch's method. Evolution 48:490497.Google Scholar
Cheverud, J. M. 1996. Developmental integration and the evolution of pleiotropy. American Zoologist 36:4450.Google Scholar
Chow, R. L., Altmann, C. R., Lang, R. A., and Hemmati-Brivanlou, A. 1999. Pax6 induces ectopic eyes in a vertebrate. Development 126:42134222.Google Scholar
Claridge, M. F., Dawah, H. A., and Wilson, M. R., eds. Species: the units of biodiversity. Chapman and Hall, London.Google Scholar
Clark, J. S. 1998. Why trees migrate so fast: Confronting theory with dispersal biology and the paleorecord. American Naturalist 152:204224.Google Scholar
Clarke, G. M. 1993. The genetic basis of developmental stability. 1. Relationships between stability, heterozygosity and genomic coadaptation. Genetica 9:1523.Google Scholar
Clarkson, E. N. K. 1988. The origin of marine invertebrate species: a critical review of microevolutionary transformations. Proceedings of the Geologists’ Association 99:153171.Google Scholar
Coates, M. I., and Cohn, M. J. 1998. Fins, limbs, and tails: outgrowths and axial patterning in vertebrate evolution. BioEssays 20:371381.Google Scholar
Cohen, B. L., Balfe, P., Cohen, M., and Curry, G. B. 1991. Molecular evolution and morphological speciation in North Atlantic brachiopods (Terebratulina spp.). Canadian Journal of Zoology 69:29032911.Google Scholar
Cohn, M. J., and Tickle, C. 1999. Developmental basis of limblessness and axial patterning in snakes. Nature 399:474479.Google Scholar
Collins, T. M., Frazer, K., Palmer, A. R., Vermeij, G. J., and Brown, W. M. 1996. Evolutionary history of northern hemisphere Nucella (Gastropoda, Muricidae): molecular, morphological, ecological, and paleontological evidence. Evolution 50:22872304.Google Scholar
Conway Morris, S. 1998. The evolution of diversity in ancient ecosystems: a review. Philosophical Transactions of the Royal Society of London B 353:327345.Google Scholar
Coope, G. R. 1995. Insect faunas in Ice Age environments: why so little extinction? Pp. 5574 in Lawton, J. H. and May, R. M., eds. Extinction rates. Oxford University Press, Oxford.Google Scholar
Coyne, J. A., and Orr, H. A. 1998. The evolutionary genetics of speciation. Philosophical Transactions of the Royal Society of London B 353:287305.Google Scholar
Cracraft, J. 1989. Speciation and its ontology: The empirical consequences of alternative species concepts for understanding patterns and processes of differentiation. Pp. 2859 in Otte and Endler 1989.Google Scholar
Cracraft, J. 1997. Species concepts in systematics and conservation biology—an ornithological viewpoint. Pp. 325339 in Claridge et al. 1997.Google Scholar
Crame, J. A. 1996. Antarctica and the evolution of taxonomic diversity gradients in the marine realm. Terra Antarctica 3:121134.Google Scholar
Crame, J. A. 1997. An evolutionary framework for the polar regions. Journal of Biogeography 24:19.Google Scholar
Crame, J. A., and Clarke, A. 1997. The historical component of marine taxonomic diversity gradients. Pp. 258273 in Ormond, R. F. G., Gage, J. D., and Angel, M. V., eds. Marine biodiversity: patterns and processes. Cambridge University Press, Cambridge.Google Scholar
Darling, K. F., Wade, C. M., Kroon, D., Brown, A. J. L., and Bijma, J. 1999. The diversity and distribution of modern planktic foraminiferal small subunit ribosomal RNA genotypes and their potential as tracers of present and past ocean circulations. Paleoceanography 14:312.Google Scholar
Darling, K. F., Wade, C. M., Stewart, I. A., Kroon, D., Dingle, R., and Brown, A. J. L. 2000. Molecular evidence for genetic mixing of Arctic and Antarctic subpolar populations of planktonic foraminifers. Nature 405:4347.Google Scholar
Davis, J. I., and Nixon, K. C. 1992. Populations, genetic variation, and the delimitation of phylogenetic species. Systematic Biology 41:421435.Google Scholar
Delph, L. F. 1990. Sex-ratio variation in the gynodioecious shrub Hebe strictissima (Scrophulariaceae). Evolution 44:134142.Google Scholar
de Queiroz, K. 1998. The general lineage concept of species, species criteria, and the process of speciation. Pp. 5775 in Howard and Berlocher 1998.Google Scholar
de Queiroz, K. 1999. The general lineage concept of species and the defining properties of the species category. Pp. 4989 in Wilson, R. A., ed. Species: new interdisciplinary essays. MIT Press, Cambridge.Google Scholar
DeSalle, R., and Carew, E. 1992. Phyletic phenocopy and the role of developmental genes in morphological evolution in the Drosophilidae. Journal of Evolutionary Biology 5:363374.Google Scholar
de Vargas, C., Norris, R., Zaninetti, L., Gibb, S. W., and Pawlowski, J. 1999. Molecular evidence of cryptic speciation in planktonic foraminifers and their relation to oceanic provinces. Proceedings of the National Academy of Sciences USA 96:28642868.Google Scholar
DiMichele, W. A., and Aronson, R. B. 1992. The Pennsylvanian-Permian vegetational transition: a terrestrial analogue of the onshore–offshore hypothesis. Evolution 46:807824.Google Scholar
Dodd, M. E., Silvertown, J., and Chase, M. W. 1999. Phylogenetic analysis of trait evolution and species diversity variation among angiosperm families. Evolution 53:732744.Google Scholar
Doebley, J., and Lukens, L. 1998. Transcriptional regulators and the evolution of plant form. Plant Cell 10:10751082.Google Scholar
Doebley, J., and Wang, R.-L. 1997. Genetics and the evolution of plant form: an example from maize. Cold Spring Harbor Symposia in Quantitative Biology 62:361367.Google Scholar
Donoghue, M. J. 1989. Phylogenies and the analysis of evolutionary sequences, with examples from the seed plants. Evolution 43:11371156.Google Scholar
Droser, M. L., Hampt, G., and Clements, S. J. 1993. Environmental patterns in the origin and diversification of rugose and deep-water scleractinian corals. Courier Forschungsinstitut Senckenberg 164:4754.Google Scholar
Droser, M. L., Bottjer, D. J., and Sheehan, P. M. 1997. Evaluating the ecological architecture of major events in the Phanerozoic history of marine invertebrate life. Geology 25:167170.Google Scholar
Duboule, D., and Wilkins, A. S. 1998. The evolution of ‘bricolage.’ Trends in Genetics 14:5459.Google Scholar
Duda, T. F., and Palumbi, S. R. 1999. Developmental shifts and species selection in gastropods. Proceedings of the National Academy of Sciences USA 96:1027210277.Google Scholar
Duffy, J. E. 1996. Species boundaries, specialization, and the radiation of sponge-dwelling alpheid shrimp. Biological Journal of the Linnean Society 58:307324.Google Scholar
Eble, G. J. 2000. Contrasting evolutionary flexibility in sister groups: disparity and diversity in Mesozoic atelostomate echinoids. Paleobiology 26:5679.Google Scholar
Eckenwalder, J. E. 1984. Natural intersectional hybridization between North American species of Populus (Salicaceae) in sections Aigeiros and Tacamahaca. III. Paleobotany and evolution. Canadian Journal of Botany 62:336342.Google Scholar
Eldredge, N. 1985. Unfinished synthesis. Oxford University Press, New York.Google Scholar
Eldredge, N. 1989. Macroevolutionary dynamics. McGraw-Hill, New York.Google Scholar
Eldredge, N. 1995. Species, speciation, and the context of adaptive change in evolution. Pp. 3963 in Erwin and Anstey 1995b.Google Scholar
Emlet, R. B., and Hoegh-Guldberg, O. 1997. Effects of egg size on postlarval performance: experimental evidence from a sea urchin. Evolution 51:141152.Google Scholar
Englemann, G. F., and Wiley, E. O. 1977. The place of ancestor-descendant relationships in phylogeny reconstruction. Systematic Zoology 26:111.Google Scholar
Erwin, D. H. 1994. Early evolution of major morphological innovations. Acta Palaeontologica Polonica 38:281294.Google Scholar
Erwin, D. H. 1998. The end and the beginning: recoveries from mass extinctions. Trends in Ecology and Evolution 13:344349.Google Scholar
Erwin, D. H. 1999. The origin of bodyplans. American Zoologist 39:617629.Google Scholar
Erwin, D. H. 2000a. Macroevolution is more than repeated rounds of microevolution. Evolution and Development 2:7884.Google Scholar
Erwin, D. H. 2000b. Lessons from the past: biotic recoveries from mass extinctions. Proceedings of the National Academy of Sciences USA (in press).Google Scholar
Erwin, D. H., and Anstey, R. L. 1995a. Speciation in the fossil record. Pp. 1138 in Erwin and Anstey 1995b.Google Scholar
Erwin, D. H., and Anstey, R. L. 1995b. New approaches to speciation in the fossil record. Columbia University Press, New York.Google Scholar
Erwin, D. H., Valentine, J. W., and Sepkoski, J. J. Jr. 1987. A comparative study of diversification events: the early Paleozoic versus the Mesozoic. Evolution 41:11771186.Google Scholar
Farrell, B. D. 1998. “Inordinate fondnessexplained: why are there so many beetles? Science 281:555559.Google Scholar
Feldmann, R. M., Tshudy, D. M., and Thomson, M. R. A. 1993. Late Cretaceous and Paleocene decapod crustaceans from James Ross Basin, Antarctic Peninsula. Paleontological Society Memoir 28. Journal of Paleontology 67(Suppl. to No. 1).Google Scholar
Feyereisen, R. 1995. Molecular biology of insecticide resistance. Toxicology Letters 82–3:8390.Google Scholar
Finnerty, J. R., and Martindale, M. Q. 1998. The evolution of the Hox cluster: insights from outgroups. Current Opinion in Genetics and Development 8:681687.Google Scholar
Fisher, D. C. 1994. Stratocladistics: Morphological and temporal patterns and their relation to phylogenetic process. Pp. 133171 in Grande, L. and Rieppel, O., eds. Interpreting the hierarchy of nature. Academic Press, Orlando, Fla.Google Scholar
Flessa, K. W., and Jablonski, D. 1996. The geography of evolutionary turnover: a global analysis of extant bivalves. Pp. 376397 in Jablonski, D., Erwin, D. H., and Lipps, J. H., eds. Evolutionary Paleobiology. University of Chicago Press, Chicago.Google Scholar
Flynn, L. J., Barry, J. C., Morgan, M. E., Pilbeam, D., Jacobs, L. L., and Lindsay, E. H. 1995. Neogene Siwalik mammalian lineages: species longevities, rates of change, and modes of speciation. Palaeogeography, Palaeoclimatology, Palaeoecology 115:249264.Google Scholar
Foote, M. 1993. Discordance and concordance between morphological and taxonomic diversity. Paleobiology 19:185204.Google Scholar
Foote, M. 1996a. Perspective: evolutionary patterns in the fossil record. Evolution 50:111.Google Scholar
Foote, M. 1996b. On the probability of ancestors in the fossil record. Paleobiology 22:141151.Google Scholar
Foote, M. 1997. The evolution of morphological diversity. Annual Review of Ecology and Systematics 28:129152.Google Scholar
Foote, M. 1999. Morphological diversity in the evolutionary radiation of Paleozoic and post-Paleozoic crinoids. Paleobiology Memoirs No. 1. Paleobiology 25(Suppl. to No. 2).Google Scholar
Fortey, R. A. 1985. Gradualism and punctuated equilibria as competing and complementary theories. Special Papers in Palaeontology 33:1728.Google Scholar
Fortey, R. A. 1988. Seeing is believing: gradualism and punctuated equilibria in the fossil record. Science Progress 72:119.Google Scholar
Frost, D. R., and Kluge, A. G. 1994. A consideration of epistemology in systematic biology, with special reference to species. Cladistics 10:259294.Google Scholar
Futuyma, D. J. 1987. On the role of species in anagenesis. American Naturalist 130:465475.Google Scholar
Futuyma, D. J., Keese, M. C., and Funk, D. J. 1995. Genetic constraints on macroevolution: the evolution of host affiliation in the leaf beetle genus Ophraella. Evolution 49:797809.Google Scholar
Gaffney, B., and Cunningham, E. P. 1988. Estimation of genetic trend in racing performance of thoroughbred horses. Nature 332:722724.Google Scholar
Garland, T., Midford, P. E., and Ives, A. R. 1999. An introduction to phylogenetically based statistical methods, with a new method for confidence intervals on ancestral values. American Zoologist 39:374388.Google Scholar
Gaston, K. J. 1998. Species-range size distributions: products of speciation, extinction and transformation. Philosophical Transactions of the Royal Society of London B 353:219230.Google Scholar
Gaston, K. J., and Blackburn, T. M. 1996. The tropics as a museum of biological diversity: an analysis of the New World avifauna. Proceedings of the Royal Society of London B 263:6368.Google Scholar
Gaston, K. J., and Blackburn, T. M. 1997. Age, area and avian diversification. Biological Journal of the Linnean Society 62:239253.Google Scholar
Gaston, K. J., and Blackburn, T. M. 1999. A critique for macroecology. Oikos 84:353368.Google Scholar
Geary, D. H. 1995. The importance of gradual change in species-level transitions. Pp. 6786 in Erwin and Anstey 1995b.Google Scholar
Gehring, W. J., and Ikeo, K. 1999. Pax-6: mastering eye morphogenesis and eye evolution. Trends in Genetics 15:371377.Google Scholar
Geller, J. B. 1998. Molecular studies of marine invertebrate biodiversity: Status and prospects. Pp. 359376 in Cooksey, K. E., ed. Molecular approaches to the study of the ocean. Chapman and Hall, London.Google Scholar
Gellon, G., and McGinnis, W. 1998. Shaping animal body plans in development and evolution by modulation of Hox expression patterns. BioEssays 20:116125.Google Scholar
Ghiselin, M. T. 1997. Metaphysics and the origin of species. State University of New York Press, Albany.Google Scholar
Ghosh, S., May, M. J., and Kopp, E. B. 1998. NF-kappa B and rel proteins: evolutionarily conserved mediators of immune responses. Annual Review of Immunology 16:225260.Google Scholar
Gibson, G. 1999. Insect evolution: Redesigning the fruitfly. Current Biology 9:R86R89.Google Scholar
Gibson, G., and Wagner, G. 2000. Canalization in evolutionary genetics: a stabilizing theory? BioEssays 22:372380.Google Scholar
Gibson, G., Wemple, M., and Helden, S. van. 1999. Potential variance affecting homeotic Ultrabithorax and Antennipedia phenotypes in Drosophila melanogaster. Genetics 151:10811091.Google Scholar
Gili, C., and Martinell, J. 1994. Relationship between species longevity and larval ecology in nassariid gastropods. Lethaia 27:291299.Google Scholar
Gingerich, P. D. 1993. Quantification and comparison of evolutionary rates. American Journal of Science 293-A:453478.Google Scholar
Goffinet, B., and Gerber, S. 2000. Quantitative Trait Loci: a metaanalysis. Genetics 155:463473.Google Scholar
Gonzalez-Crespo, S., and Levine, M. 1994. Related target enhancers for dorsal and NF-kappa-B signalling pathways. Science 264:255258.Google Scholar
Gottlieb, L. D. 1984. Genetics and morphological evolution in plants. American Naturalist 123:681709.Google Scholar
Gould, S. J. 1977a. Eternal metaphors of paleontology. Pp. 126 in Hallam, A., ed. Patterns of evolution. Elsevier, Amsterdam.Google Scholar
Gould, S. J. 1977b. Ontogeny and phylogeny. Harvard University Press, Cambridge.Google Scholar
Gould, S. J. 1982. The meaning of punctuated equilibrium and its role in validating a hierarchical approach to macroevolution. Pp. 83104 in Milkman, R., ed. Perspectives on evolution. Sinauer, Sunderland, Mass.Google Scholar
Gould, S. J. 1985. The paradox of the first tier: an agenda for paleobiology. Paleobiology 11:212.Google Scholar
Gould, S. J., and Eldredge, N. 1977. Punctuated equilibria: the tempo and mode of evolution reconsidered. Paleobiology 3:115151.Google Scholar
Gould, S. J., and Eldredge, N. 1993. Punctuated equilibrium comes of age. Nature 366:223227.Google Scholar
Gould, S. J., and Lloyd, E. A. 1999. Individuality and adaptation across levels of selection: how shall we name and generalize the unit of Darwinism? Proceedings of the National Academy of Sciences USA 96:1190411909.Google Scholar
Govindaraju, D. R. 1988. Relationship between dispersal ability and levels of gene flow in plants. Oikos 52:3135.Google Scholar
Graff, A. 1999. Population sex structure and reproductive fitness in gynodioecious Sidalcea malviflora malviflora (Malvaceae). Evolution 53:17141722.Google Scholar
Grant, P. R. 1986. Ecology and evolution of Darwin's finches. Princeton University Press, Princeton, N.J. Google Scholar
Grant, P. R., and Grant, B. R. 1995. Predicting microevolutionary responses to directional selection on heritable variation. Evolution 49:241251.Google Scholar
Grantham, T. A. 1995. Hierarchical approaches to macroevolution: recent work on species selection and the “effect hypothesis.” Annual Review of Ecology and Systematics 26:301322.Google Scholar
Greene, H. W., and Cundall, D. 2000. Limbless tetrapods and snakes with legs. Science 287:19391941.Google Scholar
Hagdorn, H., and Campbell, H. J. 1993. Paracomatula triadica sp. nov.—an early comatulid crinoid from the Otapirian (Late Triassic) of New Caledonia. Alcheringa 17:117.Google Scholar
Hageman, S. J., Bayer, M. M., and Todd, C. D. 1999. Partitioning phenotypic variation: genotypic, environmental and residual components from bryozoan skeletal morphology. Journal of Natural History 33:17131735.Google Scholar
Halder, G., Callaerts, P., and Gehring, W. J. 1995. Induction of ectopic eyes by targeted expression of the eyeless gene in Drosophila. Science 267:17881792.Google Scholar
Hall, B. K. 1983. Epigenetic control in development and evolution. Pp. 353379 in Goodwin, B. C., Holder, N., and Wylie, C. C., eds. Development and evolution. Cambridge University Press, Cambridge.Google Scholar
Hall, B. K. 1999. Evolutionary developmental biology, 2d ed. Kluwer Academic, Dordrecht, Netherlands.Google Scholar
Hallam, A. 1998. Speciation patterns and trends in the fossil record. Geobios 7:921930.Google Scholar
Hansen, T. A. 1978. Larval dispersal and species longevity in Lower Tertiary gastropods. Science 199:885887.Google Scholar
Hansen, T. A. 1982. Modes of larval development in early Tertiary neogastropods. Paleobiology 8:367372.Google Scholar
Hansen, T. A., Kelley, P. H., Melland, V. D., and Graham, S. E. 1999. Effect of climate-related mass extinctions on escalation in molluscs. Geology 27:11391142.Google Scholar
Hanski, I. 1999. Metapopulation ecology. Oxford University Press, Oxford.Google Scholar
Harrington, R. J. 1987. Skeletal growth histories of Protothaca staminea (Conrad) and Protothaca grata (Say) throughout their geographic ranges, northeastern Pacific. Veliger 30:148158.Google Scholar
Harrison, R. G. 1998. Linking evolutionary pattern and process: the relevance of species concepts for the study of speciation. Pp. 1931 in Howard and Berlocher 1998.Google Scholar
Harrison, S. 1998. Do taxa persist as metapopulations in evolutionary time? Pp. 1930 in McKinney and Drake 1998.Google Scholar
Harvey, P. H., and Rambaut, A. 1998. Phylogenetic extinction rates and comparative methodology. Proceedings of the Royal Society of London B 265:16911696.Google Scholar
Heanue, T. A., Rashef, R., Davis, R. J., Mardon, G., Oliver, G., Tomarev, S., Lassar, A. B., and Tabin, C. J. 1999. Synergistic regulation of vertebrate development by Dach2, Eya2, and Six1, homologs of genes required for Drosophila eye formation. Genes and Development 13:32313243.Google Scholar
Heaton, T. H. 1993. The Oligocene rodent Ischyromys of the Great Plains: replacement mistaken for anagenesis. Journal of Paleontology 67:297308.Google Scholar
Hendry, A. P., and Kinnison, M. T. 1999. Perspective: the pace of modern life: measuring rates of contemporary microevolution. Evolution 53:16371653.Google Scholar
Hodin, J. 2000. Plasticity and constraints in development and evolution. Journal of Experimental Zoology (Molecular and Developmental Evolution) 288:120.Google Scholar
Hoffman, A. 1989. Arguments on evolution. Oxford University Press, Oxford.Google Scholar
Hoffmann, A. A., and Hercus, M. J. 2000. Environmental stress as an evolutionary force. BioScience 50:217226.Google Scholar
Hoffmann, A. A., and Parsons, P. A. 1997. Extreme environmental change and evolution. Cambridge University Press, Cambridge.Google Scholar
Holland, P. W. H. 1999. The future of evolutionary developmental biology. Nature 402(Suppl.):C41C44.Google Scholar
Holland, S. M., and Patzkowsky, M. E. 1999. Models for simulating the fossil record. Geology 27:491494.Google Scholar
Howard, D. J., and Berlocher, S. H., eds. Endless forms: species and speciation. Oxford University Press, New York.Google Scholar
Huber, B. T., Bijma, J., and Darling, K. 1997. Cryptic speciation in the living planktonic foraminifer Globigerinella siphonifera (d'Orbigny). Paleobiology 23:3362.Google Scholar
Hughes, N. C., Chapman, R. E., and Adrain, J. M. 1999. The stability of thoracic segmentation in trilobites: a case study in developmental and ecological constraints. Evolution and Development 1:2435.Google Scholar
Hull, D. L. 1980. Individuality and selection. Annual Review of Ecology and Systematics 11:311332.Google Scholar
Hull, D. L. 1997. The ideal species concept—and why we can't get it. Pp. 357380 in Claridge et al. 1997.Google Scholar
Humphreville, R., and Bambach, R. K. 1979. Influence of geography, climate and ocean circulation on the pattern of generic diversity of brachiopods in the Permian. Geological Society of America Abstracts with Programs 11:447.Google Scholar
Jablonski, D. 1986a. Larval ecology and macroevolution of marine invertebrates. Bulletin of Marine Science 39:565587.Google Scholar
Jablonski, D. 1986b. Background and mass extinctions: the alternation of macroevolutionary regimes. Science 231:129133.Google Scholar
Jablonski, D. 1987. Heritability at the species level: analysis of geographic ranges of Cretaceous mollusks. Science 238:360363.Google Scholar
Jablonski, D. 1988. Response [to Russell and Lindberg]. Science 240:969.Google Scholar
Jablonski, D. 1993. The tropics as a source of evolutionary novelty: the post-Palaeozoic fossil record of marine invertebrates. Nature 364:142144.Google Scholar
Jablonski, D. 1995. Extinction in the fossil record. Pp. 2544 in Lawton, J. H. and May, R. M., eds. Extinction rates. Oxford University Press, Oxford.Google Scholar
Jablonski, D. 1998. Geographic variation in the molluscan recovery from the end-Cretaceous extinction. Science 279:13271330.Google Scholar
Jablonski, D. 2000a. The interplay between physical and biotic factors in evolution. In Lister, A. and Rothschild, L., eds. Evolution on planet Earth: the impact of the physical environment. Linnean Society and Academic Press, London (in press).Google Scholar
Jablonski, D. 2000b. Lessons from the past: Evolutionary impacts of mass extinctions. Proceedings of the National Academy of Sciences USA (in press).Google Scholar
Jablonski, D., and Bottjer, D. J. 1983. Soft-substratum epifaunal suspension-feeding assemblages in the Late Cretaceous: Implications for the evolution of benthic paleocommunities. Pp. 747812 in Tevesz, M. J. and McCall, P. L., eds. Biotic interactions in Recent and fossil benthic communities. Plenum, New York.Google Scholar
Jablonski, D., and Bottjer, D. J. 1990a. The ecology of evolutionary innovations: the fossil record. Pp. 253288 in Nitecki, M. H., ed. Evolutionary innovations. University of Chicago Press, Chicago.Google Scholar
Jablonski, D., and Bottjer, D. J. 1990b. The origin and diversification of major groups: environmental patterns and macroevolutionary lags. Pp. 1757 in Taylor, P. D. and Larwood, G. P., eds. Major evolutionary radiations. Clarendon, Oxford.Google Scholar
Jablonski, D., and Bottjer, D. J. 1991. Environmental patterns in the origins of higher taxa: the post-Paleozoic fossil record. Science 252:18311833.Google Scholar
Jablonski, D., and Lutz, R. A. 1983. Larval ecology of marine benthic invertebrates: paleobiological implications. Biological Reviews 58:2189.Google Scholar
Jablonski, D., and Sepkoski, J. J. Jr. 1996. Paleobiology, community ecology, and scales of ecological pattern. Ecology 77:13671378.Google Scholar
Jablonski, D., Lidgard, S., and Taylor, P. D. 1997. Comparative ecology of bryozoan radiations: origin of novelties in cyclostomes and cheilostomes. Palaios 12:505523.Google Scholar
Jackson, J. B. C., and Budd, A. F. 1996. Evolution and environment: Introduction and overview. Pp. 120 in Jackson, J. B. C., Budd, A. F., and Coates, A. G., eds. Evolution and environment in tropical America. University of Chicago Press, Chicago.Google Scholar
Jackson, J. B. C., and Cheetham, A. H. 1990. Evolutionary significance of morphospecies: a test with cheilostome Bryozoa. Science 248:579583.Google Scholar
Jackson, J. B. C., and Cheetham, A. H. 1994. Phylogeny reconstruction and the tempo of speciation in cheilostome Bryozoa. Paleobiology 20:407423.Google Scholar
Jackson, J. B. C., and Cheetham, A. H. 1999. Tempo and mode of speciation in the sea. Trends in Ecology and Evolution 14:7277.Google Scholar
Jackson, J. B. C., and Hughes, T. P. 1985. Adaptive strategies of coral-reef invertebrates. American Scientist 73:265274.Google Scholar
Jackson, S. T., and Overpeck, J. T. 2000. Responses of plant populations and communities to environmental changes of the Late Quaternary. Paleobiology 26:194220.Google Scholar
Jacobs, D. K. 1990. Selector genes and the Cambrian radiation of Bilateria. Proceedings of the National Academy of Sciences USA 87:44064410.Google Scholar
Johnson, A. L. A. 1984. The paleobiology of the bivalve families Pectinidae and Propeamussidae in the Jurassic of Europe. Zitteliana 11:235 p.Google Scholar
Johnson, A. L. A. 1985. The rate of evolutionary change in European Jurassic scallops. Special Papers in Palaeontology 33:91102.Google Scholar
Johnson, A. L. A. 1993. Punctuated equilibria vs. phyletic gradualism in European Jurassic Gryphaea evolution. Proceedings of the Geologists’ Association 104:209222.Google Scholar
Johnson, A. L. A. 1994. Evolution of European Lower Jurassic Gryphaea (Gryphaea) and contemporaneous bivalves. Historical Biology 7:167186.Google Scholar
Johnson, A. L. A., and Lennon, C. D. 1990. Evolution of gryphaeate oysters in the mid-Jurassic of western Europe. Palaeontology 33:453485.Google Scholar
Johnson, J. G. 1982. Occurrence of phyletic gradualism and punctuated equilibria through time. Journal of Paleontology 56:13291331.Google Scholar
Jones, D. S. 1988. Sclerochronology and the size versus age problem. Pp. 93108 in McKinney, M. L., ed. Heterochrony in evolution. Plenum, New York.Google Scholar
Jones, D. S. 1998. Isotopic determination of growth and longevity in fossil and modern invertebrates. In Norris, R. D. and Corfield, R. M., eds. Isotope paleobiology and paleoecology. Paleontological Society Papers 4:3767. Paleontological Society, Knoxville, Tenn.Google Scholar
Jones, D. S., and Gould, S. J. 1999. Direct measurement of age in fossil Gryphaea: the solution to a classic problem in heterochrony. Paleobiology 25:158187.Google Scholar
Judd, W. S., Sanders, R. W., and Donoghue, M. J. 1994. Angiosperm family pairs: preliminary phylogenetic analysis. Harvard Papers in Botany 5:151.Google Scholar
Kammer, T. W., Baumiller, T. K., and Ausich, W. I. 1998. Evolutionary significance of differential species longevity in Osagean-Meramecian (Mississippian) crinoid clades. Paleobiology 24:155176.Google Scholar
Keese, M. C. 1998. Performance of two monophagous leaf feeding beetles (Coleoptera: Chrysomelidae) on each other's host plant: do intrinsic factors determine host plant specialization? Journal of Evolutionary Biology 11:403419.Google Scholar
Kelley, P. H. 1989. Evolutionary trends within bivalve prey of Chesapeake Group naticid gastropods. Historical Biology 2:139156.Google Scholar
Kelley, P. H. 1991. The effect of predation intensity on rate of evolution of five Miocene bivalves. Historical Biology 5:6578.Google Scholar
Kelley, P. H., Raymond, A., and Lutken, C. B. 1990. Carboniferous brachiopod migration and latitudinal diversity; a new palaeoclimatic method. Pp. 325332 in McKerrow, W. S. and Scotese, C. R., eds. Palaeozoic palaeogeography and biogeography. Geological Society of London Memoir 12.Google Scholar
Kennish, M. J. 1980. Shell microgrowth analysis: Mercenaria mercenaria as a type example for research in population dynamics. Pp. 255294 in Rhoads, D. C. and Lutz, R. A., eds. Skeletal growth of aquatic organisms. Plenum, New York.Google Scholar
Keys, D. N., Lewis, D. L., Selegue, J. E., Pearson, B. J., Goodrich, L. V., Johnson, R. J., Gates, J., Scott, M. P., and Carroll, S. B. 1999. Recruitment of a hedgehog regulatory circuit in butterfly eye-spot evolution. Science 283:532534.Google Scholar
Kidwell, S. M., and Aigner, T. A. 1985. Sedimentary dynamics of complex shell beds: implications for ecologic and evolutionary patterns. Pp. 383395 in Bayer, U. and Seilacher, A., eds. Sedimentary and evolutionary cycles. Springer, Berlin.Google Scholar
Kidwell, S. M., and Flessa, K. W. 1995. The quality of the fossil record: populations, species, and communities. Annual Review of Ecology and Systematics 26:269299.Google Scholar
Kirkpatrick, M. 1996. Genes and adaptation: a pocket guide to the theory. Pp. 125146 in Rose, M. R. and Lauder, G. V., eds. Adaptation. Academic Press, San Diego.Google Scholar
Kirkpatrick, M., and Slatkin, M. 1993. Searching for evolutionary patterns in the shape of a phylogenetic tree. Evolution 47:11711181.Google Scholar
Kirschner, M., and Gerhart, J. 1997. Cells, embryos, and evolution. Blackwell Science, Malden, Mass.Google Scholar
Kitazato, H., Tsuchiya, M., and Takahara, K. 2000. Recognition of breeding populations in foraminifera: an example using the genus Glabratella. Paleontological Research 4:115.Google Scholar
Klingenberg, C. P. 1998. Heterochrony and allometry: the analysis of evolutionary change in ontogeny. Biological Reviews 73:79123.Google Scholar
Knoll, A. H., and Carroll, S. B. 1999. Early animal evolution: emerging views from comparative biology and geology. Science 284:21292137.Google Scholar
Knowlton, N. 1993. Sibling species in the sea. Annual Review of Ecology and Systematics 24:189216.Google Scholar
Knowlton, N., and Weigt, L. A. 1997. Species of marine invertebrates: a comparison of the biological and phylogenetic species concepts. Pp. 199219 in Claridge et al. 1997.Google Scholar
Knowlton, N., Maté, J., Guzmán, H. M., Rowan, R., and Jara, J. 1997. Direct evidence for reproductive isolation among the three species of the Montastraea annularis complex in Central America (Panamá and Honduras). Marine Biology 127:705711.Google Scholar
Koch, P. L. 1986. Clinal geographic variation in mammals: implications for the study of chronoclines. Paleobiology 12:269281.Google Scholar
Kodric-Brown, A., and Brown, J. H. 1979. Competition between distantly related taxa and the co-evolution of plants and pollinators. American Zoologist 19:11151127.Google Scholar
Kohn, A. J., and Perron, F. E. 1994. Life history and biogeography: patterns in Conus . Clarendon, Oxford.Google Scholar
Kowalewski, M., Dyreson, E., Marcot, J. D., Vargas, J. A., Flessa, K. W., and Hallman, D. P. 1997. Phenetic discrimination of biometric simpletons: paleobiological implications of morphospecies in the lingulide brachiopod Glottidia. Paleobiology 23:444469.Google Scholar
Lande, R. 1976. Natural selection and random genetic drift in phenotypic evolution. Evolution 30:314334.Google Scholar
Lande, R. 1986. The dynamics of peak shifts and the pattern of morphological evolution. Paleobiology 12:343354.Google Scholar
Langan-Cranford, K. M., and Pearse, J. S. 1995. Breeding experiments confirm species status of two morphologically similar gastropods (Lacuna spp.) in central California. Journal of Experimental Marine Biology and Ecology 186:1731.Google Scholar
Lawton, J. H. 1999. Are there general laws in ecology? Oikos 84:177192.Google Scholar
Lawton-Rauh, A. L., Alvzarez-Buylla, E. R., and Purugganan, M. D. 2000. Molecular evolution of flower development. Trends in Ecology and Evolution 15:144149.Google Scholar
Lazarus, D., Hilbrecht, H., Spencer-Cervas, C., and Thierstein, H. 1995. Sympatric speciation and phyletic change in Globorotalia truncatulinoides. Paleobiology 21:2851.Google Scholar
Leamy, L. J., Routman, E. J., and Cheverud, J. M. 1999. Quantitative trait loci for early- and late-developing skull characters in mice: a test of the genetic independence model of morphological integration. American Naturalist 153:201214.Google Scholar
Lerner, I. M. 1953. Genetic homeostasis. Wiley, New York.Google Scholar
Levin, D. A. 1970. Developmental instability and evolution in peripheral populations. American Naturalist 104:343353.Google Scholar
Levin, S. A. 1999. Fragile dominion: complexity and the commons. Perseus Books, Reading, Mass.Google Scholar
Levinton, J. 1988. Genetics, paleontology, and macroevolution. Cambridge University Press, New York.Google Scholar
Lewontin, R. C. 1970. The units of selection. Annual Review of Ecology and Systematics 1:118.Google Scholar
Li, X., and McGinnis, W. 1999. Activity regulation of Hox proteins, a mechanism for altering functional specificity in development and evolution. Proceedings of the National Academy of Sciences USA 96:68026807.Google Scholar
Lidgard, S., McKinney, F. K., and Taylor, P. D. 1993. Competition, clade replacement, and a history of cyclostome and cheilostome diversity. Paleobiology 19:352371.Google Scholar
Lieberman, B. S., Brett, C. E., and Eldredge, N. 1994. Patterns and processes of stasis in two species lineages of brachiopods from the Middle Devonian of New York State. American Museum Novitates 3114:123.Google Scholar
Lieberman, B. S., Brett, C. E., and Eldredge, N. 1995. A study of stasis and change in two species lineages of brachiopods from the Middle Devonian of New York State. Paleobiology 21:1527.Google Scholar
Lindberg, D. R. 1988. Heterochrony in gastropods: a neontological view. Pp. 197216 in McKinney, M. L., ed. Heterochrony in evolution. Plenum, New York.Google Scholar
Lister, A. M. 1989. Rapid dwarfing of red deer on Jersey in the last interglacial. Nature 342:539542.Google Scholar
Lloyd, E. A., and Gould, S. J. 1993. Species selection on variability. Proceedings of the National Academy of Sciences USA 90:595599.Google Scholar
Long, A. D., Mullaney, S. L., Mackay, T. F. C., and Langley, C. H. 1996. Genetic interactions between naturally occurring alleles at quantitative trait loci and mutant alleles at candidate loci affecting bristle number in Drosophila melanogaster. Genetics 144:14971510.Google Scholar
Lonsdale, W. M. 1999. Global patterns of plant invasions and the concept of invasibility. Ecology 80:15221536.Google Scholar
Losos, J. B., Warheit, K. I., and Schoener, T. W. 1997. Adaptive differentiation following experimental island colonization in Anolis lizards. Nature 387:7073.Google Scholar
Luckow, M. 1995. Species concepts: assumptions, methods, and applications. Systematic Botany 20:589605.Google Scholar
Ludwig, M. Z., Bergman, C., Patel, N. H., and Kreitman, M. 2000. Evidence for stabilizing selection in a eukaryotic enhancer element. Nature 403:564567.Google Scholar
Lupia, R. 1999. Discordant morphological disparity and taxonomic diversity during the Cretaceous angiosperm radiation: North American pollen record. Paleobiology 25:128.Google Scholar
Lynch, J. B. 1989. The gauge of speciation: on the frequencies of modes of speciation. Pp. 527553 in Otte and Endler 1989.Google Scholar
Lynch, M., and Walsh, J. B. 1998. Genetics and analysis of quantitative traits. Sinauer, Sunderland, Mass.Google Scholar
Mackay, T. F. C. 1996. The nature of quantitative genetic variation revisited: lessons from Drosophila bristles. BioEssays 18:113121.Google Scholar
MacLeod, N. 1991. Punctuated anagenesis and the importance of stratigraphy to paleobiology. Paleobiology 17:167188.Google Scholar
Marshall, C. R. 1991. Estimation of taxonomic ranges from the fossil record. In Gilinsky, N. L. and Signor, P. W., eds. Analytical paleobiology. Short Courses in Paleontology 4:1938. Paleontological Society, Knoxville, Tenn.Google Scholar
Marshall, C. R. 1995. Stratigraphy, the true order of species originations and extinctions, and testing ancestor-descendant-hypotheses among Caribbean Neogene bryozoans. Pp. 208235 in Erwin and Anstey 1995b.Google Scholar
Martin, P. S., and Steadman, D. W. 1999. Prehistoric extinctions on islands and continents. Pp. 1755 in MacPhee, R. D. E., ed. Extinctions in near time. Kluwer Academic/Plenum, New York.Google Scholar
Martin, R. A. 1993. Patterns of variation and speciation in Quaternary rodents. Pp. 226280 in Martin, R. A. and Barnosky, A. D., eds. Morphological change in Quaternary mammals of North America. Cambridge University Press, Cambridge.Google Scholar
Martin, R. E. 1999. Taphonomy: a process approach. Cambridge University Press, New York.Google Scholar
Maurer, B. A., and Nott, M. P. 1998. Geographic range fragmentation and the evolution of biological diversity. Pp. 3150 in McKinney and Drake 1998.Google Scholar
Mayden, R. L. 1997. A hierarchy of species concepts: the denouement in the saga of the species problem. Pp. 381424 in Claridge et al. 1997.Google Scholar
Maynard Smith, J. 1989. The causes of extinction. Philosophical Transactions of the Royal Society of London B 325:241252.Google Scholar
Mayr, E. 1992. Controversies in retrospect. Oxford Surveys in Evolutionary Biology 8:134.Google Scholar
McKenzie, J. A., and Batterham, P. 1994. The genetic, molecular and phenotypic consequences of selection for insecticide resistance. Trends in Ecology and Evolution 9:166169.Google Scholar
McKinney, F. K., Lidgard, S., Sepkoski, J. J. Jr., and Taylor, P. D. 1998. Decoupled temporal patterns of evolution and ecology in two post-Paleozoic clades. Science 281:809809.Google Scholar
McKinney, M. L. 1997. Extinction vulnerability and selectivity: combining ecological and paleontological views. Annual Review of Ecology and Systematics 28:495516.Google Scholar
McKinney, M. L. 1998. Biodiversity dynamics: Niche preemption and saturation in diversity equilibria. Pp. 116 in McKinney and Drake 1998.Google Scholar
McKinney, M. L., and Drake, J. A., eds. 1998. Biodiversity dynamics. Columbia University Press, New York.Google Scholar
McKinney, M. L., and Gittleman, J. L. 1995. Ontogeny and phylogeny: tinkering with covariation in life history, morphology and behaviour. Pp. 2147 in McNamara, K. J., ed. Evolutionary change and heterochrony. Wiley, Chichester, England.Google Scholar
McKinney, M. L., and McNamara, K. J. 1991. Heterochrony: the evolution of ontogeny. Plenum, New York.Google Scholar
McNamara, K. J. 1983. Progenesis in trilobites. Special Papers in Palaeontology 30:5968.Google Scholar
McShea, D. W. 1994. Mechanisms of large-scale evolutionary trends. Evolution 48:17471763.Google Scholar
Meyen, S. V. 1992. Geography of macroevolution in higher plants. Soviet Scientific Reviews G (Geology) 1:3970.Google Scholar
Michaux, B. 1987. An analysis of allozymic characters of four species of New Zealand Amalda (Gastropoda: Olividae: Ancillinae). New Zealand Journal of Zoology 14:359366.Google Scholar
Michaux, B. 1989. Morphological variation of species through time. Biological Journal of the Linnean Society 38:239255.Google Scholar
Miller, A. I. 1997a. Comparative diversification dynamics among palaeocontinents during the Ordovician radiation. Geobios Mémoire Spécial 20:397406.Google Scholar
Miller, A. I. 1997b. Dissecting global diversity patterns: examples from the Ordovician Radiation. Annual Review of Ecology and Systematics 28:85104.Google Scholar
Miller, A. I. 1998. Biotic transitions in global marine diversity. Science 281:11571160.Google Scholar
Miller, A. I., and Sepkoski, J. J. Jr. 1988. Modeling bivalve diversification: the effect of interaction on a macroevolutionary system. Paleobiology 14:364369.Google Scholar
Müller, G. B. 1990. Developmental mechanisms at the origin of morphological novelty: a side-effect hypothesis. Pp. 99130 in Nitecki, M. H., ed. Evolutionary innovations. University of Chicago Press, Chicago.Google Scholar
Nemeschkal, H. L. 1999. Morphometric correlation patterns of adult birds (Fringillidae: Passeriformes and Columbiformes) mirror the expression of developmental control genes. Evolution 53:899918.Google Scholar
Nevo, E., Ben-Shlomo, R., Beiles, A., Hart, C. P., and Ruddle, F. H. 1992. Homeobox DNA polymorphisms (RFLPs) in subterranean mammals of the Spalax ehrenbergi superspecies in Israel: patterns, correlates, and evolutionary significance. Journal of Experimental Zoology 263:430441.Google Scholar
Newell, N. D. 1956. Fossil populations. In Sylvester-Bradley, A. C., ed. The species concept in palaeontology. Systematics Association Publication 2:6382.Google Scholar
Nijhout, H. F., and Emlen, D. J. 1998. Competition among body parts in the development and evolution of insect morphology. Proceedings of the National Academy of Sciences USA 95:36853689.Google Scholar
Nijhout, H. F., and Paulsen, S. M. 1997. Developmental models and polygenic characters. American Naturalist 149:394405.Google Scholar
Niklas, K. J. 1997. The evolutionary biology of plants. University of Chicago Press, Chicago.Google Scholar
Nixon, K. C., and Wheeler, Q. D. 1990. An amplification of the phylogenetic species concept. Cladistics 6:211223.Google Scholar
Norell, M. A. 1996. Ghost taxa, ancestors, and assumptions: a comment on Wagner. Paleobiology 22:453455.Google Scholar
Norris, R. D., Corfield, R. M., and Cartlidge, J. E. 1996. What is gradualism? Cryptic speciation in globorotaliid foraminifera. Paleobiology 22:386405.Google Scholar
Nunn, C. L., and Smith, K. K. 1998. Statistical analyses of developmental sequences: the craniofacial region in marsupial and placental mammals. American Naturalist 152:82101.Google Scholar
Nuzhdin, S. V., Dilda, C. L., and Mackay, T. F. C. 1999. The genetic architecture of selection response: inferences from fine-scale mapping of bristle number quantitative trait loci in Drosophila melanogaster. Genetics 153:13171331.Google Scholar
Olson, E. C., and Miller, R. L. 1958. Morphological integration. University of Chicago Press, Chicago (reprinted 1999).Google Scholar
Omland, K. E. 1997. Examining two standard assumptions of ancestral reconstructions: repeated loss of dichromatism in dabbling ducks (Anatini). Evolution 51:16361646.Google Scholar
Orr, H. A. 1998. The population genetics of adaptation: the distribution of factors fixed during adaptive evolution. Evolution 52:935949.Google Scholar
Orr, H. A., and Coyne, J. A. 1992. The genetics of adaptation revisited. American Naturalist 140:725742.Google Scholar
Otte, D., and Endler, J. A. 1989. Speciation and its consequences. Sinauer, Sunderland, Mass.Google Scholar
Page, K. N. 1996. Mesozoic ammonoids in time and space. Pp. 755794 in Landman, N. H., Tanabe, K., and Davis, R. A., eds. Ammonoid paleobiology. Plenum, New York.Google Scholar
Palmer, A. R. 1985. Quantum changes in gastropod shell morphology need not reflect speciation. Evolution 39:699705.Google Scholar
Palopoli, M. F., and Patel, N. H. 1996. Neo-Darwinian developmental evolution: Can we bridge the gap between pattern and process? Current Opinion in Genetics and Development 6:502508.Google Scholar
Palumbi, S. R. 1996. Macrospatial genetic structure and speciation in marine taxa with high dispersal abilities. Pp. 101117 in Ferraris, J. and Palumbi, S. R., eds. Molecular zoology. Wiley, New York.Google Scholar
Palumbi, S. R., Grabowsky, G., Duda, T., Geyer, L., and Tachino, N. 1997. Speciation and population genetic structure in tropical Pacific sea urchins. Evolution 51:15061517.Google Scholar
Paradis, E. 1998. Detecting shifts in diversification rates without fossils. American Naturalist 152:176187.Google Scholar
Parsons, P. A. 1993. Stress, extinctions and evolutionary change: from living organisms to fossils. Biological Reviews 68:313333.Google Scholar
Parsons, P. A. 1994. Habitats, stress, and evolutionary rates. Journal of Evolutionary Biology 7:387397.Google Scholar
Paterson, A. H., Lin, Y.-R., Li, Z., Schertz, K. F., Doebley, J. F., Pinson, S. R. M., Liu, S.-C., Stansel, J. W., and Irvine, J. E. 1997. Convergent domestication of cereal crops by independent mutation at corresponding genetic loci. Science 269:17141718.Google Scholar
Patzkowsky, M. E. 1995. A hierarchical branching model of evolutionary radiations. Paleobiology 21:440460.Google Scholar
Paul, C. R. C. 1992. The recognition of ancestors. Historical Biology 6:239250.Google Scholar
Pearson, P. N. 1995. Investigating age-dependency of species extinction rates using dynamic survivorship analysis. Historical Biology 10:119136.Google Scholar
Pechenik, J. A. 1999. On the advantages and disadvantages of larval stages in benthic marine invertebrate life cycles. Marine Ecology Progress Series 177:269297.Google Scholar
Peterson, A. T., Soberon, J., and Sanchez-Cordero, V. 1999. Conservatism of ecological niches in evolutionary time. Science 285:12651267.Google Scholar
Potts, D. C., Budd, A. F., and Garthwaite, R. L. 1993. Soft tissue vs. skeletal approaches to species recognition and phylogeny reconstruction in corals. Courier Forschungsinstitut Senckenberg 16:221231.Google Scholar
Povel, G. D. E. 1993. The main branch of Miocene Gyraulus (Gastropoda; Planorbindae) of Steinheim (southern Germany): a reconsideration of Mensink's data set. Scripta Geologica Special Issue 3:371386.Google Scholar
Price, T. D., Helbig, A. J., and Richman, A. D. 1997. Evolution of breeding distributions in the Old World leaf warblers (genus Phylloscopus). Evolution 51:552561.Google Scholar
Prothero, D. R., and Heaton, T. H. 1996. Faunal stability during the Early Oligocene climatic crash. Palaeogeography, Palaeoclimatology, Palaeoecology 127:257283.Google Scholar
Prothero, D. R., and Shubin, N. 1989. The evolution of Oligocene horses. Pp. 142175 in Prothero, D. R. and Schoch, R. M., eds. The evolution of perissodactyls. Oxford University Press, New York.Google Scholar
Purugganan, M. D. 1998. The molecular evolution of development. BioEssays 20:700711.Google Scholar
Purugganan, M. D. 2000. The molecular population genetics of regulatory genes. Molecular Ecology (in press).Google Scholar
Quinn, J. C., West, J. D., and Hill, R. E. 1996. Multiple functions for Pax6 in mouse eye and nasal development. Genes and Development 10:435446.Google Scholar
Rabinowitz, D. 1981. Seven forms of rarity. Pp. 205217 in Synge, H., ed. The biological aspects of rare plant conservation. Wiley, New York.Google Scholar
Rabinowitz, D., Cairns, S., and Dillon, T. 1986. Seven forms of rarity and their frequency in the flora of the British Isles. Pp. 182204 in Soulé, M. E., ed. Conservation biology. Sinauer, Sunderland, Mass.Google Scholar
Rachootin, S. P., and Thomson, K. S. 1981. Epigenetics, paleontology, and evolution. Pp. 181193 in Scudder, G. G. E. and Reveal, J. L., eds. Evolution today. Proceedings of the second international congress of systematic and evolutionary biology. Hunt Institute for Botanical Documentation, Carnegie-Mellon University, Pittsburgh, Penn.Google Scholar
Raff, R. A. 1996. The shape of life: genes, development, and the evolution of animal form. University of Chicago Press, Chicago.Google Scholar
Raff, R. A., and Kaufman, T. C. 1983. Embryos, genes, and evolution. Macmillan, New York (reprinted 1991, Indiana University Press, Bloomington).Google Scholar
Raff, R. A., and Wray, G. A. 1989. Heterochrony: developmental mechanisms and evolutionary results. Journal of Evolutionary Biology 2:409434.Google Scholar
Raup, D. M. 1983. On the early origins of major biologic groups. Paleobiology 9:107115.Google Scholar
Raup, D. M. 1991. A kill curve for Phanerozoic marine species. Paleobiology 17:3748.Google Scholar
Raup, D. M. 1994. The role of extinction in evolution. Proceedings of the National Academy of Sciences USA 91:67586763.Google Scholar
Relaix, F., and Buckingham, M. 1999. From insect eye to vertebrate muscle: redeployment of a regulatory network. Genes and Development 13:31713178.Google Scholar
Rice, S. H. 1995. A genetical theory of species selection. Journal of Theoretical Biology 177:237245.Google Scholar
Rice, S. H. 1998. The evolution of canalization and the breaking of von Baer's laws: modeling the evolution of development with epistasis. Evolution 52:647656.Google Scholar
Ricklefs, R. E., and Latham, R. E. 1992. Intercontinental correlation of geographic ranges suggests stasis in ecological traits of relict genera of temperate perennial herbs. American Naturalist 139:13051321.Google Scholar
Riedl, R. 1978. Order in living organisms. Wiley, Chichester, England.Google Scholar
Robertson, F. W. 1959. Studies in quantitative inheritance. XIII. Interrelations between genetic behavior and development in the cellular constitution in the Drosophila wing. Genetics 44:11131130.Google Scholar
Rogers, B. T., Peterson, M. D., and Kaufman, T. C. 1997. Evolution of the insect body plan as revealed by the Sex combs reduced expression pattern. Development 124:149157.Google Scholar
Roopnarine, P. D., Byars, G., and Fitzgerald, P. 1999. Anagenetic evolution, stratophenetic patterns, and random walk models. Paleobiology 25:4157.Google Scholar
Rosenzweig, M. L., and McCord, R. D. 1991. Incumbent replacement: evidence for long-term evolutionary progress. Paleobiology 17:202213.Google Scholar
Routman, E. J., and Cheverud, J. M. 1997. Gene effects on a quantitative trait: two-locus epistatic effects measured at microsatellite markers and at estimated QTL. Evolution 51:16541662.Google Scholar
Roy, K. 1996. The roles of mass extinction and biotic interaction in large-scale replacements: a reexamination using the fossil record of stromboidean gastropods. Paleobiology 22:436452.Google Scholar
Roy, K., Valentine, J. W., Jablonski, D., and Kidwell, S. M. 1996. Scales of climatic variability and time averaging in Pleistocene biotas: implications for ecology and evolution. Trends in Ecology and Evolution 11:458463.Google Scholar
Rugh, N. S. 1997. Differences in shell morphology between the sibling species Littorina scutulata and Littorina plena (Gastropoda: Prosobranchia). Veliger 40:350357.Google Scholar
Rumrill, S. S. 1990. Natural mortality of marine invertebrate larvae. Ophelia 32:163198.Google Scholar
Rutherford, S. L., and Lindquist, S. 1998. Hsp 90 as a capacitor for morphological evolution. Nature 396:336342.Google Scholar
Salmon, W. C. 1971. Statistical explanation and statistical relevance. University of Pittsburgh Press, Pittsburgh, Penn.Google Scholar
Sanderson, M. J., and Donoghue, M. J. 1996. Reconstructing shifts in diversification rates on phylogenetic trees. Trends in Ecology and Evolution 11:1520.Google Scholar
Sato, S. 1995. Spawning periodicity and shell microgrowth patterns of the venerid bivalve Phacosoma japonicum (Reeve, 1850). Veliger 38:6172.Google Scholar
Sato, S. 1999. Temporal change of life-history traits in fossil bivalves: an example of Phacosoma japonicum from the Pleistocene of Japan. Palaeogeography, Palaeoclimatology, Palaeoecology 154:313323.Google Scholar
Saunders, W. B., Work, D. M., and Nikolaeva, S. V. 1999. Evolution of complexity in Paleozoic ammonoid sutures. Science 286:760763.Google Scholar
Schankler, D. M. 1981. Local extinction and ecological re-entry of early Eocene mammals. Nature 293:135138.Google Scholar
Scheltema, R. S. 1989. Planktonic and non-planktonic development among prosobranch gastropods and its relationship to the geographic range of species. Pp. 183188 in Ryland, J. S. and Tyler, P. A., eds. Reproduction, genetics and distribution of marine organisms. Olsen and Olsen, Fredensborg, Denmark.Google Scholar
Scheltema, R. S. 1992. Passive dispersal of planktonic larvae and the biogeography of tropical sublittoral invertebrate species. Pp. 195202 in Colombo, G. et al., eds. Marine eutrophication and population dynamics. Olsen and Olsen, Fredensborg, Denmark.Google Scholar
Schemske, D. W., and Bradshaw, H. D. 1999. Pollinator preference and the evolution of floral traits in monkeyflowers (Mimulus). Proceedings of the National Academy of Sciences USA 96:1191011915.Google Scholar
Schlichting, C. D., and Pigliucci, M. 1998. Phenotypic evolution: a reaction norm perspective. Sinauer, Sunderland, Mass.Google Scholar
Schluter, D. 1986. Character displacement between distantly related taxa? Finches and bees in the Galapagos. American Naturalist 127:95102.Google Scholar
Schluter, D. 1996. Adaptive radiation along genetic lines of least resistance. Evolution 50:17661774.Google Scholar
Schneider, J. A. 1995. Phylogenetic relationships of transisthmian Cardiidae (Bivalvia) and the usage of fossils in reinterpreting the geminate species concept. Geological Society of America Abstracts with Programs 27(6):A52.Google Scholar
Sepkoski, J. J. Jr. 1996. Competition in macroevolution: the double wedge revisited. Pp. 211255 in Jablonski, D., Erwin, D. H., and Lipps, J. H., eds. Evolutionary paleobiology. University of Chicago Press, Chicago.Google Scholar
Sepkoski, J. J. Jr. 1998. Rates of speciation in the fossil record. Philosophical Transactions of the Royal Society of London B 353:315326.Google Scholar
Sepkoski, J. J. Jr., McKinney, F. K., and Lidgard, S. 2000. Competitive displacement among post-Paleozoic cyclostome and cheilostome bryozoans. Paleobiology 26:718.Google Scholar
Shackleton, N. J., McCave, I. N., and Weedon, G. P., eds. 1999. Astronomical (Milankovitch) calibration of the geological time-scale. Philosophical Transactions of the Royal Society of London A 357:17332007.Google Scholar
Shaffer, H. B. 1984. Evolution in a paedomorphic lineage. I. An electrophoretic analysis of the Mexican ambystomatid salamanders. Evolution 38:11941206.Google Scholar
Shaw, F. H., Shaw, R. G., Wilkinson, G. S., and Turelli, M. 1995. Changes in genetic variances and covariances: G whiz! Evolution 49:12601267.Google Scholar
Sheldon, P. R. 1987. Parallel gradualistic evolution in Ordovician trilobites. Nature 330:561563.Google Scholar
Sheldon, P. R. 1988. Trilobite size-frequency distributions, recognition of instars, and phyletic size change. Lethaia 21:293306.Google Scholar
Sheldon, P. R. 1993. Making sense of microevolutionary patterns. Pp. 1931 in Lees, D. R. and Edwards, D., eds. Evolutionary patterns and processes. Academic Press, London.Google Scholar
Sheldon, P. R. 1996. Plus ça change—a model for stasis and evolution in different environments. Palaeogeography, Palaeoclimatology, Palaeoecology 127:209227.Google Scholar
Shubin, N., Tabin, C., and Carroll, S. 1997. Fossil, genes and the evolution of animal limbs. Nature 388:639648.Google Scholar
Signor, P. W. III, and Brett, C. E. 1984. The mid-Paleozoic precursor to the Mesozoic marine revolution. Paleobiology 10:229245.Google Scholar
Simms, M. J. 1988a. The role of heterochrony in the evolution of post-Paleozoic crinoids. Pp. 97102 in Burke, R. D., Mladenov, P. V., Lambert, P., and Parsley, R. L., eds. Echinoderm biology. Balkema, Rotterdam.Google Scholar
Simms, M. J. 1988b. The phylogeny of post-Palaeozoic crinoids. Pp. 269284 in Paul, C. R. C. and Smith, A. B., eds. Echinoderm phylogeny and evolution. Clarendon, Oxford.Google Scholar
Simms, M. J. 1994. Crinoids from the Chambara Formation, Pucara Group, central Peru. Palaeontographica, Abteilung A 233:169175.Google Scholar
Simpson, G. G. 1961. Principles of animal taxonomy. Columbia University Press, New York.Google Scholar
Sinervo, B., and McEdward, L. R. 1988. Developmental consequences of an evolutionary change in egg size: an experimental test. Evolution 42:885899.Google Scholar
Slatkin, M. 1981. A diffusion model of species selection. Paleobiology 7:421425.Google Scholar
Slowinski, J. B., and Guyer, C. 1993. Testing whether certain traits have caused amplified diversification: an improved method based on a model of random speciation and extinction. American Naturalist 142:10191024.Google Scholar
Smith, A. B. 1994. Systematics and the fossil record. Blackwell Scientific, Oxford.Google Scholar
Smith, G. R. 1992. Introgression in fishes: significance for paleontology, cladistics, and evolutionary rates. Systematic Biology 41:4157.Google Scholar
Smith, K. K. 1996. Integration of craniofacial structures during development in mammals. American Zoologist 36:7079.Google Scholar
Smith, K. K. 1997. Comparative patterns of craniofacial development in eutherian and metatherian mammals. Evolution 51:16631678.Google Scholar
Sneppan, K., Bak, P., Flybjerg, H., and Jensen, M. H. 1995. Evolution as a self-organized critical phenomenon. Proceedings of the National Academy of Sciences USA 92:52095213.Google Scholar
Sober, E. 1984. The nature of selection. MIT Press, Cambridge.Google Scholar
Sober, E. 1999. The multiple realizability argument against reductionism. Philosophy of Science 66:542564.Google Scholar
Solé, R. V., Manrubia, S. C., Benton, M., Kauffman, S., and Bak, P. 1999. Criticality and scaling in evolutionary ecology. Trends in Ecology and Evolution 14:156160.Google Scholar
Spencer-Cervato, C., Thierstein, H. R., Lazarus, D. B., and Beckmann, J.-P. 1994. How synchronous are Neogene marine plankton events? Paleoceanography 9:739763.Google Scholar
Stanley, S. M. 1973. An explanation for Cope's Rule. Evolution 27:126.Google Scholar
Stanley, S. M. 1979. Macroevolution. W. H. Freeman, San Francisco.Google Scholar
Stanley, S. M. 1982. Macroevolution and the fossil record. Evolution 36:460473.Google Scholar
Stanley, S. M. 1990. The general correlation between rate of speciation and rate of extinction: fortuitous causal linkages. Pp. 103127 in Ross, R. M. and Allmon, W. D., eds. Causes of evolution. University of Chicago Press, Chicago.Google Scholar
Stanley, S. M., and Yang, X. 1987. Approximate evolutionary stasis for bivalve morphology over millions of years: a multivariate, multilineage study. Paleobiology 13:113139.Google Scholar
Stanley, S. M., Wetmore, K. L., and Kennett, J. P. 1988. Macroevolutionary differences between two major clades of Neogene planktonic Foraminifera. Paleobiology 14:235249.Google Scholar
Stearns, S. C. 1992. The evolution of life histories. Oxford University Press, Oxford.Google Scholar
Stebbins, G. L. 1982. Perspectives in evolutionary theory. Evolution 36:11091119.Google Scholar
Stehli, F. G., Douglas, R. G., and Newell, N. D. 1969. Generation and maintenance of gradients in taxonomic diversity. Science 164:947949.Google Scholar
Steppan, S. J. 1998. Phylogenetic relationships and species limits within Phyllotis (Rodentia: Sigmodontinae): concordance between mtDNA sequence and morphology. Journal of Mammalogy 79:573593.Google Scholar
Sterelny, K., and Griffiths, P. E. 1999. Sex and death: an introduction to philosophy of biology. University of Chicago Press, Chicago.Google Scholar
Stevenson, R. D., Hill, M. F., and Bryant, P. J. 1995. Organ and cell allometry in Hawaiian Drosophila: how to make a big fly. Proceedings of the Royal Society of London B 259:105110.Google Scholar
Stidd, B. M., and Wade, D. L. 1995. Is species selection dependent upon emergent characters? Biology and Philosophy 10:5576.Google Scholar
Tabachnick, R. E., and Bookstein, F. L. 1990. The structure of individual variation in Miocene Globorotalia. Paleobiology 44:416434.Google Scholar
Templeton, A. R. 1989. The meaning of species and speciation: a genetic perspective. Pp. 327 in Otte and Endler 1989.Google Scholar
Templeton, A. R. 1998. Species and speciation: Geography, population structure, ecology, and gene trees. Pp. 3243 in Howard and Berlocher 1998.Google Scholar
Theissen, G., Becker, A., Rosa, A. Di, Kanno, A., Kim, J. T., Münster, T., Winter, K.-U., and Saedler, H. 2000. A short history of MADS-box genes in plants. Plant Molecular Biology 42:115149.Google Scholar
Theriot, E. 1992. Clusters, species concepts, and morphological evolution of diatoms. Systematic Biology 41:141157.Google Scholar
Thomas, R. D. K., Shearman, R. M., and Stewart, G. W. 2000. Evolutionary exploitation of design options by the first animals with hard skeletons. Science 288:12391242.Google Scholar
Thomson, K. S. 1988. Morphogenesis and evolution. Oxford University Press, New York.Google Scholar
Trussell, G. C., and Smith, L. D. 2000. Induced defenses in response to an invading crab predator: an explanation of historical and geographic phenotypic change. Proceedings of the National Academy of Sciences USA 97:21232127.Google Scholar
Turelli, M. 1988. Phenotypic evolution, constant covariances, and the maintenance of additive variance. Evolution 42:13421347.Google Scholar
Valentine, J. W. 1969. Taxonomic and ecological structure of the shelf benthos during Phanerozoic time. Palaeontology 12:684709.Google Scholar
Valentine, J. W. 1973. Evolutionary paleoecology of the marine biosphere. Prentice-Hall, Englewood Cliffs, N.J. Google Scholar
Valentine, J. W. 1980. Determinants of diversity in higher taxonomic categories. Paleobiology 6:444450.Google Scholar
Valentine, J. W. 1990. The fossil record: a sampler of life's diversity. Philosophical Transactions of the Royal Society of London B 330:251268.Google Scholar
Valentine, J. W. 1995. Why no new phyla after the Cambrian? Genome and ecospace hypotheses revisited. Palaios 10:190194.Google Scholar
Valentine, J. W. 2000. Two paths to complexity in metazoan evolution. Paleobiology 26:513519.Google Scholar
Valentine, J. W., and Jablonski, D. 1983. Speciation in the shallow sea: General patterns and biogeographic controls. In Sims, R. W., Price, J. H., and Whalley, P. E. S., eds. Evolution, time and space. Systematics Association Special Volume 23:201226. Academic Press, London.Google Scholar
Valentine, J. W., and Jablonski, D. 1993. Fossil communities: compositional variation at many time scales. Pp. 341349 in Ricklefs, R. E. and Schluter, D., eds. Species diversity in ecological communities: historical and geographical perspectives. University of Chicago Press, Chicago.Google Scholar
Valentine, J. W., Jablonski, D., and Erwin, D. H. 1999. Fossils, molecules and embryos: new perspectives on the Cambrian explosion. Development 126:851859.Google Scholar
Van Valen, L. 1976. Ecological species, multispecies, and oaks. Taxon 25:233239.Google Scholar
Vermeij, G. J. 1987. Evolution and escalation. Princeton University Press, Princeton, N.J. Google Scholar
Vermeij, G. J. 1991. When biotas meet: understanding biotic interchange. Science 253:10991104.Google Scholar
Vermeij, G. J. 1994. The evolutionary interaction among species: selection, escalation, and coevolution. Annual Review of Ecology and Systematics 25:219236.Google Scholar
Via, S., and Hawthorne, D. J. 1998. The genetics of speciation: promises and prospects of Quantitative Trait Locus mapping. Pp. 352354 in Howard and Berlocher 1998.Google Scholar
Via, S., and Shaw, A. J. 1996. Short-term evolution in size and shape of pea aphids. Evolution 50:163173.Google Scholar
Voss, S. R., and Shaffer, H. B. 1997. Adaptive evolution via a major gene effect: paedomorphosis in the Mexican axolotl. Proceedings of the National Academy of Sciences USA 94:1418514189.Google Scholar
Vrba, E. S. 1984. What is species selection? Systematic Zoology 33:318328.Google Scholar
Vrba, E. S. 1987. Ecology in relation to speciation rates: some case histories of Miocene-Recent mammal clades. Evolutionary Ecology 1:283300.Google Scholar
Vrba, E. S. 1989. Levels of selection and sorting with special reference to the species level. Oxford Surveys in Evolutionary Biology 6:111168.Google Scholar
Vrba, E. S., and Eldredge, N. 1984. Individuals, hierarchies and processes: towards a more complete evolutionary theory. Paleobiology 10:146171.Google Scholar
Vrba, E. S., and Gould, S. J. 1986. The hierarchical expansion of sorting and selection: sorting and selection cannot be equated. Paleobiology 12:217228.Google Scholar
Wagner, G. P. 1996. Homologues, natural kinds and the evolution of modularity. American Zoologist 36:3643.Google Scholar
Wagner, G. P., Booth, G., and Bagheri-Chaichian, H. 1997. A population genetic theory of canalization. Evolution 51:329347.Google Scholar
Wagner, G. P., Chiu, C. H., and Hansen, T. H. 1999. Is Hsp 90 a regulator of evolvability? Journal of Experimental Zoology (Molecular and Developmental Evolution) 285:116118.Google Scholar
Wagner, P. J. 1995. Testing evolutionary constraint hypotheses: examples with early Paleozoic gastropods. Paleobiology 21:248272.Google Scholar
Wagner, P. J. 1996a. Ghost taxa, ancestors, and assumptions: A reply to Norell. Paleobiology 22:456460.Google Scholar
Wagner, P. J. 1996b. Contrasting the underlying patterns of active trends in morphologic evolution. Evolution 50:9901007.Google Scholar
Wagner, P. J. 1998. A likelihood approach for evaluating estimates of phylogenetic relationships among fossil taxa. Paleobiology 24:430449.Google Scholar
Wagner, P. J., and Erwin, D. H. 1995. Phylogenetic patterns as tests of speciation models. Pp. 87122 in Erwin and Anstey 1995b.Google Scholar
Walsh, J. A. 1996. No second chances? New perspectives on biotic interactions in post-Paleozoic brachiopod history. Pp. 281288 in Copper, P. and Jin, J., eds. Brachiopods. Balkema, Rotterdam.Google Scholar
Watts, P. C., Thorpe, J. P., and Taylor, P. D. 1998. Natural and anthropogenic dispersal mechanisms in the marine environment: a study using cheilostome Bryozoa. Philosophical Transactions of the Royal Society of London B 353:453464.Google Scholar
Wayne, R. K. 1986. Cranial morphology of domestic and wild canids: the influence of development on morphological change. Evolution 40:243261.Google Scholar
Wayne, R. K., and Ostrander, E. A. 1999. Origin, genetic diversity, and genome structure of the domestic dog. BioEssays 21:247257.Google Scholar
Weatherbee, S. D., Nijhout, H. F., Grunert, L. W., Halder, G., Galant, R., Selegue, J., and Carroll, S. 1999. Ultrabithorax function in butterfly wings and the evolution of insect wing patterns. Current Biology 9:109115.Google Scholar
Wei, K.-Y., and Kennett, J. P. 1986. Taxonomic evolution of Neogene planktonic Foraminifera and paleoceanographic relations. Paleoceanography 1:6784.Google Scholar
Weil, E., and Knowlton, N. 1994. A multi-character analysis of the Caribbean coral Montastraea annularis (Ellis and Solander, 1786) and its two sibling species, M. faveolata (Ellis and Solander, 1786) and M. franksi (Gregory, 1895). Bulletin of Marine Science 55:151175.Google Scholar
Wen, J. 1999. Evolution of the eastern Asian and eastern North American disjunct distributions in flowering plants. Annual Review of Ecology and Systematics 30:421455.Google Scholar
Wiley, E. O. 1981. Phylogenetics. Wiley, New York.Google Scholar
Williams, G. C. 1992. Natural selection: domains, levels, and challenges. Oxford University Press, New York.Google Scholar
Williamson, M. 1996. Biological invasions. Chapman and Hall, London.Google Scholar
Wilson, A. C., Bush, G. L., Case, S. M., and King, M.-C. 1975. Social structuring of mammalian populations and rate of chromosomal evolution. Proceedings of the National Academy of Sciences USA 72:50615065.Google Scholar
Wing, S. L., and Boucher, L. D. 1998. Ecological aspects of the Cretaceous flowering plant radiation. Annual Review of Earth and Planetary Sciences 26:379421.Google Scholar
Wright, S. 1982a. Character change, speciation and higher taxa. Evolution 36:427443.Google Scholar
Wright, S. 1982b. The shifting balance theory and macroevolution. Annual Review of Genetics 16:119.Google Scholar
Zakany, J., Fromental-Ramain, C., Wartot, X., and Duboule, D. 1997. Regulation of number and size of digits by posterior Hox genes: a dose-dependent mechanism with potential evolutionary implications. Proceedings of the National Academy of Sciences USA 94:1369513700.Google Scholar
Zelditch, M. L., and Fink, W. L. 1996. Heterochrony and heterotopy: stability and innovation in the evolution of form. Paleobiology 22:241254.Google Scholar