Skip to main content
Log in

Pattern, process, and natural disturbance in vegetation

  • Published:
The Botanical Review Aims and scope Submit manuscript

Abstract

Natural disturbances have been traditionally defined in terms of major catastrophic events originating in the physical environment and, hence, have been regarded as exogenous agents of vegetation change. Problems with this view are: (1) there is a gradient from minor to major events rather than a uniquely definable set of major catastrophes for each kind of disturbance, and (2) some disturbances are initiated or promoted by the biotic component of the system. Floras are rich in disturbance-adapted species. Disturbances have probably exerted selective pressure in the evolution of species strategies.

Heathland cyclic successions and gap-phase dynamics in forests have been viewed as endogenous patterns in vegetation. When death in older individividuals imposes a rhythm on community reproduction, dynamics may indeed be the result of endogenous factors. However, documented cases of senescence in perennial plants are few and many cyclic successions and cases of gap-phase dynamics are initiated by physical factors. Forest dynamics range from those that are the result of individual tree senescence and fall, through those that are the result of blowdown of small groups of healthy trees, to those that are the result of large windstorms which level hectares of forest. The effect of wind ranges from simple pruning of dead plant parts to widespread damage of living trees. Wind speed is probably inversely proportional to occurrence frequency. Disturbances vary continuously. There is a gradient from those community dynamics that are initiated by endogenous factors to those initiated by exogenous factors. Evolution has mediated between species and environment; disturbances are often caused by physical factors but the occurrence and outplay of disturbances may be a function of the state of the community as well.

Natural disturbances in North American vegetation are: fire, windstorm, ice storm, ice push on shores, cryogenic soil movement, temperature fluctuation, precipitation variability, alluvial processes, coastal processes, dune movement, saltwater inundation, landslides, lava flows, karst processes, and biotic disturbances. Disturbances vary regionally and within one landscape as a function of topography and other site variables and are characterized by their frequency, predictability, and magnitude. The landscape level is important in assessing disturbance regime. Disturbances and cyclic successions belong to the same class of events—that of recurrent dynamics in vegetation structure—irrespective of cause. Dynamics may result from periodic, abrupt, and catastrophic environmental factors or they may result from an interaction of the changing susceptability of the community and some regular environmental factor. In any case, the dynamics result in heterogeneous landscapes; the species adapted to this heterogeneity are numerous, suggesting their long time importance.

The importance of disturbance regime as part of the environmental context of vegetation means that allogenic and autogenic models of vegetation are difficult to apply. Species composition can be seen to be a function of disturbance regime, as well as other environmental variables. Competitive replacement in succession occurs, then, only as disturbances cease to operate and can be viewed as allogenic adjustment to a new disturbance-free environment. Competitive divergence, separation of role, and competition avoidance may, in fact, underlie successional patterns traditionally viewed as the competitive replacement of inferior species by superiorly adapted climax species.

The importance of ongoing dynamics is also difficult to reconcile with the concept of climax, founded as it is on the idea of autogenesis within a stable physical environment. Climax composition is relative to disturbance regime. Climax is only arbitrarily distinguished from succession. Climax as an organizing paradigm in plant ecology has obscured the full temporal-spatial dimensions important in understanding the vegetated landscape and the evolution of species which contribute to the landscape patterns.

Whittaker’s coenocline concept is accepted with modifications: (1) natural disturbance gradients and Whittaker’s complex gradient are intimately related, (2) temporal variation in the community should be viewed as an added axis of community pattern, and (3) ongoing dynamics have important effects on specificity of species to site relations and the predictability of vegetation patterns. Recent work has suggested an r-K continuum in species strategy. In general, colonizing ability is seen as a trade-off against specialization. Frequent disruption of the community and the creation of open sites seems to result in mixes of species that are fleeting in time and do not repeat in space. Species in such mixes are often tolerant of wide environmental extremes but are compressed into early successional time if disturbance ceases. The composition of such communities is not predictable from site characteristics. Even communities with low disturbance frequency lack complete environmental determinism, and historical events are important in understanding present composition. Communities vary in level of environmental determinism and species differ in niche breadth and degree of site specificity. Management implications of vegetation dynamics are discussed.

Résumé

Les perturbations naturelles se définissent traditionnellement en fonction des événements catastrophiques majeurs dont l’origine se trouve dans l’environnement physique et que, de ce fait, l’on considère comme des agents exogènes des changements de la végétation. On peut opposer à ce point de vue les arguments suivants: (1) qu’il existe une gradation entre les événements mineurs et les événements majeurs au lieu d’un ensemble définissable de catastrophes majeurs pour chaque type de perturbation et (2) que quelques perturbations sont initiées ou promues par le composant biotique du système. Les flores sont riches en espèces adaptées aux perturbations. Les perturbations ont sans doute exercé une pression sélective sur l’évolution des stratégies que possèdent les espèces.

Les successions cyliques des landes et la dynamique du vide dans les forêts se sont considérées comme des configurations endogènes de végétation. Quand la mort des individus plus anciens impose un rythme sur la reproduction de la communauté, la dynamique peut bien être le résultat des facteurs endogènes. Cependant, il existe peu de cas documentés de la sénescence des plantes vivaces et bien des successions cyliques, et des cas de dynamique du vide sont initiés par des facteurs physiques. Les dynamiques des forêts s’étendent de celles qui résultent de la sénescence et de la chute d’un arbre individuel, à celles qui résultent de l’abattage par le vent de petits groupes d’arbres sains, jusqu’à celles qui résultent de grandes tempêtes qui nivellent des hectares de forêts. L’effet du vent va du simple élagage des parties mortes des plantes jusqu’aux dommages importants aux arbres vivants. Il est probable que la vitesse du vent est en rapport inverse avec sa fréquence. Les perturbations varient sans cesse. Il y a une gradation à partir des dynamiques de communauté initiées par les facteurs endogènes jusqu’à celles initiées par des facteurs exogènes. L’évolution a joué le rôle d’intermédiaire entre les espèces et l’environnement; les perturbations sont souvent dûes aux facteurs physiques mais il se peut bien que les perturbations aient lieu et se terminent en fonction également de l’état de la communauté.

Les perturbations naturelles dans la végétation de l’Amérique du Nord sont les suivantes: le feu, les tempêtes, les tempêtes de glace, la poussée de la glace sur les côtes, le mouvement cryogène des sols, la fluctuation des températures, la variabilité de la précipitation, les procédés alluviaux, les procédés côtiers, le mouvement des dunes, l’inondation par les eaux de mer, les éboulements de terre, le flux de lave, les procédés karstiques, et les perturbations biotiques. Ces perturbations varient suivant la région et à l’intérieur d’un même paysage en fonction de la topographie et d’autres inconstances du site et se caractérisent par leur fréquence, par leur prévisibilité et par leur grandeur. Le niveau du paysage a une importance pour l’évaluation du régime de la perturbation. Leur cause mise à part, les perturbations et les successions cycliques appartiennent à la même catégorie d’évenements—celle de la dynamique périodique dans la structure de la végétation. La dynamique peut être le résultat des facteurs périodiques, brusques et catastrophiques dans l’environnement ou d’une action réciproque entre le susceptibilité changeante de la communauté et un facteur régulier dans l’environnement. En tout cas, des paysages hétérogènes en résultent; les espèces adaptées à cette hétérogéné ité sont nombreuses, ce qui suggère leur signification à long terme.

L’importance du régime de perturbation comme partie du contexte environnant de la végétation signifie qu’il est difficile d’appliquer les modèles allogènes et autogènes de végétation. La composition de l’espèces peut se voir comme une fonction du régime des perturbations, ainsi que d’autres inconstants de l’environnement. Le remplacement compétitif dans la succession n’a donc lieu que lorsque les perturbations cessent d’exister et peut s’interpréter comme l’adaptation allogène à un nouveau environnement exempte de perturbations. La divergence compétitive, la séparation des rôles et le refus de la compétition peuvent, en fait, être à la base des configurations de succession considérées traditionnellement comme le remplacement compétitif d’une espèce inférieure par une espèce ortho-écologique adaptée de façon supérieure.

De plus, l’importance de la dynamique continue est difficile à concilier avec l’idée d’ortho-écologie qui est fondée sur l’idée de l’autogenèse dans un environnement physique stable. La composition ortho-écologique se rapporte au régime des perturbations. L’ortho-écologie ne se distingue qu’arbitrairement de la succession. L’ortho-écologie en tant que paradigme organisateur dans l’écologie des plantes a obscurci l’ensemble des dimensions temporelles et spatiales important à la compréhension du paysage végété et de l’évolution des espèces qui contribuent aux configurations du paysage.

Nous acceptons avec modifications le concept de cénocline (“coenocline”) de Whittaker: (1) les gradations des perturbations naturelles et la gradation complexe de Whittaker sont intimement liées, (2) la variation temporelle dans une communauté doit se voir comme une dimension de plus dans la configuration d’une communauté, et (3) la dynamique continue a des effets importants sur la spécificité des rapports espèces-site et sur la prévisibilité des configurations de végétation. Des travaux récents suggèrent un continuum r-K dans la stratégie des espèces. En général, la capacité de coloniser est interprétée comme un échange contre la spécialisation. La perturbation fréquente de la communauté et la création des sites ouverts paraît avoir comme résultat des mélanges d’espèces qui sont éphémères et qui ne se répètent pas dans l’espace. Les espèces dans de tels mélanges sont souvent très tolérantes de grandes extrèmes dans l’environnement mais se compriment tôt en temps de succession si la perturbation cesse. La composition de telles communautés ne peut pas se prédire suivant les caractéristiques du site. Même les communautés qui subissent une basse fréquence de perturbations manquent le plein déterminisme de l’environnement et les événements historiques sont importants à la compréhension de la composition actuelle. Les communautés varient quant au niveau de déterminisme de l’environnement et les espèces diffèrent quant à l’étendu de la niche et au dégré de spécificité au site. L’importance de la dynamique de végétation pour le management écologique est discutée.

Zusammenfassung

Natürliche Störungen werden traditionsgemäß im Sinne von großen, in der physikalischen Umwelt entstehenden katastrophalen Ereignissen definiert und daher als exogene Erreger von Vegetationsveränderungen angesehen. Problematisch an dieser Auffassung ist folgendes: (1) Es gibt keinen einmalig bestimmbaren Satz von Hauptkatastrophen für jede Störungsart, sondern einen Gradienten von kleinen bis zu großen Ereignissen, und (2) einige Störungen werden von der biotischen Komponente des Systems eingeleitet oder gefördert. Die Floren sind reich an störungsangepaßten Arten. Störungen haben wahrscheinlich einen Selektionsdruck bei der Evolution der Artenstrategien ausgeübt.

Die zyklische Sukzession im Heideland und die Dynamik der Bestandeslückenphase in Wäldern sind als endogene Vegetationsmuster betrachtet worden. Wenn das Absterben von älteren Individuen der Fortpflanzung der Pflanzengesellschaft einen Rhythmus auferlegt, dürfte die Dynamik allerdings das Resultat endogener Einflüsse sein. Die belegten Fälle von Altern in perennierenden Pflanzen sind jedoch wenige, und viele zyklische Sukzessionen und Fälle der Dynamik der Bestandeslückenphase werden durch physikalische Einflüsse hervorgerufen. Die Walddynamik umfaßt die Einflüsse, die sich aus den folgenden Situationen ergeben: Altern und Fall von einzelnen Bäumen, Windbruch von kleinen Gruppen gesunder Bäume und schließlich große Stürme, die Hektare von Wald ebene. Die Wirkung des Windes umfaßt sowohl das einfache Abbrechen von toten Pflanzenteilen als auch ausgedehnte Schäden an lebenden Bäumen. Die Windgeschwindigkeit ist der Auftrittshäufigkeit wohl verkehrt proportional. Störungen variieren kontinuierlich. Es existiert ein Gradient von der von endogenen Einflüssen eingeleiteten zu der von exogenen Einflüssen eingeleiteten Gesellschaftsdynamik. Die Evolution hat zwischen den Arten und der Umwelt vermittelt. Störungen werden oft durch physikalische Einflüsse verursacht, aber das Auftreten und das Abspielen von Störungen dürfte auch von dem Stand der Gesellschaft abhängen.

Natürliche Störungen in der Vegetation Nordamerikas sind Brand, Sturm, Eissturm, Eisschub an Ufern, Tieftemperatur-Bodenbewegung, Temperaturschwankung, Niederschlagsveränderlichkeit, alluviale und Küstenvorgänge, Dünenbewegung, Salzwasserpberflutung, Erdrutsche, Lavaströme, Karstvorgänge und biotische Störungen. Die Störungen variieren regional und innerhalb einer Landschaft in Abhängigkeit von der Oberflächengestaltung und anderen Standortsvariablen und werden durch ihre Häufigkeit, Vorausbestimmbarkeit und Größenordnung gekennzeichnet. Die Höhe des Geländes ist wichtig bei der Bewertung der Störungsbedingungen. Störungen und zyklische Sukzessionen gehören—unabhängig von der Ursache—der gleichen Ereignisklasse an: der wiederkehrenden Dynamik in der Vegetationsstruktur. Die Dynamik mag sich aus periodischen, abrupten und katastrophenartigen Umwelteinflüssen ergeben, oder aber sie mag aus der Wechselwirbung der sich ändernden Anfälligkeit der Gesellschaft und eines regelmäßigen Umweltfaktors resultieren. Auf jeden Fall führt die Dynamik zu heterogenen Landschaften. Die Arten, die sich dieser Heterogenität angepaßt haben, sind zahlreich, was auf ihre langwährende Bedeutung hinweist.

Die Wichtigkeit der Störungsbedingungen im Rahmen der Vegetationsumwelt deutet an, daß allogene und autogene Vegetationsmodelle schwer anzuwenden sind. Die Artenzusammensetzung kann in Abhängigkeit sowohl von Störungsbedingungen als auch von anderen Umweltvariablen betrachtet werden. Eine Konkurrenzersetzung in der Sukzession tritt dann nur auf, wenn die Störungen zu wirken aufhören, und kann als die allogenische Anpassung an eine neue störungsfreie Umwelt verstanden werden. Konkurrenz-Divergenz, Rollentrennung und Konkurrenzvermeidung dürften tatsächlich den Sukzessionsmustern zugrundeliegen, die traditionsgemäß als die Konkurrenzersetzung von unterlegenen Arten durch überlegen angepaßte Klimaxarten verstanden werden.

Es ist auch schwer, die Bedeutung der fortwährenden Dynamik mit dem Begriff des Klimaxes in Einklang zu bringen, da letzterer auf den Begriff der Autogenese innerhalb einer stabilen physikalischen Umwelt begründet ist. Die Klimaxzusammensetzung ist in Bezug auf die Störungsbedingungen relativ. Klimax unterscheidet sich von Sukzession nur arbiträr. Die Klimax als Organisationsparadigma in der Pflanzenökologie hat die vollen zeitlich-räumlichen Dimensionen unklar gemacht, die für ein Verständnis der Vegetationslandschaft und der Evolution der zu den Landschaftsmustern beitragenden Arten wichtig sind.

Der Begriff des “coenocline” von Whittaker wird mit einigen Abänderungen angenommen: (1) Natürliche Störungsgradiente und der komplexe Gradient von Whittaker sind eng miteinander verwandt, (2) eine zeitliche Variation in der Gesellschaft sollte als zusätzliche Achse der Gesellschaftsmuster betrachtet werden und (3) die fortwährende Dynamik hat einen wichtigen auf die Artenspezifität gegenüber den Standortsverhältnissen und auf die Vorausbestimmbarkeit von Vegetationsmustern. Neuere Arbeiten weisen auf ein r-K-Continuum in der Artenstrategie hin. Im allgemeinen nimmt man an, die Fähigkeit zur Koloniebildung existiere auf Kosten der Spezialisierung. Das häufige Auseinanderreißen der Gesellschaft und die Schaffung lichter Standorte resultiert anscheinend in Artmischungen, die kurzfristig auftreten und sich räumlich nicht wiederholen. Die Arten in solchen Mischungen tolerieren oft weite Umweltextreme, werden aber in eine frühe Sukzessionszeit zusammengedrängt, wenn die Störung sich legt. Die Zusammensetzung solcher Gesellschaften kann nicht auf Grund der Standortskenngrößen vorhergesagt werden. Sogar Gesellschaften mit niedriger Störungshäufigkeit fehlt es an einem vollkommenen Umweltdeterminismus, und historische Ereignisse sind für das Verständnis der gegenwärtigen Zusammensetzung wichtig. Gesellschaften unterscheiden sich in dem Grad des Umweltdeterminismus, und Arten unterscheiden sich in der Nischenbreite und in dem Grad der Standortsspezifität. Es wird die Einbeziehung der Vegetationsdynamik in die Forstwirtschaftsführung erörtert.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

Literature Cited

  • Ahlgren, C. E. 1974. Effects of fires on temperate forests: North Central United States. Pages 195–223in T. T. Kozlowski and C. E. Ahlgren (eds.). Fire and ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Ahlgren, I. F. andC. E. Algren. 1960. Ecological effects of forest fires. Bot. Rev.46: 304–310.

    Google Scholar 

  • Aikman, J. M. andA. W. Smelser. 1938. The structure and environment of forest communities in central Iowa. Ecology19: 141–148.

    Article  Google Scholar 

  • Albertson, F. W. andG. W. Tomanek. 1965. Vegetation changes during a 30-year period in grassland communities near Hays, Kansas. Ecology46: 714–720.

    Article  Google Scholar 

  • ——, andA. Riegel. 1957. Ecology of drought cycles and grazing intensity on grasslands of Great Central Plains. Ecol. Monogr.27: 27–44.

    Article  Google Scholar 

  • —— andJ. E. Weaver. 1945. Injury and death or recovery of trees in a prairie climate. Ecol. Monogr.15: 395–433.

    Article  Google Scholar 

  • Anderson, D. J. 1967. Studies on the structure in plant communities. IV. Cyclical succession in Dryas communities from northwest Iceland. J. Ecol.55: 629–635.

    Article  Google Scholar 

  • —,R. C. Cooke, T. T. Elkington, andD. J. Read. 1966. Studies in structure in plant communities. II. The structure of some dwarf-heath and birch-copse communities in Sjald-Fannardalur, north-west Iceland. J. Ecol.54: 781–793.

    Article  Google Scholar 

  • Anderson, G. W. andR. L. Anderson. 1963. The rate of spread of oak wilt in the Lake States. J. For.631: 823–825.

    Google Scholar 

  • Anderson, R. C. andJ. S. Fralish. 1975. An investigation of palmetto,Paurotis wrightii (Griseb. and Wendl.) Britt., communities in Belize, Central America. Turrialba25: 37–44.

    Google Scholar 

  • Antonovics, J. 1971. The effects of a heterogeneous environment on the genetics of natural populations. Am. Sci.59: 593–599.

    PubMed  CAS  Google Scholar 

  • Armson, K. A. andR. J. Fessenden. 1973. Forest windthrows and their influence on soil morphology. Soil Sci. Soc. Am. Proc.37: 781–783.

    Article  Google Scholar 

  • Armstrong, R. A. 1976. Fugitive species: experiments with fungi and some theoretical considerations. Ecology57: 953–963.

    Article  Google Scholar 

  • Ashe, W. W. 1918. Note on “Ice Storms in the Southern Appalachians” by Verne Rhoades. Mon. Weather Rev.46: 374.

    Article  Google Scholar 

  • Ashton, P. S. 1969. Speciation among tropical forest trees: some deductions in the light of recent evidence. Biol. J. Linn. Soc.1: 155–196.

    Article  Google Scholar 

  • Aubréville, A. 1971. Regeneration patterns in the closed forest of the Ivory Coast. Pages 41–55in S. R. Eyre (ed.). World vegetation types. Macmillan Press, Ltd., London. 264pp.

    Google Scholar 

  • Auclair, A. N. 1975. Sprouting response inPrunus serotina Erhr.: multivariate analysis of site, forest structure, and growth rate relationships. Am. Midl. Nat.94: 72–87.

    Article  Google Scholar 

  • — andG. Cottam. 1971. Dynamics of black cherry (Prunus serotina Erhr.) in southern Wisconsin oak forests. Ecol. Monogr.41: 153–177.

    Article  Google Scholar 

  • Austin, M. P. 1977. Use of ordination and other multivariate descriptive methods to study succession. Vegetatio35: 165–176.

    Article  Google Scholar 

  • Bailey, A. W. andC. E. Poulton. 1968. Plant communities and environmental interrelationships in a portion of the Tillamook burn, northwestern Oregon. Ecology49: 1–13.

    Article  Google Scholar 

  • Baker, H. G. 1970. Evolution in the tropics. Biotropica2: 101–111.

    Article  Google Scholar 

  • Bannister, B. A. 1970. Ecological life cycle ofEnterpe globose. Gaertn. Pages 299–314in H. T. Odum (ed.). A tropical rainforest. U.S. A.E.C., Div. Tech. Inf. Washington, D. C.

    Google Scholar 

  • Barbour, M. A. andT. M. DeJong. 1977. Response of West Coast beach taxa to salt spray, seawater inundation, and soil salinity. Bull. Torrey Bot. Club104: 29–34.

    Article  Google Scholar 

  • Barclay-Estrup, P. 1970. The description and interpretation of cyclical processes in a heath community. II. Changes in biomass and shoot production during theCalluna cycle. J. Ecol.58: 243–249.

    Article  Google Scholar 

  • — andC. H. Gimingham. 1969. The description and interpretation of cyclical processes in a heath community. 1. Vegetational change in relation to theCalluna cycle. J. Ecol.57: 737–758.

    Article  Google Scholar 

  • Barden, L. S. Tree replacement in small canopy gaps of a southern Appalachian forest. (Unpublished manuscript.)

  • Batchelder, R. B. 1967. Spatial and temporal patterns of fire in the tropical world. Proc. Annu. Tall Timbers Fire Ecol. Conf.6: 171–707.

    Google Scholar 

  • Beard, J. S. 1945. Some ecological work in the Caribbean. Emp. For. J.24: 40–46.

    Google Scholar 

  • —. 1953. The savanna vegetation of northern tropical America. Ecol. Monogr.23: 149–215.

    Article  Google Scholar 

  • —. 1955. The classification of tropical American vegetation types. Ecology36: 89–100.

    Article  Google Scholar 

  • Beasleigh, W. J. andG. A. Yarranton. 1974. Ecological strategy and tactics ofEquisetum sylvaticum during a postfire succession. Can. J. Bot.52: 2299–2318.

    Article  Google Scholar 

  • Behre, C. E. 1921. A study of windfall in the Adirondacks. J. For.19: 632–637.

    Google Scholar 

  • Bell, D. T. 1974. Tree stratum composition and distribution in a streamside forest. Am. Midl. Nat.92: 35–46.

    Article  Google Scholar 

  • — andR. del Moral. 1977. Vegetation gradients in the stream-side forest of Hickory Creek, Will County, Illinois. Bull. Torrey Bot. Club104: 127–135.

    Article  Google Scholar 

  • Benninghoff, W. S. 1952. Interaction of vegetation and soil frost phenomena. Arctic5: 34–44.

    Google Scholar 

  • Bird, R. D. 1930. Biotic communities of the aspen parkland of central Canada. Ecology11: 356–442.

    Article  Google Scholar 

  • Bishop, J. W. 1977. Effects of rainstorms on phytoplankton production in Westhampton Lake, Virginia. Abstract in Bull. Ecol. Soc. Am.58(2): 52.

    Google Scholar 

  • Biswell, H. H. 1974. Effects of fire on chaparral. Pages 321–364in T. T. Kozlowski and C. E. Ahlgren (eds.). Fire and ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Blais, J. R. 1954. The recurrence of spruce budworm infestations in the past century in the Lac Seul area of northwestern Ontario. Ecology35: 62–71.

    Article  Google Scholar 

  • Blaisdell, R. S., J. Wooten, andR. K. Godfrey. 1973. The role of magnolia and beech in forest processes in the Tallahassee, Florida, Thomasville, Georgia area. Proc. Annu. Tall Timbers Fire Ecol. Conf.13: 363.

    Google Scholar 

  • Bleakney, J. S. 1972. Ecological implications of annual variation in tidal extremes. Ecology53: 933–938.

    Article  Google Scholar 

  • Blydenstein, J. 1968. Burning and tropical American savannas. Proc. Annu. Tall Timbers Fire Ecol. Conf.8: 1–15.

    Google Scholar 

  • Bogucki, D. J. 1970. Debris slides and related flood damage with the September 1, 1951, cloudburst in the Mt. LeConte-Sugarland Mountain area. Great Smoky Mountains National Park. Ph.D. Thesis, Univ. Tennessee, Knoxville. 165 pp.

    Google Scholar 

  • —. 1977. Debris slide hazards in the Adirondack Province of New York State. Environ. Geol.1: 317–328.

    Article  Google Scholar 

  • Borchert, J. R. 1950. The climate of the central North American grassland. Assoc. Am. Geogr. Ann.40: 1–39.

    Article  Google Scholar 

  • Bormann, F. H. andG. E. Likens. 1979. Pattern and process in a forested ecosystem. Springer-Verlag, New York. 253 pp.

    Google Scholar 

  • Bray, J. R. 1956. Gap phase replacement in a maple-basswood forest. Ecology37: 598–600.

    Article  Google Scholar 

  • Brayton, R. D. andG. M. Woodwell. 1966. Effects of ionizing radiation and fire onGaylussacia baccata andVaccinium vaccilans. Am. J. Bot.53: 816–820.

    Article  Google Scholar 

  • Brewer, R. andP. G. Merritt. 1978. Windthrow and tree replacement in a climax beechmaple forest. Oikos30: 149–152.

    Article  Google Scholar 

  • Brown, J. H. 1960. The role of fire in altering the species composition of forests in Rhode Island. Ecology41: 310–316.

    Article  Google Scholar 

  • Brown, W. L. 1961. Mass insect control programs: Four case histories. Psyche68: 75–111.

    Google Scholar 

  • Budowski, G. 1963. Forest succession in tropical lowlands. Turrialba13: 42–44.

    Google Scholar 

  • —. 1966. Fire in tropical American lowlands. Proc. Annu. Tall Timbers Fire Ecol. Conf.5: 5–22.

    Google Scholar 

  • —. 1970. The distinction between old secondary and climax species in tropical Central American lowland forests. Trop. Ecol.11: 44–48.

    Google Scholar 

  • Buell, M. F. 1956. Spruce-fir, maple-basswood competition in Itasca Park, Minnesota. Ecology37: 606.

    Article  Google Scholar 

  • — andH. F. Buell. 1941. Surface level fluctuations in Cedar Creek Bog, Minnesota. Ecology22: 317–321.

    Article  Google Scholar 

  • ——. 1975. Moat bogs in the Itasca Park area, Minnesota. Bull. Torrey Bot. Club102: 6–9.

    Article  Google Scholar 

  • ——, andW. A. Reiners. 1968. Radial mat growth in Cedar Creek Bog, Minnesota. Ecology49: 1198–1199.

    Article  Google Scholar 

  • ——, andJ. A. Small. 1954. Fire in the history of Mettler’s Woods. Bull. Torrey Bot. Club81: 253–255.

    Article  Google Scholar 

  • ———, andC. D. Monk. 1961. Drought effect on radial growth of trees in the William L. Hutcheson Memorial Forest. Bull. Torrey Bot. Club88: 176–180.

    Article  Google Scholar 

  • — andW. E. Gordon. 1945. Hardwood-conifer contact zone in Itasca Park, Minnesota. Am. Midl. Nat.34: 433–439.

    Article  Google Scholar 

  • —,A. N. Langford, D. W. Davidson, andL. F. Ohmann. 1966. The upland forest continuum in northern New Jersey. Ecology47: 416–432.

    Article  Google Scholar 

  • — andW. E. Martin. 1961. Competition between maple-basswood and fir-spruce communities in Itasca Park, Minnesota. Ecology42: 428–429.

    Article  Google Scholar 

  • — andW. S. Wistendahl. 1955. Flood plain forests of the Raritan River. Bull. Torrey Bot. Club82: 463–472.

    Article  Google Scholar 

  • Carvel, K. L., E. H. Tryon, andR. P. True. 1957. Effects of glaze on the development of Appalachian hardwoods. J. For.55: 130–132.

    Google Scholar 

  • Chabreck, R. H. andA. W. Palmisano. 1973. The effects of Hurricane Camille on the marshes of the Mississippi River delta. Ecology54: 1118–1123.

    Article  Google Scholar 

  • Chapman, H. H. 1932. Is the longleaf type a climax? Ecology13: 328–334.

    Article  Google Scholar 

  • —. 1947. Natural areas. Ecology28: 193–194.

    Article  Google Scholar 

  • Chesapeake Research Consortium, Inc. 1977. The effects of tropical storm Agnes on the Chesapeake Bay estuarine system. John Hopkins Univ. Press, Baltimore. 639 pp.

    Google Scholar 

  • Christensen, N. L. andC. H. Muller. 1975. Effects of fire on factors controlling plant growth inAdenostoma Chaparral. Ecol. Monogr.45: 29–55.

    Article  Google Scholar 

  • Churchill, E. D. andH. C. Hanson. 1958. The concept of climax in arctic and alpine vegetation. Bot. Rev.24: 127–191.

    Google Scholar 

  • Clark, F. B. 1962. White ash, hackberry, and yellow poplar seed remain viable in the forest litter. Indiana Acad. Sci. Proc.72: 112–114.

    Google Scholar 

  • Cline, A. C. and S. H. Spurr. 1942. The virgin upland forest of central New England: A study of old growth stands in the Pisgah Mountain section of southwestern New Hampshire. Harv. For. Bull.21. 58 pp.

  • Cochrane, G. R. and J. S. Rowe. 1969. Fire in the tundra at Rankin Inlet NWT. Proc. Annu. Tall Timbers Fire Ecol. Conf. 292 pp.

  • Connell, J. H. 1978. Diversity in tropical rain forests and coral reefs. Science199: 1302–1310.

    Article  PubMed  CAS  Google Scholar 

  • — andR. O. Slayter. 1977. Mechanisms of succession in natural communities and their role in community stability and organization. Am. Nat.111: 1119–1144.

    Article  Google Scholar 

  • Coombe, D. E. andF. White. 1951. Notes on calcicolous communities and peat formation in Norwegian Lapland. J. Ecol.39: 33–62.

    Article  Google Scholar 

  • Cooper, C. F. 1960. Changes in vegetation, structure, and growth of southwestern pine forests since white settlement. Ecol. Monogr.30: 129–164.

    Article  Google Scholar 

  • Cooper, W. S. 1913. The climax forest of Isle Royale. Lake Superior, and its development. Bot. Gaz.55: 144; 115–140; 189–235.

    Google Scholar 

  • —. 1926. The fundamentals of vegetational change. Ecology7: 391–413.

    Article  Google Scholar 

  • Coupland, R. T. 1974. Fluctuations in North American grassland vegetation. Pages 235–241in R. Knapp (ed.). Vegetation dynamics. Dr. W. Junk, The Hague.

    Google Scholar 

  • Cowles, H. C. 1899. The ecological relations of the vegetation on the sand dunes of Lake Michigan. Part I. Geographical relations of the dune floras. Bot. Gaz.27: 95–117; 167–201; 281–308; 361–391.

    Article  Google Scholar 

  • Craighead, F. C., Sr. 1971. The trees of South Florida. I. The natural environments and succession. University of Miami Press, Coral Gables, Florida. 212 pp.

  • Curtis, J. T. 1959. The vegetation of Wisconsin. University of Wisconsin Press, Madison. 657 pp.

  • Cwynar, L. C. 1978. Recent history of fire and vegetation from laminated sediment of Greenleaf Lake, Algonquin Park, Ontario. Can. J. Bot.56: 10–21.

    Google Scholar 

  • Dansereau, P. 1946. L’érabliere laurentienne. II. Les successions et leurs indicateurs. Can. J. Res. Sect. C, Bot. Sci.24: 235–291.

    Google Scholar 

  • —. 1957. Biogeography, an ecological perspective. Ronald Press Co., New York. 394 pp.

    Google Scholar 

  • —. 1974. Classification of successions and of their terminal stages. Pages 125–135in R. Knapp (ed.). Vegetation dynamics. Dr. W. Junk, The Hague.

    Google Scholar 

  • Daubenmire, R. F. 1968a. Plant communities: A textbook of plant synecology. Harper and Row, New York. 30 pp.

    Google Scholar 

  • —. 1968b. Ecology of fire in grasslands. Adv. Ecol. Res.5: 209–266.

    Google Scholar 

  • Davis, M. B. 1976. Pleistocene biogeography of temperate deciduous forests. Geosci. Man13: 13–26.

    Google Scholar 

  • Dayton, P. K. 1971. Competition, disturbance, and community organization: the provision and subsequent utilization of space in a rocky intertidal community. Ecol. Monogr.41: 351–389.

    Article  Google Scholar 

  • DeGraaf, R. M. 1978. New life from dead trees. Nat. Wildl. (June 1978): 29–31.

  • Delcourt, H. R. andP. A. Delcourt. 1977. Presettlement magnolia-beech climax of the Gulf Coastal Plain: Quantitative evidence from the Apalachicola River Bluffs, north-central Florida. Ecology58: 1085–1093.

    Article  Google Scholar 

  • Demaree, D. 1932. Submerging experiments with Taxidium. Ecology13: 258–262.

    Article  Google Scholar 

  • Denny, C. S. and J. C. Goodlett. 1956. Microrelief resulting from fallen trees.In C. S. Denny (ed.). Surificial geology and geomorphology of Potter County, Pennsylvania. U.S. Geol. Surv. Prof. Pap. 288.

  • Dittus, W. P. J. 1977. The ecology of a semi-evergreen forest community in Sri Lanka. Biotropica9: 268–286.

    Article  Google Scholar 

  • Dix, R. L. 1957. Sugar maple in the climax forests of Washington, DC. Ecology38: 663–665.

    Article  Google Scholar 

  • — andJ. M. A. Swan. 1971. The roles of disturbance and succession in upland forest at Candle Lake, Saskatchewan. Can. J. Bot.49: 657–676.

    Google Scholar 

  • Dolan, R., B. P. Hayden, andG. Soucie. 1978. Environmental dynamics and resource management in the U.S. National Parks. Environ. Manage.2: 249–258.

    Article  Google Scholar 

  • Douglas, G. W. 1974. Montane zone vegetation of the Alsek River region, Southwestern Yukon. Can. J. Bot.52: 2505–2532.

    Article  Google Scholar 

  • — andT. M. Ballard. 1971. Effects of fire on alpine plant communities in the North Cascades, Washington. Ecology52: 1058–1064.

    Article  Google Scholar 

  • Downs, A. A. 1938. Glaze damage in the birch-beech-maple-hemlock type of Pennsylvania and New York. J. For.36: 63–70.

    Google Scholar 

  • Drury, W. H. andI. C. T. Nisbet. 1971. Inter-relations between developmental models in geomorphology, plant ecology, and animal ecology. Gen. Syst.16: 57–68.

    Google Scholar 

  • ——. 1973. Succession. J. Arnold Arbor. Harv. Univ.54: 331–368.

    Google Scholar 

  • Eggler, W. A. 1971. Quantitative studies of vegetation on sixteen young lava flows on the Island of Hawaii. Trop. Ecol.12: 66–100.

    Google Scholar 

  • Egler, F. E. 1947. Arid Southeast Oahu vegetation, Hawaii. Ecol. Monogr.17: 383–435.

    Article  Google Scholar 

  • Ehrlich, P. R., D. E. Breedlove, P. F. Brussard, andM. A. Sharp. 1972. Weather and “ regulation” of subalpine populations. Ecology53: 243–247.

    Article  Google Scholar 

  • Fahey, J. J. 1975. Fire in the forests of Maine and New Hampshire. Abstract in Bull. Ecol. Soc. Am.56: 41.

    Google Scholar 

  • Flaccus, E. 1959. Revegetation of landslides in the White Mountains of New Hampshire. Ecology40: 692–703.

    Article  Google Scholar 

  • Foldats, E. andE. Rutkis. 1975. Ecological studies of chapparo (Curatella americana L.) and manteco (Byrsonima crassifolia HBK) in Venezuela. J. Biogeogr.2: 159–178.

    Article  Google Scholar 

  • Forcier, L. K. 1975. Reproductive strategies and the co-occurrence of climax tree species. Science189: 808–810.

    Article  PubMed  CAS  Google Scholar 

  • Foster, R. B. 1977.Tachigalia versicolor is a suicidal neotropical tree. Nature268: 624–626.

    Article  Google Scholar 

  • Fox, J. F. 1977. Alternation and coexistence of tree species. Am. Nat.111: 69–89.

    Article  Google Scholar 

  • Franklin, J. F. andC. T. Dyrness. 1969. Vegetation of Oregon and Washington. U.S.D.A. For. Serv. Res. Pap. PEW80: 1–216.

    Google Scholar 

  • Franz, E. H. andF. A. Bazzaz. 1977. Simulation of vegetation response to modified hydrologic regimes: A probabilistic model based on niche differentiation in a flood plain forest. Ecology58: 176–183.

    Article  Google Scholar 

  • Gams, H. 1918. Principienfragen der Vegetations forschung Viertelsjahrsschr. Naturf. Ges. Zürich67: 132–156.

    Google Scholar 

  • Gardiner, L. M. 1975. Insect attack and value loss in wind-damaged spruce and jack pine stands in northern Ontario. Can. J. For. Res.5: 387–398.

    Google Scholar 

  • Garren, K. H. 1943. Effects of fire on vegetation of the southeastern United States. Bot. Rev.9: 617–654.

    Google Scholar 

  • Gentry, A. H. 1974. Flowering phenology and diversity in tropical Bignoniaceae. Biotropica6: 64–68.

    Article  Google Scholar 

  • Ghent, A. W., D. A. Fraser, andJ. B. Thomas. 1957. Studies of regeneration of forest stands devastated by the spruce budworm. I. For. Sci.3: 184–208.

    Google Scholar 

  • Giesel, J. T. 1976. Reproductive strategies as adaptations to life in temporally heterogeneous environments. Ann. Rev. Ecol. Syst.7: 57–80.

    Article  Google Scholar 

  • Gill, A. M. 1977. Plant traits adapted to fires in Mediterranean land ecosystems. Pages 17–26in H. Mooney and C. E. Conrad (eds.). Proceedings of the symposium on the environmental consequences of fire and fuel management in Mediterranean ecosystems. U.S.D.A. For. Serv. Gen. Tech. Rep. WO-3. Washington, D.C.

  • Gill, D. E. 1975. Spatial patterning of pines and oaks in the New Jersey pine barrens. J. Ecol.63: 291–298.

    Article  Google Scholar 

  • Gleason, H. A. 1926. The individualistic concept of the plant association. Bull. Torrey Bot. Club53: 7–26.

    Article  Google Scholar 

  • —. 1927. Further views on the succession concept. Ecology8: 299–327.

    Article  Google Scholar 

  • Gómez-Pompa, A. 1971. Posible papel de la vegetación secundaria en la evolución de la flora tropical. Biotropica3: 125–135.

    Article  Google Scholar 

  • -,S. del Amo R., C. Vazgnez-Yanes, and A. Butando (eds.). 1976. Investigaciones sabre la regeneración de selvas altas en Veracruz, México. Cia. Editorial Continental, SA. Calz. de Tlalpan Núrn, 4620, México 22, D.F. 676 pp.

  • Goodlett, J. C. 1954. Vegetation adjacent to the border of Wisconsin drift in Potter County, Pennsylvania. Harv. For. Bull.25. 93 pp.

  • —. 1969. Vegetation and the equilibrium concept of landscape. Pages 33–44in K. N. H. Greenidge (ed.). Essays in plant geography and ecology. Nova Scotia Museum, Halifax.

    Google Scholar 

  • Grigal, D. F. andL. F. Ohmann. 1975. Classification, description, and dynamics of upland plant communities within a Minnesota wilderness area. Ecol. Monogr.45: 389–407.

    Article  Google Scholar 

  • Griggs, R. F. 1936. The vegetation of the Katmai district. Ecology17: 380–417.

    Article  Google Scholar 

  • Grime, J. P. 1977. Evidence for the existance of three primary strategies in plants and its relevance to ecological and evolutionary theory. Am. Nat.111: 1169–1194.

    Article  Google Scholar 

  • — andR. Hunt. 1975. Relative growth-rate: its range and adaptive signicance in a local flora. J. Ecol.63: 393–422.

    Article  Google Scholar 

  • Grubb, P. J. 1977. The maintenance of species-richness in plant communities: The importance of the regeneration niche. Biol. Rev.52: 107–145.

    Article  Google Scholar 

  • Habeck, J. R. andR. W. Mutch. 1973. Fire-dependent forests in the northern Rocky Mountains. Quat. Res.3: 408–424.

    Article  Google Scholar 

  • Hack, J. T. 1960. Interpretation of erosional topography in humid temperate regions. Am. J. Sci.258A: 80–97.

    Google Scholar 

  • Komarek, E. V. 1974. Effects of fire on temperate forests and related ecosystems: South.

  • -.and J. C. Goodlett. 1960. Geomorphology and forest ecology of a mountain region in the central Appalachians. U.S. Geol. Surv. Prof. Pap. 347.

  • Hall, T. F. andG. E. Smith. 1955. Effects of flooding on woody plants, West Sandy Dewatering Project, Kentucky Reservoir. J. For.53: 281–285.

    Google Scholar 

  • Hanes, T. L. 1971. Succession after fire in the chaparral of southern California. Ecol. Monogr.41: 27–52.

    Article  Google Scholar 

  • Hanson, H. C. 1950. Vegetation and soil profiles in some solifluction and mound areas in Alaska. Ecology31: 606–630.

    Article  Google Scholar 

  • — andW. Whitman. 1937. Plant succession on solonetz soils in western North Dakota. Ecology18: 516–522.

    Article  Google Scholar 

  • Harper, J. L. 1967. A Darwinian approach to plant ecology. J. Ecol.55: 247–270.

    Article  Google Scholar 

  • —. 1977. Population biology of plants. Academic Press, New York. 892 pp.

    Google Scholar 

  • —. andJ. Ogden. 1970. The reproductive strategy of higher plants: I. the concept of strategy with specific reference toSenecio vulgaris L. J. Ecol.58: 681–689.

    Article  Google Scholar 

  • — andJ. White. 1974. The demography of plants. Ann. Rev. Ecol. Syst.5: 419–463.

    Article  Google Scholar 

  • Harris, S. W. andW. H. Marshall. 1963. Ecology of water-level manipulations on a northern marsh. Ecology44: 331–343.

    Article  Google Scholar 

  • Hartshorn, G. 1977. Neotropical forest dynamics. Abstract in Bull. Ecol. Soc. Am.58: 29.

    Google Scholar 

  • Hayes, G. I. 1942. Difference in fire danger with altitude, aspect, and time of day. J. For.40: 318–323.

    Google Scholar 

  • Hedrick, P. W., M. E. Ginevan, andE. P. Ewing. 1976. Genetic polymorphism in heterogeneous environments. Ann. Rev. Ecol. Syst.7: 1–32.

    Article  Google Scholar 

  • — 1970. Landscape evolution, peatland types, and the environment in the Lake Agassiz Peatlands. Ecol. Monogr.40: 235–261.

    Article  Google Scholar 

  • -. 1971. The natural role of fire in northern conifer forests. Pages 61–72in C. W. Slaughter, R. J. Barney, and G. M. Hansen (ed.). Fire in the northern environmenta symposium. Pac. Northwest For. Range Exp. Stn., Portland, Oregon.

  • —. 1973. Fire in the virgin forests of the Boundary Waters Canoe Area, Minnesota. Quat. Res.3: 329–382.

    Article  Google Scholar 

  • -Heinselman, M. L. and H. E. Wright, Jr. (eds.). 1973. The ecological role of fire in natural conifer forests of western and northern North America. Quat. Res.3: 317–318.

  • Henry, J. D. andJ. M. A. Swan. 1974. Reconstructing forest history from live and dead plant material—An approach to the study of forest succession in southwest New Hampshire. Ecology55: 772–783.

    Article  Google Scholar 

  • Hett, J. M. andO. L. Loucks. 1976. Age structure models of balsam fir and eastern hemlock. J. Ecol.64: 1029–1044.

    Article  Google Scholar 

  • Hewetson, C. E. 1956. A discussion on the “climax” concept in relation to the tropical rain and deciduous forest. Emp. For. Rev.35: 274–291.

    Google Scholar 

  • Heyward, F. 1939. Relation of fire to stand composition of longleaf pine forests. Ecology20: 287–304.

    Article  Google Scholar 

  • Hibbs, D. H., B. C. Fisher, andB. F. Wilson. 1977. The survival strategy of striped maple. Abstract in Bull. Ecol. Soc. Am.58: 47.

    Google Scholar 

  • Horn, H. S. 1976. Succession. Pages 187–204in R. M. May (ed.). Theoretical ecology: Principles and applications. Blackwell, London.

    Google Scholar 

  • Horn, H. 1971. The adaptive geometry of trees. Princeton University Press, Princeton, New Jersey. 144 pp.

    Google Scholar 

  • Hosner, J. F. 1957. Effects of water on seed germination of bottomland trees. For. Sci.3: 67–69.

    Google Scholar 

  • —. 1960. Relative tolerance to complete inundation of fourteen bottomland tree species. For. Sci.6: 246–251.

    Google Scholar 

  • — andS. G. Boyce. 1962. Relative tolerance to water saturated soil of various bottomland hardwoods. For. Sci.8: 180–186.

    Google Scholar 

  • Hough, A. F. andR. D. Forbes. 1943. The ecology and silvics of forests in the high plateaus of Pennsylvania. Ecol. Monogr.13: 299–320.

    Article  Google Scholar 

  • Hough, A. F. andR. F. Taylor. 1946. Response of Allegheny northern hardwoods to partial cutting. J. For.44: 30–38.

    Google Scholar 

  • Houston, D. B. 1973. Wildfires in northern Yellowstone National Park. Ecology54: 1111–1117.

    Article  Google Scholar 

  • Humphrey, R. R. 1974. Fire in the deserts and desert grassland of North America. Pages 365–400in T. T. Kozlowski and C. E. Ahlgren (eds.). Fire and ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Hutnik, R. J. 1952. Reproduction on windfalls in a northern hardwood stand. J. For.50: 693–694.

    Google Scholar 

  • Illick, J. S. 1916. A destructive snow and ice storm. For. Leaves15: 103–107.

    Google Scholar 

  • Ives, R. L. 1942. The beaver-meadow complex. J. Geomorphol.5: 191–203.

    Google Scholar 

  • Janzen, D. H. 1967. Why mountain passes are higher in the tropics. Am. Nat.101: 233–249.

    Article  Google Scholar 

  • —. 1976. Why bamboos wait so long to flower. Ann. Rev. Ecol. Syst.7: 347–392.

    Article  Google Scholar 

  • Johnson, A. W., L. A. Viereck, R. E. Johnson, and H. Melchior. 1966. Vegetation and flora.In N. J. Wilimorsky and J. N. Wolfe (eds.). Environment of the Cape Thompson region, Alaska. U.S. A.E.C.

  • Johnson, E. A. 1975. Buried seed populations in the subarctic forest east of Great Slave Lake, Northwest Territories. Can. J. Bot.53: 2933–2941.

    Google Scholar 

  • — andJ. S. Rowe. 1975. Fire in the subarctic wintering ground of the Beverley caribou herd. Am. Midl. Nat.94: 1–14.

    Article  Google Scholar 

  • Johnson, P. L. andW. D. Billings. 1962. The alpine vegetation of the Beartooth Plateau in relation to Cropedogenic processes and patterns. Ecol. Monogr.32: 102–135.

    Article  Google Scholar 

  • Johnson, W. C., R. L. Burgess, andW. R. Keammerer. 1976. Forest overstory vegetation and environment on the Missouri River Floodplain in North Dakota. Ecol. Monogr.46: 59–84.

    Article  Google Scholar 

  • Jones, E. W. 1945. The structure and reproduction of the virgin forest of the North Temperate zone. New Phytol.44: 130–148.

    Article  Google Scholar 

  • —. 1955. Ecological studies on the rain forest of Southern Nigeria. I. J. Ecol.43: 564–594.

    Google Scholar 

  • —. 1956. Ecological studies on the rain forest of Southern Nigeria. II. J. Ecol.44: 83–117.

    Article  Google Scholar 

  • Keatinge, T. H. 1975. Plant community dynamics in wet heathland. J. Ecol.63: 163–172.

    Article  Google Scholar 

  • Keeley, J. E. 1977. Seed production, seed populations in soil, and seedling production after fire for two congeneric pairs of sprouting and nonsprouting chaparral shrubs. Ecology58: 820–829.

    Article  Google Scholar 

  • — andP. H. Zedler. 1978. Reproduction of chaparral shrubs after fire: a comparison of sprouting and seeding strategies. Am. Midl. Nat.99: 142–161.

    Article  Google Scholar 

  • Kellman, M. 1974. Preliminary seed budgets for two plant communities in coastal British Columbia. J. Biogeogr.1: 123–133.

    Article  Google Scholar 

  • Kershaw, K. A. 1960. Cyclic and pattern phenomena as exhibited byAlchemilla alpina. J. Ecol.48: 443–453.

    Article  Google Scholar 

  • Kessell, S. R. 1976. Gradient modeling: a new approach to fire modeling and wilderness resource management. Environ. Manage.1: 39–48.

    Article  Google Scholar 

  • Kilgore, B. M. 1973. The ecological role of fire in Sierran conifer forests. Quat. Res.3: 496–513.

    Article  Google Scholar 

  • Kimmerer, R. W. andT. F. H. Allen. 1978. The role of disturbance in the structure of a riparian bryophyte community. Abstract in Bull. Ecol. Soc. Ann.59: 88.

    Google Scholar 

  • Knapp, R. (ed.). 1974. Vegetation dynamics. Dr. W. Junk, The Hague. 356 pp.

    Google Scholar 

  • —. 1974. Cyclic successions and ecosystem approaches in vegetation dynamics. Pages 92–100in R. Knapp (ed.). Vegetation dynamics. Dr. W. Junk, The Hague.

    Google Scholar 

  • Knight, D. H. 1975. A phytosociological analysis of species-rich tropical forest on Barro Colorado Island, Panama. Ecol. Monogr.45: 259–284.

    Article  Google Scholar 

  • Koevenig, J. L. 1976. Effect of climate, soil physiography, and seed germination on the distribution of river birch (Betulu nigra). Rhodora78: 420–437.

    Google Scholar 

  • Komarek, E. V., Sr. 1966. The meteorological basis for fire ecology. Pages 85–126 in Proc. Annu. Tall Timbers Fire Ecol. Conf.

  • —. 1968. Lightning and lightning fires as ecological forces. Proc. Annu. Tall Timbers Fire Ecol. Conf.8: 169–197.

    Google Scholar 

  • Komarek, E. V. 1974. Effects of fire on temperate forests and related ecosystems: South-eastern United States. Pages 251–277in T. T. Kozlowski and C. E. Ahlgren (eds.). Fire and Ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Korchagin, A. A. andV. G. Karpov. 1974. Fluctuations in coniferous taiga communities. Pages 227–231in R. Knapp (ed.). Vegetation dynamics. Dr. W. Junk, The Hague.

    Google Scholar 

  • Kozlowski, T. T. andC. E. Ahlgren (eds.). 1974. Fire and ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Krefting, L. W. andC. E. Ahlgren. 1974. Small mammals and vegetation changes after fire in a mixed conifer-hardwood forest. Ecology55: 1391–1398.

    Article  Google Scholar 

  • Kucera, C. L. andS. C. Martin. 1957. Vegetation and soil relationships in the glade region of the Southwestern Missouri Ozarks. Ecology38: 285–291.

    Article  Google Scholar 

  • Küchler, A. W. 1964. Potential natural vegetation of the conterminous United States. Am. Geogr. Soc. Spec. Publ. 36. 116 pp.

  • Langenheim, J. H. 1956. Plant succession on a subalpine earthflow in Colorado. Ecology37: 301–317.

    Article  Google Scholar 

  • Langford, A. N. andM. F. Buell. 1969. Integration, identity, and stability in the plant association. Adv. Ecol. Res.6: 83–135.

    Article  Google Scholar 

  • Larson, F. 1940. The role of the bison in maintaining the short grass plain. Ecology21: 113–121.

    Article  Google Scholar 

  • Leak, W. B. 1963. Delayed germination of white ash seeds under forest conditions. J. For.61: 768–772.

    Google Scholar 

  • Lejeune, R. R. 1955. Population ecology of the larch sawfly. Can. Entomol.87: 111–117.

    Article  Google Scholar 

  • Lemon, P. C. 1949. Successional responses of herbs in the longleaf-slash pine forest after fire. Ecology30: 135–145.

    Article  Google Scholar 

  • —. 1961. Forest ecology of ice storms. Bull. Torrey Bot. Club88: 21–29.

    Article  Google Scholar 

  • —. 1968. Effects of fire on African Plateau grassland. Ecology49: 316–322.

    Article  Google Scholar 

  • Levin, S. A. 1976. Population dynamic models in heterogeneous environments. Ann. Rev. Ecol. Syst.7: 287–310.

    Article  Google Scholar 

  • -.and R. J. Paine. 1975. The role of disturbance in models of community structure. Pages 56–67in S. A. Levin (ed.). Ecosystem analysis and prediction. Philadelphia, Pennsylvania.

  • Lewis, F. J. andE. S. Dowding. 1926. The vegetation and retrogressive changes of peat areas (“ muskegs”) in central Alberta. J. Ecol.14: 317–341.

    Article  Google Scholar 

  • Lewis, K. P. 1975. Community analysis and the dynamics of establishmentof Acer saccharinum L. (silver maple) on flood plains of the unglaciated Appalachian Plateau. Abstract in Bull. Ecol. Soc. Am.56: 489.

    Google Scholar 

  • Lichens, G. E. andM. B. Davis. 1975. Post-glacial history of Mirror Lake and its watershed in New Hampshire, USA: An initial report. Verh. Int. Ver. Limnol.19: 982–993.

    Google Scholar 

  • Lindsey, A. A., R. O. Petty, D. K. Sterling, andW. Van Asdall. 1961. Vegetation and environment along the Wabash and Tippecanoe Rivers. Ecol. Monogr.31: 105–156.

    Article  Google Scholar 

  • Little, S. 1946. The effects of forest fires on the stand history of New Jersey’s pine region. Northeast. For. Exp. Stn., For. Manage. Pap.2. 45 pp.

  • — 1974. Effects of fire on temperate forests: Northeastern United States. Pages 225–250in T. T. Kozlowski and C. E. Ahlgren (eds.). Fire and ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Lloyd, L. L. andG. E. Gruell. 1973. The ecological role of fire in the Jackson Hole Area, northwestern Wyoming. Quat. Res.3: 425–443.

    Article  Google Scholar 

  • Lorimer, C. G. 1977a. The presettlement forest and natural disturbance cycle of northeastern Maine. Ecology58: 139–148.

    Article  Google Scholar 

  • —. 1977b. Stand history and dynamics of a southern Appalachian virgin forest. Ph.D. Thesis, Duke Univ. 200 pp. Univ. Microfilms No. 77-18. 778. Ann Arbor, Michigan.

    Google Scholar 

  • Loucks, O. L. 1970. Evolution of diversity, efficiency, and community stability. Am. Zool.10: 17–25.

    PubMed  CAS  Google Scholar 

  • Loya, Y. 1976. Recolonization of red sea corals affected by natural catastrophes and manmade perturbations. Ecology57: 278–289.

    Article  Google Scholar 

  • Lugo, A. 1977. Mangroves: succession or steady-state? Abstract in Bull. Ecol. Soc. Am.58: 29.

    Google Scholar 

  • Lunan, J. S. andJ. R. Habeck. 1973. The effects of fire exclusion on ponderosa pine communities in Glacier National Park, Montana. Can. J. For. Res.3: 574.

    Article  Google Scholar 

  • Lutz, H. J. 1940. Disturbance of soil resulting from uprooting of trees. Yale Univ. Sch. For. Bull.45. 37 pp.

  • — andA. L. McComb. 1935. Origin of white pine in virgin forest stands in north-western Pennsylvania as indicated by stem and basal branch features. Ecology16: 252–256.

    Article  Google Scholar 

  • Lyford, W. H. and D. W. MacLean. 1966. Mound and pit microrelief in relation to soil disturbance and tree distribution in New Brunswick, Canada. Harv. For. Pap.15.

  • Macaloney, H. J. 1966. The impact of insects in the northern hardwoods type. U.S.D.A. For. Serv. Res. Note NC-10. 3 pp.

  • Madison, M. 1977. A revision of Monstera (Araceae). Contrib. Gray Herb. Harv. Univ.207. 100 pp.

  • Maissurow, D. K. 1935. Fire as a necessary factor in the perpetuation of white pine. J. For.33: 373–378.

    Google Scholar 

  • —. 1941. The role of fire in the perpetuation of virgin forests in northern Wisconsin. J. For.39: 201–207.

    Google Scholar 

  • Major, J. 1974. Kinds and rates of changes in vegetation and chronofunctions. Pages 8–18in R. Knapp (ed.). Vegetation dynamics. Dr. W. Junk, The Hague.

    Google Scholar 

  • Marks, P. L. 1974. The role of pin cherry (Prunus pensylvanica L.) in the maintenance of stability in northern hardwood ecosystems. Ecol. Monogr.44: 73–88.

    Article  Google Scholar 

  • —. 1975. On the relation between extension growth and successional status of deciduous trees of the northeastern United States. Bull. Torrey Bot. Club102: 172–177.

    Article  Google Scholar 

  • Marquis, D. A. 1975. Seed storage and germination under northern hardwood forest. Can. J. For. Res.5: 478–484.

    Google Scholar 

  • Martin, W. E. 1959. The vegetation of Island Beach State Park, New Jersey. Ecol. Monogr.29: 1–46.

    Article  Google Scholar 

  • McCleod, K. W. andJ. K. McPherson. 1973. Factors limiting the distribution ofSalix nigra. Bull. Torrey Bot. Club100: 102–110.

    Article  Google Scholar 

  • McDermott, R. E. 1954. Effects of saturated soil in seedling growth of some bottomland hardwood species. Ecology35: 36–41.

    Article  Google Scholar 

  • McGinty, D. T. andE. J. Christy. 1977. Turkey oak ecology on a Georgia sand hill. Am. Midl. Nat.98: 487–491.

    Article  Google Scholar 

  • McIntosh, R. P. 1961. Windfall in forest ecology. Ecology42: 834.

    Article  Google Scholar 

  • Monk, C. D. 1961a. The vegetation of the William L. Hutcheson Memorial Forest, New Jersey. Bull. Torrey Bot. Club88: 156–166.

    Article  Google Scholar 

  • —. 1961b. Past and present influences on reproduction in the William L. Hutcheson Memorial Forest, New Jersey. Bull. Torrey Bot. Club88: 167–175.

    Article  Google Scholar 

  • Mooney, H., J. M. Bonnicksen, N. L. Christensen, J. E. Lotan, and W. A. Reiners (eds.). 1979. Fire regimes and ecosystem properties. U.S.D.A. For. Serv. Gen. Tech. Rept. Washington, D.C. (In press.)

  • -and C. E. Conrad (eds.). 1977. Proceedings of the symposium on the environmental consequences of fire and fuel management in Mediterranean ecosystems. U.S.D.A. For. Serv. Gen. Tech. Rept. WO-3. Washington, D.C. 498 pp.

  • Morris, R. F. (ed.). 1963. The dynamics of epidemic spruce budworm dynamics. Entomol. Soc. Can. Mem.31. 352 pp.

    Google Scholar 

  • Mosquin, T. 1971. Competition for pollinators as a stimulus for the evolution of flowering time. Oikos22: 398–402.

    Article  Google Scholar 

  • Mount, A. B. 1969. Eucalypt ecology as related to fire. Proc. Annu. Tall Timbers Fire Ecol. Conf.9: 75–108.

    Google Scholar 

  • Mueller-Dombois, D. andH. Ellenberg. 1974. Aims and methods of vegetation ecology. John Wiley and Sons, New York. 547 pp.

    Google Scholar 

  • Muller, C. H. 1952. Plant succession in arctic heath and tundra in northern Scandinavia. Bull. Torrey Bot. Club79: 296–309.

    Article  Google Scholar 

  • Munger, T. T. 1940. The cycle from douglas fir to hemlock. Ecology21: 451–459.

    Article  Google Scholar 

  • Munro, N. 1966. The fire ecology of Caribbean pine in Nicaragua. Proc. Annu. Tall Timbers Fire Ecol. Conf.5: 67–84.

    Google Scholar 

  • Mutch, R. W. 1970. Wildland fires and ecosystems—a hypothesis. Ecology51: 1046–1051.

    Article  Google Scholar 

  • Nanson, G. C. andH. F. Beach. 1977. Forest succession and sedimentation on a meandering river flood plain, northeast British Columbia, Canada. J. Biogeogr.4: 229–252.

    Article  Google Scholar 

  • Newton, M., B. A. elHassan, andJ. Zavitkovski. 1968. Role of red alder in western Oregon forest succession. Pages 73–84in J. M. Trappe, J. F. Franklin, A. F. Tarrant, and G. M. Hansen (eds.). Biology of alder. Pac. Northwest. For. Range Exp. Stn., U.S.D.A. For. Serv. Portland, Oregon.

    Google Scholar 

  • Nichols, G. E. 1920a. The plant associations of eroding areas along the seacoast. Bull. Torrey Bot. Club47: 89–117.

    Article  Google Scholar 

  • —. 1920b. The associations of depositing areas along the seacoast. Bull. Torrey Bot. Club47: 511–548.

    Article  Google Scholar 

  • Nixon, E. S., R. L. Willett, andP. W. Cox. 1977. Woody vegetation of a virgin forest in an Eastern Texas river bottom. Castanea42: 227–237.

    Google Scholar 

  • Noble, R. E. andP. K. Murphy. 1975. Short term effects of prolonged backwater flooding on understory vegetation. Castanea40: 228–238.

    Google Scholar 

  • Noy-Meir, I. 1973. Desert ecosystems: Environment and producers. Ann. Rev. Ecol. Syst.4: 25–51.

    Article  Google Scholar 

  • O’Cinneide, M. S. 1975. Aspect and wind direction as factors in forest stability: The case of Northern Ireland. J. Biogeogr.2: 127–139.

    Article  Google Scholar 

  • Odum, E. P. 1971. Fundamentals of Ecology, 3rd ed. Saunders, Philadelphia. 574 pp.

    Google Scholar 

  • Old, S. M. 1969. Microclimate, fire, and plant production in an Illinois prairie. Ecol. Monogr.39: 355–383.

    Article  Google Scholar 

  • Oliver, C. D. andE. P. Stephens. 1977. Reconstruction of a mixed species forest in Central New England. Ecology58: 562–572.

    Article  Google Scholar 

  • Olson, J. S. 1958. Rates of succession and soil changes on southern Lake Michigan sand dunes. Bot. Gaz.119: 125–170.

    Article  CAS  Google Scholar 

  • Oosting, H. J. 1944. The comparative effect of surface and crown fire of a loblolly pine community. Ecology25: 61–69.

    Article  Google Scholar 

  • Paine, R. T. 1974. Intertidal community structure. Experimental studies on the relationship between a dominant competitor and its principal predator. Oecologia15: 93–120.

    Article  Google Scholar 

  • Peek, J. M. 1974. Initial response of moose to a forest fire in northeastern Minnesota. Am. Midl. Nat.91: 435–438.

    Article  Google Scholar 

  • Pethick, J. S. 1974. The distribution of salt pans on tidal salt marshes. J. Biogeogr.1: 57–61.

    Article  Google Scholar 

  • Phillips, J. 1974. Effects of fire in forest and savanna ecosystems of Sub-Sahara Africa. Pages 435–481in T. T. Kozlowski and C. E. Ahlgren (eds.). Fire and Ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Philpot, C. W. 1977. Vegetative features as determinants of fire frequency and intensity. Pages 12–16in H. Mooney and C. E. Conrad (eds.). Proceedings of the symposium on the environmental consequences of fire and fuel management in Mediterranean ecosystems. U.S.D.A. For. Serv. Gen. Tech. Rept. WO-3. Washington, D.C.

  • Pickett, S. T. A. 1976. Succession: an evolutionary interpretation. Am. Nat.110: 107–119.

    Article  Google Scholar 

  • Place, I. C. M. 1964. Structure of old growth forest stands in eastern Canada. Abstract 10th Int. Bot. Congr. 1964, pp. 273–274.

  • Platt, W. J. 1975. The colonization and formation of equilibrium plant species associations on badger disturbances in a tall-grass prairie. Ecol. Monogr.45: 285–305.

    Article  Google Scholar 

  • — andI. M. Weis. 1977. Resource partitioning and competition within a guild of fugitive prairie plants. Am. Nat.111: 479–513.

    Article  Google Scholar 

  • Polunin, N. 1934. The vegetation of Akpatok Island, I. J. Ecol.22: 337–395.

    Article  Google Scholar 

  • —. 1935. The vegetation of Akpatok Island, II. J. Ecol.23: 161–209.

    Article  Google Scholar 

  • —. 1936. Plant succession in Norwegian Lapland. J. Ecol.24: 372–391.

    Article  Google Scholar 

  • Poore, M. E. D. 1968. Studies in Malaysian rain forest, I. The forest on triassic sediments in Jengka Forest Reserve. J. Ecol.56: 143–196.

    Article  Google Scholar 

  • Price, L. W. 1971. Vegetation, microtopography, and depth of active layer on different exposures in subarctic alpine tundra. Ecology52: 638–647.

    Article  Google Scholar 

  • Quarterman, E. andC. Keever. 1962. Southern mixed hardwood forest: Climax in the southeastern coastal plain, USA. Ecol. Monogr.32: 167–185.

    Article  Google Scholar 

  • Rabotnov, T. A. 1974. Differences between fluctuations and successions: examples in grassland phytocoenoses of the U.S.S.R. Pages 20–24in R. Knapp (ed.). Vegetation dynamics. Dr. W. Junk, The Hague.

    Google Scholar 

  • Raup, H. M. 1941a. Botanical problems in Boreal America. Bot. Rev.7: 147–248.

    Google Scholar 

  • —. 1941b. An old forest in Stonington, Connecticut. Rhodora43: 67–71.

    Google Scholar 

  • —. 1951. Vegetation and cryoplanation. Ohio J. Sci.51: 105–116.

    Google Scholar 

  • —. 1954. Some botanical problems of arctic and subarctic regions. Arctic7: 229–235.

    Google Scholar 

  • —. 1956.In report of meeting Int. Un. Conserv. Nature, Edinburgh, 1956. Nature178: 175–177.

    Article  Google Scholar 

  • —. 1957. Vegetational adjustment to the instability of site. Pages 36–48in Proc., 6th Tech. Meet. 1956, Int. Un. Conserv. Nature Nat. Resources, Edinburgh.

    Google Scholar 

  • —. 1964. Some problems in ecological theory and their relation to conservation. J. Ecol.52(suppl.): 19–28.

    Article  Google Scholar 

  • —. 1967. American forest biology. J. For.65: 800–803.

    Google Scholar 

  • —. 1975. Species versatility in shore habitats. J. Arnold Arbor. Harv. Univ.55: 126–165.

    Google Scholar 

  • Reiners, N. M. andW. A. Reiners. 1965. Natural harvesting of trees. William L. Hutcheson Mem. For. Bull.2: 9–17.

    Google Scholar 

  • Reiners, W. A. and G. E. Lang. 1979. Vegetational patterns and processes in the balsam fir zone, White Mountains, New Hampshire. Ecology (In press.)

  • Rhoades, R. W. 196I. The distribution of White Spruce,Picea glauca, in New Hampshire. M.S. Thesis, Univ. New Hampshire, Durham. 52 pp.

    Google Scholar 

  • Richards, P. W. 1952. The tropical rain forest. Cambridge, England. 450 pp.

  • —. 1955. The secondary succession in the tropical rain forest. Sci. Progr.43(169): 45–57.

    Google Scholar 

  • Richards, P. andG. B. Williamson. 1975. Treefalls and patterns of understory species in a wet lowland tropical forest. Ecology56: 1226–1229.

    Article  Google Scholar 

  • Ricklefs, R. E. 1977. Environmental heterogeneity and plant species diversity: a hypothesis. Am. Nat.111: 376–381.

    Article  Google Scholar 

  • Rogers, W. E. 1922. Ice storms and trees. Torreya22: 61–63.

    Google Scholar 

  • —. 1923. Resistance of trees to ice storm injury. Torreya23: 95–99.

    Google Scholar 

  • Rothermel, R. C. 1972. A mathematical model for predicting fire spread in midland fuels. USDA For. Serv. Res. Pap. INT-115. 40 pp.

  • Rowe, J. S. 1961. Critique of some vegetational concepts as applied to forests of north-western Alberta. Can. J. Bot.39: 1007–1017.

    Google Scholar 

  • — andG. W. Scotter. 1973. Fire in the boreal forest. Quat. Res.3: 444–464.

    Article  Google Scholar 

  • Rundel, P. W. 1973. The relationship between basal fire scars and crown damage in giant sequoia. Ecology54: 210–213.

    Article  Google Scholar 

  • Runkle, J. R. 1979. Gap phase dynamics in climax mesic forests. Ph.D. Thesis, Cornell Univ. Ithaca, New York. 289 pp.

    Google Scholar 

  • Sauer, J. D. 1962. Effects of recent tropical cyclones on the coastal vegetation of Mauritius. J. Ecol.50: 275–290.

    Article  Google Scholar 

  • Schroeder, P. M., R. Dolan, andB. P. Hayden. 1976. Vegetation changes associated with barrier-dune construction on the Outer Banks of North Carolina. Environ. Manage.1: 105–114.

    Article  Google Scholar 

  • Schwintzer, C. R. andG. Williams. 1974. Vegetation changes in a small Michigan bog from 1917 to 1972. Am. Midl. Nat.92: 447–459.

    Article  Google Scholar 

  • Scotter, G. W. 1972. Fire as an ecological factor in boreal forest ecosystems of Canada. Pages 15–24in Fire and the Environment, Symp. Proc. USFS. USDA. Denver, Colorado.

    Google Scholar 

  • Selleck, G. W. 1960. The climax concept. Bot. Rev.26: 534–545.

    Google Scholar 

  • Shafi, M. I. andG. A. Yarranton. 1973. Diversity,floristic richness, and species evenness during a secondary (post-fire) succession. Ecology54: 897–902.

    Article  Google Scholar 

  • Shelford, V. E. andS. Olson. 1935. Sere climax and influent animals with special reference to the transcontinental coniferous forest of North America. Ecology16: 375–402.

    Article  Google Scholar 

  • Sherman, R. J. andW. W. Chilcote. 1972. Spatial and chronological patterns ofPurshia tridentata as influenced byPinus ponderosa. Ecology72: 294–298.

    Article  Google Scholar 

  • Siccama, T. G., G. Weir, andK. Wallace. 1976. Ice damage in a mixed hardwood forest in Connecticut in relation toVitis infestation. Bull. Torrey Bot. Club103: 180–183.

    Article  Google Scholar 

  • Sigafoos, R. S. 1951. Soil instability in tundra vegetation. Ohio J. Sci.51: 281–298.

    Google Scholar 

  • —. 1952. Frost action as a primary physical factor in tundra plant communities. Ecology33: 480–487.

    Article  Google Scholar 

  • Silberbauer-Gottsberger, I., W. Morawetz, andG. Gottsberger. 1977. Frost damage of Cerrado plants in Botucatu, Brazil, as related to the geographical distribution of species. Biotropica9: 253–261.

    Article  Google Scholar 

  • Skunk, I. V. 1939. Oxygen requirements for germination of seeds ofNyssa aquatica, tupelo gum. Science90: 565–566.

    Article  Google Scholar 

  • Slatkin, M. andR. Lange. 1976. Niche width in a fluctuating environment—density independent model. Am. Nat.110: 31–55.

    Article  Google Scholar 

  • Slaughter, C. W., R. J. Barney, and G. M. Hansen (eds.). 1971. Fire in the northern environment—a symposium. Pac. Northwest. For. Range Exp. Stn., For. Serv., USDA, Portland, Oregon. 275 pp.

  • Smathers, G. A. and D. Mueller-Dombois. 1974. Invasion and recovery of vegetation after a volcanic eruption in Hawaii. National Park Service, Sci. Monogr. Series No. 5.

  • Smith, D. H. 1946. Storm damage in New England forests. M.S. Thesis, Yale Univ., New Haven, Connecticut. 173 pp.

    Google Scholar 

  • Snow, D. W. 1965. A possible selective factor in the evolution of fruiting seasons in a tropical forest. Oikos15: 274–281.

    Article  Google Scholar 

  • Sobey, D. G. andP. Barkhouse. 1977. The structure and rate of growth of the rhizomes of some forest herbs and dwarf shrubs of the New Brunswick-Nova Scotia border region. Can. Field-Nat.91: 377–383.

    Google Scholar 

  • Sollers, S. C. 1974. Substrate conditions, community structure, and succession in a portion of the flood plain of Wissahickon Creek. Bartonia42: 24–42.

    Google Scholar 

  • Sprugel, D. G. 1976. Dynamic structure of wave-generatedAbies balsamea forests in the northeastern United States. J. Ecol.64: 889–912.

    Article  Google Scholar 

  • Spurr, S. H. 1956a. Natural restocking of forests following the 1938 hurricane in central New England. Ecology37: 443–451.

    Article  Google Scholar 

  • —. 1956b. Forest associations in the Harvard Forest. Ecol. Monogr.26: 245–262.

    Article  Google Scholar 

  • Starker, J. T. 1934. Fire resistance in the forest. J. For.32: 462–467.

    Google Scholar 

  • Stearns, F. W. 1949. Ninety years change in a northern hardwood forest in Wisconsin. Ecology30: 350–358.

    Article  Google Scholar 

  • Stelfox, J. G. andH. G. Vriend. 1977. Prairie fires and pronghorn use of cactus. Can. Field-Nat.91: 282–285.

    Google Scholar 

  • Stephens, E. P. 1955a. The historical-developmental method of determining forest trends. Ph.D. Thesis, Harvard Univ., Cambridge. 36 pp.

    Google Scholar 

  • —. 1955b. Research in the biological aspects of forest production. J. For.53: 183–186.

    Google Scholar 

  • —. 1956. The uprooting of trees: a forest process. Proc. Soil Sci. Soc. Am.20: 113–116.

    Article  Google Scholar 

  • Stiles, E. H. andL. E. Melchers. 1935. The drought of 1934 and its effect on trees in Kansas. Trans. Kans. Acad. Sci.38: 107–127.

    Article  Google Scholar 

  • Stone, E. C. andR. B. Vasey. 1968. Preservation of coast redwood on alluvial flats. Science159: 157–161.

    Article  PubMed  CAS  Google Scholar 

  • Stoneburner, D. L. 1978. Evidence of hurricance influence on Barner Island slash pine forests in the Northern Gulf of Mexico. Am. Midl. Nat.99: 234–238.

    Article  Google Scholar 

  • Stout, B. B., J. M. Deschenes, andL. F. Ohmann. 1975. Multispecies model of a deciduous forest. Ecology56: 226–331.

    Article  Google Scholar 

  • Strong, D. R., Jr. 1977. Epiphyte loads, tree-falls, and perennial disruption: a mechanism for maintaining higher tree species richness in the tropics without animals. J. Biogeogr.4: 215–218.

    Article  Google Scholar 

  • Sutherland, J. P. 1974. Multiple stable points in natural communities. Am. Nat.108: 859–873.

    Article  Google Scholar 

  • Swaine, J. M. 1933. The relation of insect activities to forest development as exemplified in the forests of eastern North America. Sci. Agric.14: 8–31.

    Google Scholar 

  • Swanston, D. N. andF. J. Swanson. 1976. Timber-harvesting, mass erosion, and steep land form geomorphology in the Pacific Northwest. Pages 199–221in D. R. Coats (ed.). Geomorphology and Engineering, Dowden, Hutchinson, and Ross, Inc., Stroudsburg, Pennsylvania.

    Google Scholar 

  • Tansley, A. G. 1935. The use and abuse of vegetational concepts and terms. Ecology16: 284–308.

    Article  Google Scholar 

  • Taylor, B. W. 1963. An outline of the vegetation of Nicaragua. J. Ecol.51: 27–54.

    Article  Google Scholar 

  • Taylor, D. L. 1973. Some ecological implications of forest fire control in Yellowstone National Park, Wyoming. Ecology51: 1394–1396.

    Article  Google Scholar 

  • Thom, B. G. 1967. Mangrove ecology and deltaic geomorphology: Tabasco, Mexico. J. Ecol.55: 301–343.

    Article  Google Scholar 

  • Transeau, E. N. 1935. The prairie peninsula. Ecology16: 423–437.

    Article  Google Scholar 

  • Tredici, P. del. 1977. The buried seeds ofComptonia peregrina, the sweet fern. Bull. Torrey Bot. Club104: 270–275.

    Article  Google Scholar 

  • Trimble, G. R., Jr. and D. W. Seegrist. 1973. Epicormic branching on hardwood trees boardering forest openings. U.S.D.A. For. Serv. Res. Pap. NE-261. 6 pp.

  • Turner, L. M. 1935. Catastrophes and pure stands of southern shortleaf pine. Ecology16: 213–215.

    Article  Google Scholar 

  • U.S.D.A. 1965. Silvics of forest trees of the United States. For. Serv., USDA Agric. Handb.271. 762 pp.

  • Van der Valk, A. G. andC. B. Davis. 1978. The role of seed banks in the vegetation dynamics of prairie glacial marshes. Ecology59: 322–335.

    Article  Google Scholar 

  • Vankat, J. L., W. H. Blackwell, Jr., andW. E. Hopkins. 1975. The dynamics of Hueston Woods and a review of the question of the successional status of the southern beechmaple forest. Castanea40: 290–308.

    Google Scholar 

  • Van Steenis, G. G. G. J. 1956. Basic principles of rain forest sociology. Pages 159–163in Study of tropical vegetation. Proc. of the Kandy Symp., UNESCO, Paris.

    Google Scholar 

  • Veblen, T. T., D. H. Ashton, F. M. Schlegel, andA. T. Veblen. 1977. Plant succession in a timberline depressed by vulcanism in south-central Chile. J. Biogeogr.4: 275–294.

    Article  Google Scholar 

  • Veno, P. A. 1976. Successional relationships of five Florida plant communities. Ecology57: 498–508.

    Article  Google Scholar 

  • Viereck, L. A. 1966. Plant succession and soil development on gravel outwash of the Muldrow Glacier, Alaska. Ecol. Monogr.36: 181–199.

    Article  Google Scholar 

  • —. 1973. Wildfire in the taiga of Alaska. Quat. Res.3: 465–495.

    Article  Google Scholar 

  • Visher, S. S. 1949. American dry seasons: their intensity and frequency. Ecology30: 365–370.

    Article  Google Scholar 

  • Vogl, R. J. 1974. Effects of fire on grasslands. Pages 139–194in T. T. Kozlowski and C. E. Ahlgren (eds.). Fire and ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Wace, N. M. 1961. The vegetation of Gough Island. Ecol. Monogr.31: 337–367.

    Article  Google Scholar 

  • Wadsworth, F. H. andG. H. Englerth. 1959. Effects of the 1956 hurricane on forests in Puerto Rico. Caribb. For.20: 38–51.

    Google Scholar 

  • Wagner, R. H. 1964. The ecology ofUniola paniculata L. in the dune-strand habitat of North Carolina. Ecol. Monogr.34: 79–96.

    Article  Google Scholar 

  • Walker, B. H. 1972. Vegetation-site relationships in the Harvard Forest. Vegetatio29: 169–178.

    Article  Google Scholar 

  • Ware, G. H. andW. T. Penfound. 1949. The vegetation of the lower levels of the flood plain of the South Canadian River in central Oklahoma. Ecology30: 478–484.

    Article  Google Scholar 

  • Watt, A. S. 1923. On the ecology of British beechwoods with special reference to their regeneration. J. Ecol.11: 1–48.

    Article  Google Scholar 

  • —. 1947a. Pattern and process in the plant community. J. Ecol.35: 1–22.

    Article  Google Scholar 

  • —. 1947b. Contributions to the ecology of bracken. IV. The structure of the community. New Phytol.46: 97–121.

    Article  Google Scholar 

  • —. 1955. Bracken versus heather: a study in plant sociology. J. Ecol.43: 490–506.

    Article  Google Scholar 

  • —. 1974. Senescence and rejuvenation in ungrazed chalk grassland (grassland B) in Breckland: The significance of litter and of moles. J. Appl. Ecol.11: 1157–1172.

    Article  Google Scholar 

  • —,R. M. S. Perrin, andR. G. West. 1966. Patterned ground in Breckland: structure and composition. J. Ecol.64: 239–258.

    Google Scholar 

  • Weaver, H. 1974. Effects of fire on temperate forests: Western United States. Pages 279–319in T. T. Kozlowski and C. E. Ahlgren (eds.). Fire and ecosystems. Academic Press, New York. 542 pp.

    Google Scholar 

  • Weaver, J. E. 1950. Stabilization in midwestern grassland. Ecol. Monogr.20: 251–270.

    Article  Google Scholar 

  • —. 1958. Summary and interpretation of underground development in natural grassland communities. Ecol. Monogr.28: 55–78.

    Article  Google Scholar 

  • —. 1960. Flood plain vegetation of the central Missouri Valley and contacts of woodland with prairie. Ecol. Monogr.30: 37–64.

    Article  Google Scholar 

  • —. 1968. Prairie plants and their environment: a 50-year study in the Midwest. Univ. Nebr. Press, Lincoln. 276 pp.

    Google Scholar 

  • — andF. W. Albertson. 1943. Resurvey of grasses, forbs, and underground plant parts at the end of the great drought. Ecol. Monogr.13: 63–117.

    Article  Google Scholar 

  • ——. 1944. Nature and degree of recovery of grassland from the great drought of 1933 to 1940. Ecol. Monogr.14: 393–479.

    Article  Google Scholar 

  • — andW. E. Bruner. 1945. A seven-year quantitative study of succession in grassland. Ecol. Monogr.15: 297–319.

    Article  Google Scholar 

  • — andF. E. Clements. 1929. Plant ecology, 2nd ed., 1938. McGraw-Hill, New York. 520 pp.

    Google Scholar 

  • —,L. A. Stoddart, andW. Noll. 1935. Response of the prairie to the Great Drought of 1934. Ecology16: 612–629.

    Article  Google Scholar 

  • Webb, L. J. 1958. Cyclones as an ecological factor in tropical lowland rain forest, north Queensland. Aust. J. Bot.6: 220–228.

    Article  Google Scholar 

  • —. 1968. Environmental relationships of the structural types of Australian rain forest vegetation. Ecology49: 296–311.

    Article  Google Scholar 

  • —,L. G. Tracey, andW. T. Williams. 1972. Regeneration and pattern in the subtropical rain forest. J. Ecol.60: 675–695.

    Article  Google Scholar 

  • Wein, R. W. andL. C. Bliss. 1973. Changes in arcticEriophorum tussock communities following fire. Ecology54: 845–852.

    Article  Google Scholar 

  • Wells, B. W. 1942. Ecological problems of the southeastern United States coastal plain. Bot. Rev.8: 533–561.

    Article  Google Scholar 

  • Wells, H. W. 1961. The fauna of oyster beds, with special reference to the salinity factor. Ecol. Monogr.31: 239–266.

    Article  Google Scholar 

  • Wells, P. V. 1965. Scarp woodlands, transported grassland soils, and concept of grassland climate in the Great Plains region. Science148: 246–249.

    Article  PubMed  CAS  Google Scholar 

  • Wendel, G. W. 1972. Longevity of black cherry seed in the forest floor. U.S.D.A. For. Serv. Res. Note NE-149. 4 pp.

  • West, O. 1971. Fire, man, and wildlife as interacting factors limiting the development of climax vegetation in Rhodesia. Proc. Annu. Tall Timbers Fire Ecol. Conf.11: 121–145.

    Google Scholar 

  • Westman, W. E. 1968. Invasion of fir forest by sugar maple in Itasca Park, Minnesota. Bull. Torrey Bot. Club95: 172–186.

    Article  Google Scholar 

  • White, P. S. 1976. The upland forest vegetation of the Second College Grant, New Hampshire. Ph.D. Thesis, Dartmouth College, Hanover, New Hampshire. 295 pp.

    Google Scholar 

  • Whitmore, T. C. 1974. Change with time and the role of cyclones in tropical rain forest of Kolombangara, Solomon Islands, Univ. Oxford Commonwealth Forestry Institute, Institute Paper No. 46.

  • —. 1975. Tropical rain forests of the Far East. Clarendon Press, Oxford. 282 pp.

    Google Scholar 

  • Whitney, G. G. 1976. The bifurcation ratio as an indicator of adaptive strategy in woody plant species. Bull. Torrey Bot. Club103: 67–72.

    Article  Google Scholar 

  • Whittaker, R. H. 1953. A consideraton of climax theory: the climax as a population and pattern. Ecol. Monogr.23: 41–78.

    Article  Google Scholar 

  • —. 1972. Evolution and measurement of species diversity. Taxon21: 213–251.

    Article  Google Scholar 

  • —. 1974. Climax concepts and recognition. Pages 138–154in R. Knapp (ed.). Vegetation dynamics. Dr. W. Junk, The Hague.

    Google Scholar 

  • —. 1975. Communities and ecosystems, 2nd ed., Macmillan. New York. 385 pp.

    Google Scholar 

  • —. 1977. Evolution of species diversity in land communities. Evol. Biol.10: 1–67.

    Google Scholar 

  • — andD. Goodman. 1979. Classifying species according to their demographic strategy. I. Population fluctuations and environmental heterogeneity. Am. Nat.113: 185–200.

    Article  Google Scholar 

  • Wiens, J. A. 1976. Population responses to patchy environments. Ann. Rev. Ecol. Syst.7: 81–120.

    Article  Google Scholar 

  • Williamson, G. G. 1975. Pattern and seral composition in an old-growth beech-maple forest. Ecology56: 727–731.

    Article  Google Scholar 

  • Wilson, J. W. 1952. Vegetation patterns associated with soil movement on Jan Mayen Island. J. Ecol.40: 249–264.

    Article  Google Scholar 

  • Wistendahl, W. A. 1958. The Hood plain of the Raritan River, New Jersey. Ecol. Monogr.28: 129–153.

    Article  Google Scholar 

  • Woodin, S. A. 1978. Refuges, disturbance, and community structure: a marine soft-bottom example. Ecology59: 274–284.

    Article  Google Scholar 

  • Wright, H. E., Jr. 1974. Landscape development, forest fires, and wilderness management. Science186: 487–495.

    Article  PubMed  Google Scholar 

  • — andM. L. Heinselman. 1973. The ecological role of fire in natural conifer forests of western and northern North America—Introduction. Quat. Res.3: 319–328.

    Article  Google Scholar 

  • Yarranton, M. andG. A. Yarranton. 1975. Demography of a jack pine stand. Can. J. Bot.53: 310–314.

    Article  Google Scholar 

  • Zach, L. W. 1950. A northern climax, forest or muskey? Ecology31: 304–306.

    Article  Google Scholar 

  • Zackrisson, O. 1977. Influence of forest fires on the north Swedish boreal forest. Oikos29: 22–32.

    Article  Google Scholar 

  • Zavitkovski, J. andM. Newton. 1968. Ecological improtance of snowbrush (Ceanothus velutinus) in the Oregon Cascades. Ecology49: 1134–1145.

    Article  Google Scholar 

  • Zedler, P. H. andT. A. Ebert. 1977. Shrub seedling establishment and survival following an unusual September rain in the Colorado desert. Abstract in Bull. Ecol. Soc. Am.58(2): 47.

    Google Scholar 

  • — andF. G. Goff. 1973. Size-association analysis of forest successional trends in Wisconsin. Ecol. Monogr.43: 79–94.

    Article  Google Scholar 

  • Zobel, D. B. 1969. Factors affecting the distribution ofPinus pungens, an Appalachian endemic. Ecol. Monogr.39: 303–333.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

About this article

Cite this article

White, P.S. Pattern, process, and natural disturbance in vegetation. Bot. Rev 45, 229–299 (1979). https://doi.org/10.1007/BF02860857

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1007/BF02860857

Keywords

Navigation